Physics 112 The Classical Ideal Gas

Similar documents
Grand Canonical Formalism

We already came across a form of indistinguishably in the canonical partition function: V N Q =

Ex3009: Entropy and heat capacity of quantum ideal gases

i=1 n i, the canonical probabilities of the micro-states [ βǫ i=1 e βǫn 1 n 1 =0 +Nk B T Nǫ 1 + e ǫ/(k BT), (IV.75) E = F + TS =

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

The properties of an ideal Fermi gas are strongly determined by the Pauli principle. We shall consider the limit:

Advanced Thermodynamics. Jussi Eloranta (Updated: January 22, 2018)

5. Systems in contact with a thermal bath

Introduction Statistical Thermodynamics. Monday, January 6, 14

[S R (U 0 ɛ 1 ) S R (U 0 ɛ 2 ]. (0.1) k B

HWK#7solution sets Chem 544, fall 2014/12/4

Monatomic ideal gas: partition functions and equation of state.

PHYS 352 Homework 2 Solutions

summary of statistical physics

VII.B Canonical Formulation

Chapter 14. Ideal Bose gas Equation of state

Quantum ideal gases: bosons

Ideal gases. Asaf Pe er Classical ideal gas

5. Systems in contact with a thermal bath

Many-Particle Systems

Physics 505 Homework No.2 Solution

UNIVERSITY OF SOUTHAMPTON

E = 0, spin = 0 E = B, spin = 1 E =, spin = 0 E = +B, spin = 1,

13. Ideal Quantum Gases I: Bosons

Lecture 6: Ideal gas ensembles

( ) ( )( k B ( ) ( ) ( ) ( ) ( ) ( k T B ) 2 = ε F. ( ) π 2. ( ) 1+ π 2. ( k T B ) 2 = 2 3 Nε 1+ π 2. ( ) = ε /( ε 0 ).

INDIAN INSTITUTE OF TECHNOLOGY GUWAHATI Department of Physics MID SEMESTER EXAMINATION Statistical Mechanics: PH704 Solution

2. Thermodynamics. Introduction. Understanding Molecular Simulation

9.1 System in contact with a heat reservoir

4. Systems in contact with a thermal bath

Thermal and Statistical Physics Department Exam Last updated November 4, L π

(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 607 Final Exam

Quantum Grand Canonical Ensemble

The state of a quantum ideal gas is uniquely specified by the occupancy of singleparticle

Solution to the exam in TFY4230 STATISTICAL PHYSICS Wednesday december 21, 2011

Part II Statistical Physics

CHEM-UA 652: Thermodynamics and Kinetics

Part II: Statistical Physics

The Second Virial Coefficient & van der Waals Equation

Recitation: 10 11/06/03

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term Solutions to Problem Set #10

+ 1. which gives the expected number of Fermions in energy state ɛ. The expected number of Fermions in energy range ɛ to ɛ + dɛ is then dn = n s g s

ChE 210B: Advanced Topics in Equilibrium Statistical Mechanics

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

PHYS 328 HOMEWORK 10-- SOLUTIONS

Phys Midterm. March 17

Assignment 8. Tyler Shendruk December 7, 2010

2m + U( q i), (IV.26) i=1

Grand-canonical ensembles

CHAPTER 4. Cluster expansions

I. BASICS OF STATISTICAL MECHANICS AND QUANTUM MECHANICS

Statistical thermodynamics L1-L3. Lectures 11, 12, 13 of CY101

Physics 607 Final Exam

CHEM-UA 652: Thermodynamics and Kinetics

14. Ideal Quantum Gases II: Fermions

Supplement: Statistical Physics

Physics 607 Exam 2. ( ) = 1, Γ( z +1) = zγ( z) x n e x2 dx = 1. e x2

The non-interacting Bose gas

Thermal & Statistical Physics Study Questions for the Spring 2018 Department Exam December 6, 2017

Second Quantization: Quantum Fields

A Brief Introduction to Statistical Mechanics

Solutions to Problem Set 8

Statistical Mechanics Notes. Ryan D. Reece

Noninteracting Particle Systems

System-Size Dependence in Grand Canonical and Canonical Ensembles

Statistical. mechanics

liquid He

1 Fundamentals of Statistical Physics

Physics 408 Final Exam

ChE 503 A. Z. Panagiotopoulos 1

V.E Mean Field Theory of Condensation

Quantum statistics: properties of the Fermi-Dirac distribution.

Handout 10. Applications to Solids

PHY 5524: Statistical Mechanics, Spring February 11 th, 2013 Midterm Exam # 1

Physics Oct Reading. K&K chapter 6 and the first half of chapter 7 (the Fermi gas). The Ideal Gas Again

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

1 The fundamental equation of equilibrium statistical mechanics. 3 General overview on the method of ensembles 10

( ) Lectures on Hydrodynamics ( ) =Λ Λ = T x T R TR. Jean-Yve. We can always diagonalize point by point. by a Lorentz transformation

Thermodynamics & Statistical Mechanics

Exam TFY4230 Statistical Physics kl Wednesday 01. June 2016

Definite Integral and the Gibbs Paradox

Microcanonical Ensemble

Outline Review Example Problem 1 Example Problem 2. Thermodynamics. Review and Example Problems. X Bai. SDSMT, Physics. Fall 2013

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

8.333: Statistical Mechanics I Re: 2007 Final Exam. Review Problems

Physics 127b: Statistical Mechanics. Lecture 2: Dense Gas and the Liquid State. Mayer Cluster Expansion

V.C The Second Virial Coefficient & van der Waals Equation

PHYSICS 210A : EQUILIBRIUM STATISTICAL PHYSICS HW ASSIGNMENT #4 SOLUTIONS

Many-body physics 2: Homework 8

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

Outline Review Example Problem 1. Thermodynamics. Review and Example Problems: Part-2. X Bai. SDSMT, Physics. Fall 2014

The perfect quantal gas

Statistical Mechanics

84 Quantum Theory of Many-Particle Systems ics [Landau and Lifshitz (198)] then yields the thermodynamic potential so that one can rewrite the statist

Quantum ideal gases: fermions

Solution to the exam in TFY4230 STATISTICAL PHYSICS Friday december 19, 2014

1. Thermodynamics 1.1. A macroscopic view of matter

8.333: Statistical Mechanics I Re: 2007 Final Exam. Review Problems

Introduction. Chapter The Purpose of Statistical Mechanics

Transcription:

Physics 112 The Classical Ideal Gas Peter Young (Dated: February 6, 2012) We will obtain the equation of state and other properties, such as energy and entropy, of the classical ideal gas. We will start with quantum statistical mechanics, and take the classical limit, since this avoids certain ambiguities. Furthermore, our result for the entropy will involve h and so what we are doing is not entirely a classical theory. A state of an ideal gas is identified by the occupation numbers, n k, of the single particle states k with energy ǫ k. In particular, the total number of particles N and energy E are given by N = k E = k n k, (1) n k ǫ k. (2) We will use the Gibbs distribution, in which the number of particles is allowed to vary and, the results are expressed in terms of the chemical potential µ rather than the mean number of particles N. The reason for this choice is that, as we shall now show, for non-interacting, identical particles, the grand partition of the whole system, Z, factorizes into a a product of grand partition functions for the single-particle states. The grand partition function is given by Z = exp[β(nµ E)] = [ β {n k } {n k }exp ] (µ ǫ k )n k = ] [e β(µ ǫ k)n k = k {n k } k k [ ], n k e β(µ ǫ k)n k where µ is the chemical potential, and {n k } refers to the set of all allowed values of the occupation numbers n k. To get the third expression we used that the exponential of a sum is the product of exponentials, and to get the fourth expression we used that each exponential only depends on one of the n k. The summations over the different n k can be done independently because there is no constraint on the total number N = k n k in the Gibbs distribution. Hence, as stated above, the grand partition of the whole system, Z, factorizes into a a product of grand partition functions for the single-particle states, i.e. (3) Z = k Zk, (4)

2 where Zk = n k e n kβ(µ ǫ k ). (5) The factorization of the grand partition function for non-interacting particles is the reason why we use the Gibbs distribution (also known as the grand canonical ensemble ) for quantum, indistinguishable particles. For fermions, n k in the sum in Eq. (5) only takes values 0 and 1, while for bosons n k takes values from 0 to and Eq. (5) gives a geometric series which is easy to sum. The result is Zk = 1+λe βǫ k, fermions, ( ) 1 λe βǫ k 1 (6), bosons, where λ = e βµ (7) is called the activity (sometimes the word fugacity is also used). As discussed in class, the classical limit is when the occupancy of any single-particle state is very much less that unity. This means that the statistical weight (i.e. the exponential factor in Eq. (5)) for n k = 1, is much less than the statistical weight for n k = 0, i.e. λe βǫ k 1. (8) In this case, we work to first order in λe βǫ k, so the difference between bosons and fermions disappears and for both we get Zk = 1+λe βǫ k, (9) lnzk = λe βǫ k. (10) where we used ln(1+x) = x+ for small x. Taking the log of Eq. (4), and using Eq. (10), gives lnz = k lnzk = λ k e βǫ k. (11) As shown in class, the grand potential Ω is related to Z by Ω (= PV) = k B T lnz, (12) where P is the pressure, and combining this with Eq. (11) gives Ω = k B Tλz (1) (13)

3 where z (1) is the (canonical) partition function for a single particle: z (1) = k e βǫ k. (14) To carry out the sum in Eq. (14) we use the result for the single particle density of states discussed earlier in class, and covert the sum over states to an integral over ǫ. Considering spin-0 (in the class we will discuss the trivial changes that come from a non-zero spin) the density of states is given by Hence z (1) in Eq. (14) is given by z (1) = V 4π 2 ρ(ǫ) = V 4π 2 ( ) 2m 3/2 h 2 ǫ 1/2. (15) ( ) 2m 3/2 h 2 ǫ 1/2 e βǫ dǫ = V ( ) 2mkB T 3/2 0 4π 2 h 2 x 1/2 e x dx. (16) 0 As discussed in Math. Methods classes, the integral is Γ(3/2) = π/2, and so z (1) = V ( mkb T 2π h 2 ) 3/2 = V V Q, (17) where ( 2π h 2 ) 3/2 V Q = (18) mk B T is the quantum volume already discussed in class. Hence, from Eqs. (11), (14) and (17) lnz = λ V V Q, (19) so Z = e λv/v Q. (20) Using Eq. (12), then gives the desired expression for the grand potential Ω Ω(T,V,µ) = k B Tλ V V Q, (21) (remember λ = e βµ ). Since we are using the Gibbs distribution with a variable number of particles, we have calculated the grand potential, which is expressed in terms of the chemical potential rather than the number of particles. However, in practice, experiments are carried out with a given number of particles, not a given value of the chemical potential.

4 Hence we have to convert expressions like Eq. (21), written in terms of µ, to ones written in terms of the mean number of particles N. This is a general problem with the Gibbs distribution. Fortunately this conversion is very easy here for the classical ideal gas because we will find a closed form expression for µ(t) as a function of n, Eq. (24) below. It will be less easy when we consider quantum ideal gases. To obtain the result for µ(t), recall that the mean number of particles is given by ( ) Ω N =, (22) µ which, according to Eq. (21), yields T,V N = e βµ V V Q, (23) (remember that λ = e βµ ). Although fluctuations in the number of particles are allowed in the Gibbs distribution, their relative size is small (of order N 1/2 ). Hence, for conciseness of notation, from now on, we will omit the brackets and denote the mean number of particles simply by N. Eq. (23) gives the desired expression for the chemical potential: [ ( ) ] 1 mkb T 3/2 µ(t) = k B T ln(nv Q ) = k B T ln n 2π h 2 = 3 ( ) T 2 k BT ln T Q (24) where n = N/V is the particle density, and T Q = 2π h2 mk B n 2/3, is a temperature below which quantum effects are important. Since the the condition to be in the classical regime is nv Q 1, see Eq. (25) below, (or equivalently T T Q ), it follows that µ must be negative. Furthermore, Eq. (24) shows that µ gets more negative with increasing T. The variation of µ(t) with T in the classical regime is sketched in the figure below. µ (T) 0 T Q T Solid line Classical result (T >> T Q) At lower T (dashed line and still lower T) µ depends on the statistics of the particles (fermions or bosons)

5 To describe a classical ideal gas with a FIXED density of particles n using the Gibbs distribution, we have to let µ(t) vary with T according to Eq. (24). The condition that we are in the classical regime, i.e. the occupancy is much less than unity even for the lowest single particle level which we take to be at zero energy, is λ 1, see Eq. (8). From Eq. (23), this is equivalent to nv Q 1 (condition to be in the classical regime). (25) As discussed in class this corresponds to the mean separation between particles (n 1/3 ) being much greater than the thermal de Broglie wavelength (the de Broglie wavelength of a particle whose energy is k B T). Equivalently, from Eq. (24), we can write T T Q (condition to be in the classical regime), (26) showing that classical regime is a high temperature regime. Eq. (23) can be substituted into Eq. (21), and noting that Ω = PV, we get PV = Nk B T, (27) the famous ideal gas law. The (Helmholtz) free energy is F = Ω+µN which, using Eqs. (21) and (23) and recalling that F should be expressed in terms of N not µ, is given by F = k B TN +µn or F(T,V,N) = Nk B T [ln(nv Q ) 1]. (28) Now that we have determined F, it is more convenient to use it, rather than Ω, to determine other quantities, because F is expressed in terms of N, which is specified and kept fixed, whereas Ω is expressed in terms of µ, which varies with T and V (in such a way as to keep N constant). For example, the entropy is given by ( ) [ F S = = Nk B 1 ln(nv Q ) T ] T V,N T lnv Q. (29) Since V Q T 3/2, ( / T)lnV Q = 3/(2T). Hence we have [ ] 5 S = Nk B 2 ln(nv Q), (30)

6 a well known result, which is called the Sackur-Tetrode equation. The entropy is proportional to the number of particles N if the density n is kept constant, i.e. the entropy is extensive. Note too that although we have taken the classical limit, the expression for the entropy involves h. The reason is that only in quantum mechanics can one define precisely the sum over states and hence give an absolute definition for the entropy (rather than just entropy differences). More discussion of this, along with a discussion of experimental verification of Eq. (30) will be given in class. The energy can be obtained from U = ( ) (βf) The only quantity which depends on β is V Q, and V Q β 3/2, so β N,V = N β [ln(nv Q) 1] N,V. (31) U = 3 2 Nk BT. (32) This is an example of a well known result of classical statistical mechanics, called the equipartition theorem. This states that every quadratic term in the microscopic expression for the energy contributes (1/2)k B T to the average energy. Here we have N atoms for each of which the energy is 1 2 m(v2 x +v 2 y +v 2 z), (33) which has 3 quadratic terms. Hence the total average energy is (3/2)Nk B T, in agreement with Eq. (32) The specific heat at constant volume, C V, can be obtained from C V = ( U/ T) N,V or equivalently from C V = (T S/ T) N,V, and the result is C V = 3 2 Nk B. (34)