I. BASICS OF STATISTICAL MECHANICS AND QUANTUM MECHANICS

Similar documents
I. BASICS OF STATISTICAL MECHANICS AND QUANTUM MECHANICS

Quantum Grand Canonical Ensemble

Monatomic ideal gas: partition functions and equation of state.

V. FERMI LIQUID THEORY

1 Fluctuations of the number of particles in a Bose-Einstein condensate

A. Field theoretical methods for the partition function, perturbative expansions

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

N independent electrons in a volume V (assuming periodic boundary conditions) I] The system 3 V = ( ) k ( ) i k k k 1/2

Physics 127a: Class Notes

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

d 3 r d 3 vf( r, v) = N (2) = CV C = n where n N/V is the total number of molecules per unit volume. Hence e βmv2 /2 d 3 rd 3 v (5)

First Problem Set for Physics 847 (Statistical Physics II)

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

Chapter 14. Ideal Bose gas Equation of state

(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

5. Systems in contact with a thermal bath

Second Quantization: Quantum Fields

Introduction Statistical Thermodynamics. Monday, January 6, 14

1 Quantum field theory and Green s function

summary of statistical physics

Statistical Mechanics

QUANTUM FIELD THEORY, GREEN S FUNCTION, AND PERTURBATION THEORY

VII.B Canonical Formulation

Introduction to Path Integral Monte Carlo. Part I.

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

Introduction. Chapter The Purpose of Statistical Mechanics

Elements of Statistical Mechanics

LECTURES ON STATISTICAL MECHANICS E. MANOUSAKIS

5. Systems in contact with a thermal bath

Many-body physics 2: Homework 8

PRINCIPLES OF PHYSICS. \Hp. Ni Jun TSINGHUA. Physics. From Quantum Field Theory. to Classical Mechanics. World Scientific. Vol.2. Report and Review in

The non-interacting Bose gas

Physics 505 Homework No.2 Solution

Electrons in a periodic potential

International Physics Course Entrance Examination Questions

Quantum ideal gases: bosons

The general solution of Schrödinger equation in three dimensions (if V does not depend on time) are solutions of time-independent Schrödinger equation

Part II: Statistical Physics

PHYS3113, 3d year Statistical Mechanics Tutorial problems. Tutorial 1, Microcanonical, Canonical and Grand Canonical Distributions

Lecture 6: Ideal gas ensembles

We start with some important background material in classical and quantum mechanics.

Assignment 8. Tyler Shendruk December 7, 2010

Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

2 Statistical Mechanics of Non-Interacting Particles

Quantum statistics: properties of the Fermi-Dirac distribution.

9.1 System in contact with a heat reservoir

Quantum Mechanics: Fundamentals

Solid Thermodynamics (1)

Grand Canonical Formalism

Physics 127c: Statistical Mechanics. Application of Path Integrals to Superfluidity in He 4

Identical Particles. Bosons and Fermions

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

The Ideal Gas. One particle in a box:

-state problems and an application to the free particle

Quantum field theory and Green s function

Statistical physics and light-front quantization. JR and S.J. Brodsky, Phys. Rev. D70, (2004) and hep-th/

Part II Statistical Physics

We can then linearize the Heisenberg equation for in the small quantity obtaining a set of linear coupled equations for and :

Coordinate and Momentum Representation. Commuting Observables and Simultaneous Measurements. January 30, 2012

221B Lecture Notes Quantum Field Theory II (Fermi Systems)

The Partition Function Statistical Thermodynamics. NC State University

Bose-Einstein Condensation

Grand-canonical ensembles

CHAPTER 9 Statistical Physics

Fluctuations of Trapped Particles

6.730 Physics for Solid State Applications

Physics 112 The Classical Ideal Gas

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

1.1 Creation and Annihilation Operators in Quantum Mechanics

561 F 2005 Lecture 14 1

I. QUANTUM MONTE CARLO METHODS: INTRODUCTION AND BASICS

Ex3009: Entropy and heat capacity of quantum ideal gases

ChE 503 A. Z. Panagiotopoulos 1

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

Attempts at relativistic QM

Infinitely long-range nonlocal potentials and the Bose-Einstein supersolid phase

Elementary Lectures in Statistical Mechanics

Imperial College London BSc/MSci EXAMINATION May 2008 THERMODYNAMICS & STATISTICAL PHYSICS

liquid He

221B Lecture Notes Quantum Field Theory II (Fermi Systems)

Advanced Topics in Equilibrium Statistical Mechanics

Low-dimensional Bose gases Part 1: BEC and interactions

Lecture 6: Fluctuation-Dissipation theorem and introduction to systems of interest

Statistical Methods and Thermodynamics - Spring Prerequisites: CHEM 130 or 330 and or 332, or equivalents.

Quantum statistics of ideal gases in confined space

Second quantization (the occupation-number representation)

Lecture 4: Entropy. Chapter I. Basic Principles of Stat Mechanics. A.G. Petukhov, PHYS 743. September 7, 2017

SECOND QUANTIZATION. Lecture notes with course Quantum Theory

+1-1 R

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Bose-Einstein condensation: static and dynamical aspects

UNIVERSITY OF MARYLAND Department of Physics College Park, Maryland. PHYSICS Ph.D. QUALIFYING EXAMINATION PART II

A Brief Introduction to Statistical Mechanics

Formulation of Quantum Statistics

Noninteracting Particle Systems

30 Photons and internal motions

Many-Body Problems and Quantum Field Theory

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

Superfluidity and Superconductivity

Lecture 6 Photons, electrons and other quanta. EECS Winter 2006 Nanophotonics and Nano-scale Fabrication P.C.Ku

Transcription:

I. BASICS OF STATISTICAL MECHANICS AND QUANTUM MECHANICS Marus Holzmann LPMMC, Maison de Magistère, Grenoble, and LPTMC, Jussieu, Paris marus@lptl.jussieu.fr http://www.lptl.jussieu.fr/users/marus (Dated: May 8, 0) We want to describe basic properties of systems containing a large number N of particles, from mesoscopic (N 0 3 ) up to macroscopic (N 0 4 ) systems in d spatial dimensions. (In general we have d 3.) Liquids and solids are typical many-body systems. In many cases, the atoms/ molecules of the system are very well described by classical dynamics. Only at very low temperatures, the quantum nature of the particles becomes important. Liquid helium has been for long time the example of a quantum liquid, until experiments on alcaline gases have archieved the neccessary densities/ temperatures to reach quantum degeneracy. Whereas obervation of quantum effects for real particles neccessitates quite some experimental efforts, a different ind of quantum liquid is dominating every day s life, at least for condensed matter and solid state physicists: the electron gas. Many body properties are best described in the framewor of statistical mechanics, and we will give a brief introduction in quantum statistics first. A. Quantum Mechanics Single partice quantum mechanics. The state of the system is characterized by by a vector Ψ, and the mean value of an observable with operator Ô in this state is given by Ψ Ô Ψ () The state can be representated using any complete basis set, e.g. ˆr r r r is the position representation Ψ Ô Ψ dr dr Ψ (r) r Ô r Ψ(r) () Ψ(r) r Ψ, Ψ (r) Ψ r (3) In the Schrödinger eqution the time evolution of the system s state is determined by Schrödinger s equation i h Ψ Ĥ Ψ (4) t where Ĥ is the Hamiltonian of the system, and operators are time-independant. Eigenstates of Ĥ with energiegeladenes E i are stationary, and form a basis E i with Ĥ E i E i E i. Many particle quantum mechanics: Bosons and Fermions. We can form a basis for many quantum particles by labelling each of them as we do for classical particles. If we have single particle states, e.g. energy eigenstates of a single particle hamiltonian E i we can form a many-body state Ψ N E E... E N N (5) putting the first particle in an energy state E, the second into E, etc.. However, there is no possibility of distinguishing quantum particles, e.g. all Hamiltonians nown up to now are symmetric with respect to particle permutations. Therefore, all wavefunctions needed to describe the world so far are seperated in two main classes which do not mix: Total symmetric and total antisymmetric wavefunctions Ψ N S/A (±) P E P () E P ()... E P (N) N (6) N! P where the summation is over all possible N! permutations P of particle labels. Particles which are described by symmetrical wavefunctions are called Bosons, antisymmetric wavefunctions describe Fermions. The framewor for formal calculations with symmetric/antisymmetric many-body wavefunctions is called second quantization.

B. Statistical mechanics Micro-canonical ensemble. Since any energy eigenstate is equally probably in the micro-canonical ensemble, (N, V, E are constant), the expectation value of an obersable, e.g. the one-body density, ˆρ () (r) i δ(r ˆr i), is given by [ ] ˆρ () Tr δĥ,e ˆρ () (r) ] Tr [δĥ,e (7) i δ E,E i E i ˆρ () (r) E i i δ E,E i (8) where Ĥ is the Hamiltonian of the system and E i are its eigenvalues. The denominator counts just the degeneracy of the energy level under consideration, and it is common to define the entropy S of the system which is proportional to that degeneracy S(E, N, V ) B log TrδĤ,E (9) B log i δ E,Ei (0) Canonical ensemble. Let us consider now that the system is connected to a bigger bath and can exchange energy with the bath. Bath and system together form an isolated system which can be described by the micro-canonical ensemble. The probability, p i, to find the system at energy E i is equal to the probability that the bath has energy E E i p i es B(E E i)/ B i es B(E E i)/ B () where S B (E) is the entropy of the bath for a given bath-energy E. If the bath is much bigger than the system, we can expand S B (E E i ) around E i 0. We get or S B (E E i ) S B (E) ds B(E) de E i +... () p i e βei i e βei (3) where we have defined the inverse temperature from the thermodynamic relation /T S, and β / B T,. V,N Since the total entropy S B (E E S ) + S(E S ) must be maximal for the most liely energy of the system E s, the temperature of the bath must equal the temperature of the system in equilibrium. The canonical ensemble describes a system in contact with an energy bath. It is therefore characterized by the temperature T (instead of the energy), and the probability p α to find the system in a state α, p α α ˆρ(β) α (4) E The probability is expressed introducing the density operator ˆρ(β) e βĥ i E i E i e βei Z(β) The normalisation factor Z(β) is called partition function of the canonical ensemble (5) i e βei. (6) The mean value, Ô, of some obersable Ô can now be calculated ] [Ôˆρ(β) Ô Tr α Ôˆρ(β) α (7) α

One important observable is the energy E itself and the mean energy of the system at temperature T writes α H E αe βeα log Z(β) α e βeα β can be calculated directly from the partition function. The partition function, indeed, plays a central role in statistical mechanics and the connections to thermodynamics is made by introducing the free energy, F (β), via From the thermodynamic definition of the free energy, 3 (8) exp [ βf (β)]. (9) F E T S (0) we can define the entropy S from thermodynamics. Again, we see that e S is connected to the typical number of states with energy E in the system. Classical Statistical Mechanics. The classical limit of the partition function for a system with Hamiltonian Ĥ i ˆp i m + i u(ˆr i ) + i<j v(ˆr ij ) () can be obtained by assuming that the momentum and position operators commute. In this limit we have e βĥ e β ˆp i β i m e [i u(ri)+ ] v(rij) i<j () Using position eigenfunctions r i as basis of indistinguishable particles to evaluate the trace, we get for the partition function N! (±) P +P dr r P () dr r P () dr N r P (N) N e β ˆp i i m P P [ e β u(ˆri)+ v(ˆrij) i i<j ] r P () r P () r P (N) N (3) (±) P ˆp ˆp ˆp β dr r P () e m β r dr r P () e m β N r dr N r P (N) N e m rn N N! Using P e β [ i u(ri)+ i<j v(rij) ] (4) r p (π h) d/ e ip r/ h (5) we have r ˆp β e m r and we get dp r p β p e m p r N! h dn (±) P P dr dp p (π h) d e β m e ip(r r )/ h dp (π h) d e β dp [ (π) dn e β p i i /m e m (r i P (i) r i) / h β e β p m e m(r r ) / h β i u(ri)+ i<j v(rij) ] (6) (7) As long as the mean distance n /d between particle satisfies n /d λ T where λ T (π/mt) / is the thermal wave length, all permutations different from the identity will be exponentially suppressed, and we obtain the classical partition function where the energy writes E(R, P) N! h dn N dp (π) dn p i m + N i i u(r i ) + dre βe(r,p) (8) N v(r ij ) (9) i<j

The coordinates R (r, r,..., r N ) and momenta P (p, p,..., p N ) define the phase space of the classical system, and the probability of finding the particles at the point (R, P) is given by ρ(r, P)dRdP e βe(r,p) drdp (π h) dn Z(β) where dr i dr i. The momentum integration in the partition function N!(π h) dn dr dpe βe(r,p) (3) can be done explicitly in d spatial dimensions, ( ) dn/ m π h β N! 4 (30) dre β[u(r)+v (R)] (3) and only the configuration integral over the potential energy in position space remains, U(R) i u(r i) and V (R) i<j v(r ij) It is an important feature of a classical system, that inetic and potential energy separates in the partition function. Independantly of strength and form of the interactions, the distribution of momenta is always gaussian ρ(p)dp ρ(r, P)dRdP e β i p i /m (33) and the inetic energy is a direct measure of the temperature d BT p i N m p m (34) i This result is the equipartition theorem for classical particles. This relation allows us to determine the temperature in a Molecular Dynamics simulation of in experiments. Galilean invariance. Let us consider a translational invariant system of interacting particles, e.g. particles on a large ring without explicit single body confining potential, so that the total momentum, P p i, is conserved, H P M + H (35) where H is independent of the center-of mass motion, and M Nm is the total mass. If we observe the system from a different frame (denoted by a prime) moving with a velocity u, the total energy and the total momentum in the primed frame is related to that of the lab frame by P P Mu (36) H H P u + Mu [P Mu] M + H (37) If the system is stationary in the primed frame (one has to imagine that the ring contains small impurity potentials which are important to determine the stationary frame, but which are neglected in the calculations), the partition function is given by Z u Tre βh (38) If we now consider the total momentum in the lab frame P u [ Tr Pe βh ] (39) Z u we can expand for very small velocities u 0, P u β P[P u] u0 + O(u ) (40) For a classical system, momenta are always Boltzmann distributed exp[ βp /M], as we have seen above, and we have P u Mu. In general, quantum systems can show deviations from the classical result. However, as one expects the center of mass motion to become classical for macroscopic systems so that deviations vanish in this limit. As we will see, important deviations from the normal behavior may rest even for macroscopic systems, leading to the phenomena of superfluidity and supraconductivity.

5 C. Ideal Bose gas Let us consider an ideal Bose gas at low temperatures, and denote the energy eigenstates of energy ε i of the corresponding single particle Hamiltonian, h, by ε i. Since the particles do not interact, any state of the form : ε i : ε i N : ε in (4) is an eigenstate of total energy j ε i j of the N-particle Hamiltonian H i h i. However, whereas for (ficticious) quantum particles obeying Boltzmann statistics, : ε i : ε i N : ε in (4) is a new eigenstate, degenerate in energy with the previous, the symmetric states obtained from both are identical. All possible bosonic eigenstates of the Hamiltonian are therefore simply labelled by the occupation number n i of the single-particle eigenstate ε i, whereas Boltzmann statistics would also require the actual particle number. For an ideal Bose gas the total energy of the system is then given by a summation of the occupation number n i of each energy eigenstate of energy ε i E({ni}) n i ε i (43) and the partition function at temperature T /β in the canonical ensemble writes Z N i e βe({ni}) (44) {n i} where the summation over all occupation numbers is done with the constraint that they sum up to the total number of Bosons N, δ n i,n (45) {n i} n 0 n 0 Going over to the grand canonical ensemble, we can avoid this constraint introducing the chemical potential µ and we get Z N e βµn Z N (46) i n 0 e βn(ε µ) n 0 [ ] e β(εi µ) We get the mean occupation number of mode i by varying ε i, e.g., N i e βn(ε µ) (47) (48) n i δ log Z (49) δβε i (50) e β(εi µ) which is the familar Bose-Einstein distribution. Bose Einstein condensation. The number of particles N i n i is an increasing function for increasing (negative) chemical potential. However, at µ ε 0 a singularity appears, where N 0 β ɛ 0 µ, for µ ε 0 (5) It is important to distinguish if this singularity is approached for a system of finite density at finite temperature so that the singularity indicates a transition to a new phase, e.g. in contrast to a singularity occuring as a consequence of a diverging density. Therefore, let us calculate the total number of particles for a fixed system, N i>0 N i δ dɛn(ɛ) e β(ɛ (µ ε0)) (5)

6 where we have introduced the density of states N(ɛ) i δ(ɛ (ε i ε 0 )) (53) and introduced a small offset ε ε 0 > δ > 0. The convergence properties of the integral in Eq. (5) in the limit δ 0 determine if the singularity for N 0 is approached for finite parameters of the system. In the case where the density of states is non vanishing for small energies, N(ɛ) c > 0 for ɛ 0, the integral diverges Nc T log δ/t for µ ɛ 0, whereas for any N(ɛ) ɛ x with x > 0, it approaches a finite limiting value N c, and we may expect a phase transition to occur at finite temperature. Homogenous system. For a homogeneous system in 3 dimensions, where the energy eigenstates are plane waves with energies ε h /m, and the density can be expressed by ( m n V N d 3 (π) 3 e β( h /m µ) where we have introduced the thermal wavelength λ ) 3/ 8π 3 d 3 x h β π h /(mt ). Defining n e n(x βµ) λ 3 n e nβµ n 3/ (54) g z (x) n n z x n (55) we have nλ 3 g 3/ (e βµ ) (56) Einstein condensation occurs at a finite density/ temperature in three dimensions, since the density at µ 0 is finite n c λ 3.6... (57) This is the critical density of Bose Einstein condensation. Note that we approached µ 0 such that the ground state occupation density n 0 N 0 /V remains zero. Below the critical temperature (or above the critical density), βµ V (58) and the ground state density n 0. However, whereas the condensate is macroscopically occupied, the excited state occupation remains of order one. Since the density of states is non-vanishing in two or one dimensions, Bose-Einstein condensation (in the sense of a true phase transition in the thermodynamic limit) does not occur in these lower dimensions. Superfluidity. Let us now discuss the response of the homogeneous d-dimensional Bose gas to a moving system with small velocities. From the discussion above, we need to calculate the total momentum squared, e.g. in the x-direction [ ] P x x n [ x n n ] + x p x n n p p d ( βε ) n + x p x n n p (59) βµ p Whereas the last term on the rhs is clearly zero, we now use that ε /m, to simplify P x m βµ N mv n mv d d β βµ d β λ d g d/ (e βµ ) T nv m T M (60) βµ We see that for N N c the system behaves normal, howeever, for N > N c the particles in the condensate have zero momenta and will not contribute to the fluctuations of the total momenta and we have P u (N N 0 )Mu (6) and the normal mass is smaller than the total mass leading to a the superfluid mass density N 0 M/V in the noninteracting system. Non-ideal gases. Important questions arises concerning what properties of the ideal gas survive when the particles interact. How is Bose condensation and superfluidity affected and what are the changes in dilute gases? How to describe dense fluids, e.g. 4 He?

7 D. Ideal Fermi gas Let us now discuss the ideal Fermi gas where we have to antisymmetrize the Boltzmann energy-eigenstates. Similar to a Bose system, two Boltzmann states with the same set of occupation numbers but occupied by different permutation of particle labels will describe exactly the same state after antisymmetrization. In addition, we can see that if two or more particles occupy the same state, the state vanishes after antisymmetrization, e.g. Therefore, similar to bosons, the eigenenergies of the ideal system writes [ P ij ] i : ε j : ε 0 (6) E({n i }) i n i ε i (63) but, now, all occupation numbers are either zero or one, n i 0,. It is again advantageous to consider the grand canonical ensemble where we have Z e βµn Z N (64) N e βn(ε µ) e βn(ε µ) (65) i n 0 n 0 We then get the mean occupation number of mode i by varying ε i, e.g., [ + e β(εi µ)] (66) N i n i δ log Z (67) δβε i (68) e β(εi µ) + which is the Fermi-Dirac distribution. We see that, in contrast to the Bose distribution, no singularity can occur at finite temperature. In the zero temperature limit we have N i θ(µ ε i ) (69) so that all single particle levels with energies less than the chemical potential are occupied. Appendix: gaussian integrals Gaussian integrals play a central role in statistical mechanics, in d dimensions we have I d (γ) d d xe γx (70) where x x + x +... + x d. We can easily relate the integrals beween different dimensions using I d (γ) [I (γ)] d (7) In d dimensions, the integral can be canlculated easily changing to radial variables, r x + x and cos θ x /r, I d cos θ drre γr π d(r )e γr π (7) γ So we get the general result I d (γ) ( ) d/ π (73) γ