Statistical Mechanics: Lecture 1

Similar documents
CLASSICAL CONCEPT REVIEW 8

9. Momentum and Collisions in One Dimension*

PHYS 1624 University Physics I. PHYS 2644 University Physics II

Topic 3b: Kinetic Theory

The derivation of the balance equations

Physics 9e/Cutnell. correlated to the. College Board AP Physics 1 Course Objectives

ENERGY CONSERVATION The First Law of Thermodynamics and the Work/Kinetic-Energy Theorem

3. Derive the partition function for the ideal monoatomic gas. Use Boltzmann statistics, and a quantum mechanical model for the gas.

Name Partners Date. Energy Diagrams I

Mechanics 1: Conservation of Energy and Momentum

HEAT UNIT 1.1 KINETIC THEORY OF GASES Introduction Postulates of Kinetic Theory of Gases

The Quantum Harmonic Oscillator Stephen Webb

Lecture L2 - Degrees of Freedom and Constraints, Rectilinear Motion

momentum change per impact The average rate of change of momentum = Time interval between successive impacts 2m x 2l / x m x m x 2 / l P = l 2 P = l 3

Appendix A: Science Practices for AP Physics 1 and 2

Orbits of the Lennard-Jones Potential

Numerical analysis of Bose Einstein condensation in a three-dimensional harmonic oscillator potential

Unified Lecture # 4 Vectors

Entropy and the Kinetic Theory: the Molecular Picture

Isentropic flow. Wikepedia

Kinetic Theory & Ideal Gas

What is Statistical Mechanics?

) and mass of each particle is m. We make an extremely small

Lecture L3 - Vectors, Matrices and Coordinate Transformations

People have thought about, and defined, probability in different ways. important to note the consequences of the definition:

State Newton's second law of motion for a particle, defining carefully each term used.

2. Spin Chemistry and the Vector Model

Kinetic Theory of Gases. Chapter 33. Kinetic Theory of Gases

4 Microscopic dynamics

(1) The size of a gas particle is negligible as compared to the volume of the container in which the gas is placed.

Define the notations you are using properly. Present your arguments in details. Good luck!

Chapter 4 One Dimensional Kinematics

6-2. A quantum system has the following energy level diagram. Notice that the temperature is indicated

RMS Power. The Meaning of Average

arxiv:physics/ v1 [physics.ed-ph] 14 Apr 2000

5.61 Physical Chemistry 25 Helium Atom page 1 HELIUM ATOM

AP1 Electricity. 1. A student wearing shoes stands on a tile floor. The students shoes do not fall into the tile floor due to

Statistical Mechanics, Kinetic Theory Ideal Gas. 8.01t Nov 22, 2004

1. Degenerate Pressure

Online Courses for High School Students

Conservation of Momentum and Energy

THE IDEAL GAS LAW AND KINETIC THEORY

Adequate Theory of Oscillator: A Prelude to Verification of Classical Mechanics Part 2

arxiv: v1 [quant-ph] 3 Mar 2016

FUNDAMENTALS OF ENGINEERING THERMODYNAMICS

Theory of electrons and positrons

Thermodynamics: Lecture 8, Kinetic Theory

Chapter 6 Work and Energy

Let s first see how precession works in quantitative detail. The system is illustrated below: ...

1. The Kinetic Theory of Matter states that all matter is composed of atoms and molecules that are in a constant state of constant random motion

Rotation: Moment of Inertia and Torque

Topic 2: Energy in Biological Systems

Assessment Plan for Learning Outcomes for BA/BS in Physics

Precession of spin and Precession of a top

KINETIC THEORY AND THERMODYNAMICS

Work and Conservation of Energy

PRELAB: NEWTON S 3 RD LAW AND MOMENTUM CONSERVATION

Chapter 22 The Hamiltonian and Lagrangian densities. from my book: Understanding Relativistic Quantum Field Theory. Hans de Vries

Proof of the conservation of momentum and kinetic energy

1... From Microscopic to Macroscopic Behavior. We explore the fundamental differences between microscopic and macroscopic. 1.

3. Mathematical Induction

Energy comes in many flavors!

The rate of change of velocity with respect to time. The average rate of change of distance/displacement with respect to time.

Work and Energy. Work = Force Distance. Work increases the energy of an object. Energy can be converted back to work.

Prerequisite: High School Chemistry.

Free Fall: Observing and Analyzing the Free Fall Motion of a Bouncing Ping-Pong Ball and Calculating the Free Fall Acceleration (Teacher s Guide)

Thermodynamics: Lecture 2

2 Session Two - Complex Numbers and Vectors

Thermodynamics & Statistical Mechanics:

CBE 6333, R. Levicky 1 Differential Balance Equations

Vectors. Objectives. Assessment. Assessment. Equations. Physics terms 5/15/14. State the definition and give examples of vector and scalar variables.

Fric-3. force F k and the equation (4.2) may be used. The sense of F k is opposite

Figure 1.1 Vector A and Vector F

Chapter 6 The first law and reversibility

2.2 Magic with complex exponentials

Basic Concepts in Nuclear Physics

State Newton's second law of motion for a particle, defining carefully each term used.

PHY1020 BASIC CONCEPTS IN PHYSICS I

c 2008 Je rey A. Miron We have described the constraints that a consumer faces, i.e., discussed the budget constraint.

1 The basic equations of fluid dynamics

Solutions to Problems in Goldstein, Classical Mechanics, Second Edition. Chapter 7

Science Standard Articulated by Grade Level Strand 5: Physical Science

Physics Notes Class 11 CHAPTER 3 MOTION IN A STRAIGHT LINE

Indiana's Academic Standards 2010 ICP Indiana's Academic Standards 2016 ICP. map) that describe the relationship acceleration, velocity and distance.

Lab 2: Vector Analysis

DO PHYSICS ONLINE FROM QUANTA TO QUARKS QUANTUM (WAVE) MECHANICS

The content is based on the National Science Teachers Association (NSTA) standards and is aligned with state standards.

Gravitational Potential Energy

Free Electron Fermi Gas (Kittel Ch. 6)

PHYSICS TEST PRACTICE BOOK. Graduate Record Examinations. This practice book contains. Become familiar with. Visit GRE Online at

Review D: Potential Energy and the Conservation of Mechanical Energy

Eðlisfræði 2, vor 2007

F B = ilbsin(f), L x B because we take current i to be a positive quantity. The force FB. L and. B as shown in the Figure below.

Time dependence in quantum mechanics Notes on Quantum Mechanics

Chapter 10 Rotational Motion. Copyright 2009 Pearson Education, Inc.

FLAP P11.2 The quantum harmonic oscillator

Chapter 15 Collision Theory

Lecture L22-2D Rigid Body Dynamics: Work and Energy

When the fluid velocity is zero, called the hydrostatic condition, the pressure variation is due only to the weight of the fluid.

Differential Relations for Fluid Flow. Acceleration field of a fluid. The differential equation of mass conservation

Transcription:

1 Statistical Mechanics: Lecture 1 Tabish Qureshi Thermodynamics and microscopic laws Thermodynamics describes the macroscopic properties of matter phenomenologically with the aid of equations of state which are derived empirically. For thermodynamics it is of no importance, how a certain equation of state comes about. This, of course, is consistent with the large universality of thermodynamics, namely that same laws of thermodynamics hold for different materials. However, this doesn t tell us what makes, say, the specific heat of one material different from the other. It is intuitively obvious that specific heat of one material is different from that of the other because microscopically the materials are different. The macroscopic quantities in materials must obviously be resulting from the microscopics properties. For example, the pressure of a gas is due to the collisions of the molecules with a surface, whereas temperature is directly given by the mean kinetic energy of the particles. On the other end, the microscopic laws of physics describe the behaviour of individual particles with their interaction, very well. For classical particles we use Newton s equations of motion, and for quantum

2 particles we use the Schrödinger equation. Although microscopics laws describe the behavior of particles accurately, they do not tell us how a huge collection of particles, of the order of 10 23, would behave on the average. It is the task of statistical mechanics to answer this question, namely how microscopic behavior of particles or small constituents, leads to a particular macroscopic property of the material. Thus, statistical mechanics provides a connection between microscopic physics and thermodynamics. One may take the view that microscopic laws describe the physics of one particles, and hence, they should describe the behavior of an assembly of particles (however large). A gas of N classical particles can be described by a set of coupled equations of motion d 2 r 1 = 1 F1 ( r dt 2 1, r 2... r N ) m 1 d 2 r 2 = 1 F2 ( r dt 2 1, r 2... r N ) m 2... =... d 2 r N = 1 FN ( r dt 2 1, r 2... r N ) (1) m N where r i is the position vector of the i th particle, with mass m i, and F i is the force acting on it because of interaction with all other particles, and any external

3 influence. A solution of these equations would allow us to know exact position and velocity of each particle at any future time. That information allows us to know the microscopic state (which we will henceforth call microstate) of the gas at every instant of time. Solving these coupled equations analytically is generally not possible. However, one can solve such equations numerically, using computers. However, for a realistic situation N is of the order of 10 23, and solving such a large number of equations numerically is beyond the capacity of any existing computer. One should realize that for describing the macroscopic state of a gas, characterized by pressure, volume, temperature (which we will henceforth call macro we do not need information on every microscopic detail which these coupled equations could provide. A huge number of microstates may correspond to the same, single macrostate. We would, for instance, be interested in knowing the pressure of the gas, and not bother about what a particular atom of the gas is doing at every instance. In other words, we only need some average macroscopic quantities, and not every microscopic detail. So, solving these coupled differential equations would anyway be an overkill, if at all we were able to solve them. We need a conceptually different approach to this problem.

4 Phase space Let us first be more specific about the the concept of microstates. For a classical system it is sufficient to know at a time t all generalized coordinates q i (t) and momenta p i (t) to uniquely specify the state of motion of the system. Thus for a mechanical system we can interpret the set {q i, p i, i = 1, 2,...N} as the microstate of this system. For a single particle in one dimension, there is only one position variable x and one momentum variable p x. If we plot x on the x-axis and p x on the y-axis, a point on the graph will represent one state of the particle. As the particle moves in time, the point will follow a trajectory. We will call the space described by x and p x, phase space, the point representing a particular value of x and p x, a phase point, and the trajectory followed by the point, the phase trajectory. In this particular case, the phase space is 2-dimensional. If a single particle moves in 3 dimensions, we would need 6 coordinate axes for x, y, z, p x, p y, p z. For N particles in 3-dimensions, the phase space will be 6Ndimensional. So, the set {q i, p i } can now be understood as a point in a 6N-dimensional phase space. A point in this phase space describes particular value of position and momentum values of all N particles. Hence, a definite point in this phase space exactly corresponds to one microscopic state of motion of the whole system.

5 PHASE TRAJECTORIES p 2mE E ω 2 2mE 2E mω2 p 0 x 0 x L 2mE 2mE Harmonic Oscillator Particle in a box The trajectory in phase space is governed by Hamilton s equations q i = H p i, ṗ i = H q i where the Hamiltonian H(q,(t), p,(t)) corresponds to the (possibly time-dependent) total energy of the system. It is a function of the phase-space point (q, p), and of time. In a closed global system, in which the Hamiltonian does not depend explicitly on time, the total energy E = H(p(t), q(t)) is a conserved quantity. Thus, the phase trajectory always moves on a constant-energy curve or multidimensional surface.

6 Statistical Ensembles Each microstate of a N-particle gas is represented by a point in the 6N-dimensional phase space. So, if one considers all possible microstates which the gas can have, one will have a huge collection of points in the phase space. This collection of points in the phase space, of all possible microstates of the system, is called an ensemble. Or one can imagine each point in the phase space representing an imaginary copy of the system, each in a different microstate. This collection of imaginary copies of the system, each in a different microstate, is called an ensemble. When the gas evolves in time, in the phase space it basically goes from one phase point to the next, in a specific sequence. Any macroscopic quantity of the gas which we measure, is not measured instantaneously. Rather it is measured over a finite time, which is very long compared to the timescale of motion of the particles of the gas. So, the measured quantity is actually a time-averaged quantity. The basic idea of statistical mechanics is the following. In doing a time-average of a quantity, one is basically looking at different values the quantity takes, as it goes from one microstate to the other, during its time evolution. And then one takes an average of all the values of the quantity. As far as taking the

7 average is concerned, it is not important what the sequence in going from one microstate to the other is. One can just take the phase points of the ensemble, over which the system goes, and take the average. In other words, the time average can be replaced by the average over the whole ensemble. To enable one to replace time-average by ensembleaverage, some conditions have to be satisfied. First of all, average over the ensemble implicitly assumes that the system visits all phase points during its time evolution. Not only that, it is also assumed that all microstates are equally likely to be visited. So it better happen so in reality. If the system spends more time in certain microstates, and less in some others, our assumption will break down. This assumption constitutes, what is called, the ergodic hypothesis. The ergodic hypothesis says that, over long periods of time, the time spent by a system in some region of the phase space of microstates with the same energy is proportional to the phase-volume of this region, i.e., all accessible microstates are equiprobable over a long period of time. Ergodic hypothesis is a pillar of statistical mechanics, however, it cannot be proven in general. It is assumed to be true, and finds justification in the fact that statistical mechanics turns out to be a successful theory, in agreement with experiments.

8 We will first consider an isolated system, typically a gas enclosed in a box, which is thermally insulated. So, any time evolution of the system will be subject to the constraint that the total energy remains constant. Left for a long time, it is believed to be in equilibrium. We further assume that given an isolated system in equilibrium, it is found with equal probability in each of its accessible microstates. This is the postulate of equal a priory probability. This postulate is at the very core of statistical mechanics. Now each macrostate comprises of numerous microstates. For example, all the gas confined to only one half of the box, is a macrostate. There is a huge number of ways this can happen, by various arrangements of particles and their momenta. The gas uniformly occupying the whole volume of the box, is another macrostate. And again, there are a huge lot of microstates associated with this macrostate. Now each microstate is equally probable, but we never actually see a gas occupying only one half of its container. Why does that happen? It happens because the number of microstates associated with the gas occupying the whole volume are overwhelmingly large, compared to the microstates associated with the gas occupying only one half of the box. One can get an idea of the numbers involved in such situations, by a simple example. Let there be an array of 4 noninteracting magnetic moments, each of

9 which can only take values +1 or -1, in some suitable units. Now, assume that each magnetic moment is free to flip up and down, i.e., +1 or -1. We can use the total magnetic moment of the array to describe a macrostate. There is only one microstate associated with the macrostate with total magnetic moment 4, which is (+1+1+1+1). If one considers the macrostate with total magnetic moment zero, the microstates associated with it are (+1+1-1-1), (-1-1+1+1), (+1-1+1-1), (+1-1-1+1), (-1+1+1-1), (-1+1-1+1). So, there are 6 microstates associated with total magnetic moment 0, while only 1 with magnetic moment 4. One can easily see that if there were 10 magnetic moments, there would still be only one microstate associated with total magnetic moment 10. However, the number of microstates associated with total magnetic moment 0 in that case, will be overwhelmingly large. So, in a system of 10 magnetic moments which are freely flipping, one will almost never see magnetic moment 10, and the magnetic moment will appear to be zero or very small all the time. A close analogy can be drawn from probability theory. Suppose there are different people doing 100 coin-tosses each. Almost all of them will get nearly equal number of heads and tails. It is next to impossible for anyone to obtain 100 heads or 100 tails. Probability of a 100 heads is (1/2) 100, which is in-

10 finitisimally small. Probability of any single configuration is also exactly the same. For example, the probability of getting alternate heads and tails for all 100 tosses is also (1/2) 100. However, the number of configurations in which the number of heads and tails are 50-50 is overwhelmingly large. In the light of the above argument, we conclude that equilibrium state is the one in which the number of microstates is maximum. In reality, the system will go over all microstates, as the ergodic hypothesis states, but it will be mostly found in certain macrostates