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)

Similar documents
The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 20, March 8, 2006

Lecture #13 1. Incorporating a vector potential into the Hamiltonian 2. Spin postulates 3. Description of spin states 4. Identical particles in

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m.

Rotations in Quantum Mechanics

Identical Particles in Quantum Mechanics

Quantum Physics Lecture 9

Are these states normalized? A) Yes

For example, in one dimension if we had two particles in a one-dimensional infinite potential well described by the following two wave functions.

Solution Set of Homework # 6 Monday, December 12, Textbook: Claude Cohen Tannoudji, Bernard Diu and Franck Laloë, Second Volume

Coupling of Angular Momenta Isospin Nucleon-Nucleon Interaction

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

CHM Physical Chemistry II Chapter 9 - Supplementary Material. 1. Constuction of orbitals from the spherical harmonics

P3317 HW from Lecture and Recitation 10

Preliminary Quantum Questions

Quantum Physics II (8.05) Fall 2002 Assignment 12 and Study Aid

P3317 HW from Lecture and Recitation 7

Quantum Statistics (2)

Brief review of Quantum Mechanics (QM)

International Physics Course Entrance Examination Questions

LS coupling. 2 2 n + H s o + H h f + H B. (1) 2m

Chemistry 881 Lecture Topics Fall 2001

Sommerfeld (1920) noted energy levels of Li deduced from spectroscopy looked like H, with slight adjustment of principal quantum number:

Chemistry 120A 2nd Midterm. 1. (36 pts) For this question, recall the energy levels of the Hydrogenic Hamiltonian (1-electron):

CHEM3023: Spins, Atoms and Molecules

3.024 Electrical, Optical, and Magnetic Properties of Materials Spring 2012 Recitation 8 Notes

Quantum Physics 2006/07

1 Measurement and expectation values

1 Mathematical preliminaries

3 Quantization of the Dirac equation

Next topic: Quantum Field Theories for Quantum Many-Particle Systems; or "Second Quantization"

C/CS/Phys 191 Uncertainty principle, Spin Algebra 10/11/05 Fall 2005 Lecture 13

Basic Quantum Mechanics

Notes on Spin Operators and the Heisenberg Model. Physics : Winter, David G. Stroud

1 Multiplicity of the ideal gas

1 Recall what is Spin

Chemistry 483 Lecture Topics Fall 2009

Potential energy, from Coulomb's law. Potential is spherically symmetric. Therefore, solutions must have form

Quantum Theory of Many-Particle Systems, Phys. 540

For the case of S = 1/2 the eigenfunctions of the z component of S are φ 1/2 and φ 1/2

( ) dσ 1 dσ 2 + α * 2

PHY331 Magnetism. Lecture 8

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11

Page 712. Lecture 42: Rotations and Orbital Angular Momentum in Two Dimensions Date Revised: 2009/02/04 Date Given: 2009/02/04

(1.1) In particular, ψ( q 1, m 1 ; ; q N, m N ) 2 is the probability to find the first particle

1 Quantum field theory and Green s function

Atomic Structure. Chapter 8

( ) ( ) ( ) Invariance Principles & Conservation Laws. = P δx. Summary of key point of argument

Intermission: Let s review the essentials of the Helium Atom

Section 10: Many Particle Quantum Mechanics Solutions

C. Show your answer in part B agrees with your answer in part A in the limit that the constant c 0.

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

Introduction to Electronic Structure Theory

Applied Nuclear Physics (Fall 2006) Lecture 8 (10/4/06) Neutron-Proton Scattering

16.1. PROBLEM SET I 197

Derivation of the Boltzmann Distribution

G : Quantum Mechanics II

Chapter Electron Spin. * Fine structure:many spectral lines consist of two separate. lines that are very close to each other.

QFT. Unit 1: Relativistic Quantum Mechanics

Introduction to particle physics Lecture 3: Quantum Mechanics

Quantum Mechanics: Fundamentals

ψ(t) = U(t) ψ(0). (6.1.1)

P3317 HW from Lecture 15 and Recitation 8

Quantum Mechanics of Atoms

Degeneracy & in particular to Hydrogen atom

Luigi Paolasini

Angular Momentum set II

Multi-Particle Wave functions

Quantum mechanics of many-fermion systems

Lecture 11 Spin, orbital, and total angular momentum Mechanics. 1 Very brief background. 2 General properties of angular momentum operators

in-medium pair wave functions the Cooper pair wave function the superconducting order parameter anomalous averages of the field operators

Electron States of Diatomic Molecules

Lecture 24 Origins of Magnetization (A number of illustrations in this lecture were generously provided by Prof. Geoffrey Beach)

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

Energy Level Energy Level Diagrams for Diagrams for Simple Hydrogen Model

Problem Set # 1 SOLUTIONS

CHAPTER 6: AN APPLICATION OF PERTURBATION THEORY THE FINE AND HYPERFINE STRUCTURE OF THE HYDROGEN ATOM. (From Cohen-Tannoudji, Chapter XII)

Path Integral for Spin

221B Lecture Notes Many-Body Problems I

Atoms, Molecules and Solids. From Last Time Superposition of quantum states Philosophy of quantum mechanics Interpretation of the wave function:

11 Group Theory and Standard Model

3: Many electrons. Orbital symmetries. l =2 1. m l

129 Lecture Notes More on Dirac Equation

r R A 1 r R B + 1 ψ(r) = αψ A (r)+βψ B (r) (5) where we assume that ψ A and ψ B are ground states: ψ A (r) = π 1/2 e r R A ψ B (r) = π 1/2 e r R B.

Lecture 14 The Free Electron Gas: Density of States

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Quantum Physics II (8.05) Fall 2002 Outline

The Quantum Heisenberg Ferromagnet

St Hugh s 2 nd Year: Quantum Mechanics II. Reading. Topics. The following sources are recommended for this tutorial:

Clebsch-Gordan Coefficients

Lecture 3: Helium Readings: Foot Chapter 3

( R) MOLECULES: VALENCE BOND DESCRIPTION VALENCE BOND APPROXIMATION METHOD. Example: H 2. at some fixed R, repeat for many different values

Many Body Quantum Mechanics

Physics 125 Course Notes Identical Particles Solutions to Problems F. Porter

3.024 Electrical, Optical, and Magnetic Properties of Materials Spring 2012 Recitation 3 Notes

Identical Particles. Bosons and Fermions

Isospin. K.K. Gan L5: Isospin and Parity 1

Transcription:

LECTURE 12 Maxwell Velocity Distribution Suppose we have a dilute gas of molecules, each with mass m. If the gas is dilute enough, we can ignore the interactions between the molecules and the energy will not depend on the positions of the molecules. Let f( r, v)d 3 rd 3 v be the mean number of molecules with center of mass position between r and r + d r and velocity between v and v + d v. Then f( r, v)d 3 rd 3 v = Ce β(mv2 /2) d 3 rd 3 v (1) where C is a constant determined by normalization. Here we have used the Boltzmann distribution e βe with the energy E being the kinetic energy of the system mv 2 /2. The contribution to the energy from internal states or degrees of freedom doesn t affect the velocity; these internal states are summed over ( r exp( βɛ r )) and will just contribute to C. The constant C is determined by the condition d 3 r d 3 vf( r, v) = N (2) We can evaluate the integrals to obtain C: or C d 3 r d 3 ve β(mv2 /2) C = n = CV ( e β(mv2 x /2) dv x ) 3 ) 3/2 = ( 2π CV βm = N (3) ( ) 3/2 βm (4) 2π where n N/V is the total number of molecules per unit volume. Hence f( v)d 3 rd 3 v = n ( ) 3/2 βm e βmv2 /2 d 3 rd 3 v (5) 2π or ( ) m 3/2 f( v)d 3 rd 3 v = n e mv 2 /2kT d 3 rd 3 v (6) 2πkT We have omitted r from the argument of f since f doesn t depend on r. Equation (6) is known as the Maxwell velocity distribution for a dilute gas in thermal equilibrium. Notice that it is a Gaussian distribution centered at v = 0. Recall that the exponential factor in a Gaussian distribution has the form exp( x 2 /2σ 2 ) where 2σ is the width of the distribution. Then the width of the Maxwell velocity distribution is σ = kt/m. Notice that the width increases as the temperature increases. This means there are more hot molecules at higher temperatures. By hot molecules I mean ones with large v.

This applies for both positive and negative velocities because in (6) v 2 = v 2 x + v 2 y + v 2 z and each component of velocity can be positive or negative. Quantum Statistics So far we have mostly concentrated on systems where classical mechanics is applicable. We have invoked quantum mechanics only to the extent that we had to acknowledge that quantum mechanically systems have discrete states and discrete energy levels. This becomes particularly apparent at low temperatures. We also noted that identical particles are indistinguishable. That s how we resolved the Gibbs paradox. Now we need to include some more quantum mechanics in order to treat the statistics of systems of identical particles where the classical approximation no longer is valid. Before we get to quantum statistics, let s go over some preliminaries. Quantum Numbers In your quantum mechanics course, when you solved the problem of a particle in a one dimensional box, you labelled the eigenvalues and eigenvectors with an integer n. where Hψ n = E n ψ n (7) E n = n2 π 2 h 2 2ma 2 (8) with n = 1, 2, 3,... n is an example of a quantum number. It s a number which characterizes an eigenstate of the system. It changes from eigenstate to eigenstate, but if the system is in that eigenstate, it remains in that eigenstate and the quantum numbers describing that state don t change either. This is why an eigenstate is called a stationary state. Typically the quantum numbers are associated with symmetries. Symmetry means that the system looks the same under certain operations. For example if the system is invariant under time translation, then the Hamiltonian is time independent, and energy is conserved. This was true of the case of a particle in a box, and we labelled the energy eigenstates by n. Another possible symmetry is spatial translation. If the potential is invariant under spatial translation, then momentum is a good quantum number. This would be true of a constant potential or if there were no potential. Classically momentum is conserved as long as the system is not subjected to a force. So if F = d p/dt = 0, then p = constant. The force is the gradient of the potential: F = V ( r). So momentum is a good quantum number as long as the potential does not vary spatially, i.e, as long as it s constant. (We don t usually refer to forces in quantum mechanics.) If the potential is constant or zero, then we have a free particle with momentum p, energy E = p 2 /2m and ψ exp(i p r/ h). Another way to define a conserved quantity is to recall the Heisenberg equation of motion. For wavefunctions, it takes the form i h ψ t = Hψ (9) 2

There is also an equation of motion for operators. Let  be some operator. Its equation of motion is i h  = [Â, Ĥ] (10) t So if i h Â/ t = 0, then we must have [Â, Ĥ] = 0. This is the condition that A is a constant of the motion and is a conserved quantity. So if momentum is a conserved quantity, then it commutes with the Hamiltonian: [Ĥ, ˆp] = Ĥ ˆp ˆpĤ = 0 (11) But this is getting too technical. Angular Momentum Angular momentum is another quantity that is conserved. There are two basic types of angular momentum: orbital and spin. Classically the orbital angular momentum is conserved as long as there is no torque on the system. In quantum mechanics we don t usually talk about torque. Rather we say that orbital angular momentum is a good quantum number if the system has continuous rotational symmetry. By rotational symmetry, I mean that if you rotate the system in some way, like by an angle θ about the z axis, it still looks the same. An atom is a spherical kind of object, and it has rotational symmetry. Spin angular momentum is an internal angular momentum that is associated with a particle. (If the particle has rotational symmetry in spin space, then spin is a good quantum number.) The spin operator is denoted by S. There are 2 quantum numbers associated with spin angular momentum: s and s z. Sometimes s z is denoted by m s. If spin is a good quantum number, then the energy eigenstate ψ n is also an eigenstate of Ŝ 2 and Ŝz: Ŝ 2 ψ n = h 2 s(s + 1)ψ n (12) and Ŝ z ψ n = hs z ψ n (13) (In terms of commutators, [Ĥ, Ŝ2 ] = 0, [Ĥ, ˆ S] = 0, and [ Ŝ 2, ˆ S] = 0.) It turns out that sz has a range of values: s z = s, s + 1,..., s 1, s (14) There are 2s + 1 values of s z. s can be either an integer or a half integer. Particles with integer spin are called bosons and particles with half integer spin are called fermions. An example of a fermion is an electron. An electron is a spin-1/2 particle, i.e., s = 1/2, and it has 2 spin states: spin up (which corresponds to s z = +1/2) and spin down (which corresponds to s z = 1/2). Protons and neutrons are also spin 1/2 particles and are therefore fermions. An example of a boson is a photon. A photon is a spin 1 particle, i.e, s = 1, and it has s z = 1 or s z = +1. It turns out not to have s z = 0. This is related to the fact that 3

there are 2 possible polarizations of the electric field E and that E k, where k points in the direction of propagation of the electromagnetic wave. Adding Angular Momenta One can add angular momenta. For example one can add the spin angular momenta of two independent particles, say the spin of the proton and the neutron in a deuteron: ˆ S = ˆ S1 + ˆ S2 (15) The total angular momentum also has 2 quantum numbers associated with it: S and S z. If these are good quantum numbers, then the energy eigenstate ψ n satisfies: and Ŝ 2 ψ n = h 2 S(S + 1)ψ n (16) Ŝ z ψ n = hs z ψ n (17) One can also add orbital angular momenta. The rules for angular momentum addition are the same for all types of angular momentum. But the rules for adding angular momenta in quantum mechanics are a little tricky. You might think that if one particle has angular momentum s 1 (be it spin or orbital or total) and another independent particle has angular momentum s 2, the total is s = s 1 + s 2. This is not necessarily so. Classically you don t add two vectors by adding their magnitudes: s 1 + s 2 s 1 + s 2. You don t do this in quantum mechanics either. In quantum mechanics the rule is that if you add s 1 and s 2, the total s obeys s 1 s 2 s s 1 + s 2 (18) The z component s z still runs from s to +s in integer steps; there are 2s + 1 values of s z. So if we add the spin angular momentum of 2 spin 1/2 particles, the total spin S of the system is S = 0 (we call this a singlet) or S = 1 (we call this a triplet). The singlet state has s z = 0 and the triplet state has 3 possible values of s z : 1, 0, +1. In this way one can make a composite particle that is a boson out of fermions. For example, 4 He is a boson because if you add the spin of the proton, neutron, and 2 electrons, you always will get an integer. On the other hand 3 He is a fermion. Many Particle States versus Single Particle States Before proceeding, I want to clarify the difference between eigenstates of the entire system and single particle eigenstates. When we summed over microstates r in the past, we were thinking that each microstate was a many particle eigenstate. However, in quantum mechanics it is often easier to solve Schrodinger s equation for a single particle and ignore many body interactions. Then we approximate a many particle system as a sum of single particle systems. Thus we often speak in terms of single particle states. Pauli Exclusion Principle As we mentioned earlier, fermions are particles which have half integer spin. Electrons are an example of fermions. All electrons are the same. They are indistinguishable particles. If we have a system with many electrons, as in a multielectron atom, the Pauli 4

exclusion principle states that there can never be more than one electron in the same quantum state. In other words, a given set of quantum numbers can only be assigned to at most one electron. This is true for each type of fermion. For example, there cannot be more than one proton in a given quantum state. Consider the quantum mechanical problem of a box with infinitely high walls. We can label the states with quantum numbers n, spin s and the z component of spin s z. Suppose we put 4 electrons in this box. Two electrons, one spin up and the other spin down, go into the n = 1 state. The quantum numbers of these two electrons is (n = 1, s = 1/2, s z = 1/2) and (n = 1, s = 1/2, s z = 1/2). The other two electrons go into the n = 2 state. They have quantum numbers (n = 2, s = 1/2, s z = 1/2) and (n = 2, s = 1/2, s z = 1/2). V = infinity n=2 n=1 x = 0 V = 0 x = a Exchange Symmetry of Bosons and Fermions The wavefunction of a collection of fermions is antisymmetric under the exchange of any 2 fermions. For example, if we have 2 fermions with coordinates 1 and 2, and we put them into 2 states φ a and φ b, an antisymmetric wavefunction for them is: ψ(1, 2) = φ a (1)φ b (2) φ a (2)φ b (1) (19) Notice that if we interchange 1 and 2, we get ψ(2, 1) = ψ(1, 2). This is what we mean by antisymmetry. If fermions 1 and 2 were both in the same state, say φ a, then we would get φ a (1)φ a (2) φ a (2)φ a (1) = 0 (20) Thus antisymmetry enforces the Pauli exclusion principle. In general a wavefunction describing a collection of N fermions must be antisymmetric and satisfy ψ(1, 2,..., r,..., s,..., N) = ψ(1, 2,..., s,..., r,..., N) (21) Bosons are symmetric under exchange. For example, if we have 2 bosons with coordinates 1 and 2, and we put them into 2 states φ a and φ b, a symmetric wavefunction for them is: ψ(1, 2) = φ a (1)φ b (2) + φ a (2)φ b (1) (22) 5

If we interchange 1 and 2, we get ψ(2, 1) = ψ(1, 2). This is what we mean by a symmetric wavefunction. If bosons 1 and 2 were both in the same state, say φ a, then we would get φ a (1)φ a (2) + φ a (2)φ a (1) = 2φ a (1)φ a (2) 0 (23) Thus it s ok to put more than one boson in the same single particle state. In general a wavefunction describing a collection of N bosons must be symmetric and must satisfy ψ(1, 2,..., r,..., s,..., N) = +ψ(1, 2,..., s,..., r,..., N) (24) 6