HAMILTON S PRINCIPLE

Similar documents
for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Curves in the configuration space Q or in the velocity phase space Ω satisfying the Euler-Lagrange (EL) equations,

REVIEW. Hamilton s principle. based on FW-18. Variational statement of mechanics: (for conservative forces) action Equivalent to Newton s laws!

Time-Dependent Statistical Mechanics 5. The classical atomic fluid, classical mechanics, and classical equilibrium statistical mechanics

Physics 5153 Classical Mechanics. Canonical Transformations-1

Chapter 1. Principles of Motion in Invariantive Mechanics

NIU PHYS 500, Fall 2006 Classical Mechanics Solutions for HW6. Solutions

Canonical transformations (Lecture 4)

The first order formalism and the transition to the

Sketchy Notes on Lagrangian and Hamiltonian Mechanics

Physics 106a, Caltech 13 November, Lecture 13: Action, Hamilton-Jacobi Theory. Action-Angle Variables

Liouville Equation. q s = H p s

PHYS 705: Classical Mechanics. Hamiltonian Formulation & Canonical Transformation

Under evolution for a small time δt the area A(t) = q p evolves into an area

Second quantization: where quantization and particles come from?

06. Lagrangian Mechanics II

Physics 235 Chapter 7. Chapter 7 Hamilton's Principle - Lagrangian and Hamiltonian Dynamics

Lecture 11 : Overview

= 0. = q i., q i = E

4.1 Important Notes on Notation

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

Lecture 5. Alexey Boyarsky. October 21, Legendre transformation and the Hamilton equations of motion

M2A2 Problem Sheet 3 - Hamiltonian Mechanics

The Principle of Least Action

The kinetic equation (Lecture 11)

L(q, q) = m 2 q2 V (q) 2 m + V (q)

15. Hamiltonian Mechanics

(Most of the material presented in this chapter is taken from Thornton and Marion, Chap. 7) p j . (5.1) !q j. " d dt = 0 (5.2) !p j . (5.

Homework 4. Goldstein 9.7. Part (a) Theoretical Dynamics October 01, 2010 (1) P i = F 1. Q i. p i = F 1 (3) q i (5) P i (6)

A Short Essay on Variational Calculus

Physics 106b: Lecture 7 25 January, 2018

Classical mechanics of particles and fields

Nonlinear Single-Particle Dynamics in High Energy Accelerators

Physical Dynamics (SPA5304) Lecture Plan 2018

2 Canonical quantization

Forces of Constraint & Lagrange Multipliers

PHYS2100: Hamiltonian dynamics and chaos. M. J. Davis

Chapter 11. Hamiltonian formulation Legendre transformations

Legendre Transforms, Calculus of Varations, and Mechanics Principles

Lecture 19: Calculus of Variations II - Lagrangian

EULER-LAGRANGE TO HAMILTON. The goal of these notes is to give one way of getting from the Euler-Lagrange equations to Hamilton s equations.

Statistical Mechanics Solution Set #1 Instructor: Rigoberto Hernandez MoSE 2100L, , (Dated: September 4, 2014)

Use conserved quantities to reduce number of variables and the equation of motion (EOM)

LAGRANGIAN AND HAMILTONIAN

Goldstein Problem 2.17 (3 rd ed. # 2.18)

Dirac Structures and the Legendre Transformation for Implicit Lagrangian and Hamiltonian Systems

Physical Dynamics (PHY-304)

Conservation of total momentum

Variation Principle in Mechanics

Hamiltonian Field Theory

Motion under the Influence of a Central Force

Part II. Classical Dynamics. Year

Dynamics and Stability application to submerged bodies, vortex streets and vortex-body systems

M3-4-5 A16 Notes for Geometric Mechanics: Oct Nov 2011

Hamiltonian flow in phase space and Liouville s theorem (Lecture 5)

arxiv: v1 [quant-ph] 4 Nov 2008

INC 693, 481 Dynamics System and Modelling: Lagrangian Method III Dr.-Ing. Sudchai Boonto Assistant Professor

Solution Set Two. 1 Problem #1: Projectile Motion Cartesian Coordinates Polar Coordinates... 3

Classical Mechanics and Electrodynamics

Canonical transformations and exact invariants for time-dependent Hamiltonian systems

The Microcanonical Approach. (a) The volume of accessible phase space for a given total energy is proportional to. dq 1 dq 2 dq N dp 1 dp 2 dp N,

Classical Equations of Motion

Lecture 4. Alexey Boyarsky. October 6, 2015

Housekeeping. No announcement HW #7: 3.19

III. Kinetic Theory of Gases

D Alembert s principle of virtual work

The Simple Double Pendulum

221A Lecture Notes Notes on Classica Mechanics I

The Klein-Gordon equation

The Geometry of Euler s equation. Introduction

University of Utrecht

Averaging II: Adiabatic Invariance for Integrable Systems (argued via the Averaging Principle)

Physics 106a, Caltech 4 December, Lecture 18: Examples on Rigid Body Dynamics. Rotating rectangle. Heavy symmetric top

Classical Mechanics and Electrodynamics

On singular lagrangians and Dirac s method

Hamiltonian Mechanics

Physics 351, Spring 2015, Homework #6. Due at start of class, Friday, February 27, 2015

Homework 3. 1 Goldstein Part (a) Theoretical Dynamics September 24, The Hamiltonian is given by

Lecture I: Constrained Hamiltonian systems

Hamiltonian Solution I

Chaotic motion. Phys 750 Lecture 9

Noether s Theorem. 4.1 Ignorable Coordinates

LSZ reduction for spin-1/2 particles

REVIEW REVIEW. A guess for a suitable initial state: Similarly, let s consider a final state: Summary of free theory:

7 Pendulum. Part II: More complicated situations

CE 530 Molecular Simulation

Seminar 6: COUPLED HARMONIC OSCILLATORS

Summary of free theory: one particle state: vacuum state is annihilated by all a s: then, one particle state has normalization:

Lecture 10. Central potential

From quantum to classical statistical mechanics. Polyatomic ideal gas.

The two body problem involves a pair of particles with masses m 1 and m 2 described by a Lagrangian of the form:

B7.1 Classical Mechanics

Electric and Magnetic Forces in Lagrangian and Hamiltonian Formalism

Lecture 1: Historical Overview, Statistical Paradigm, Classical Mechanics

FINAL EXAM GROUND RULES

8. INTRODUCTION TO STATISTICAL THERMODYNAMICS

Illustrating Dynamical Symmetries in Classical Mechanics: The Laplace-Runge-Lenz Vector Revisited

Concepts in Theoretical Physics

Variational principles and Hamiltonian Mechanics

Supplement on Lagrangian, Hamiltonian Mechanics

Transcription:

HAMILTON S PRINCIPLE In our previous derivation of Lagrange s equations we started from the Newtonian vector equations of motion and via D Alembert s Principle changed coordinates to generalised coordinates ending up with Lagrange s equations of motion. There is another way to express the basic laws of mechanics in a single statement which is equivalent to Lagrange s equations; this statement is called Hamilton s Principle or sometimes the Principle of Least Action. It differs fundamentally from the differential equation formulation in that it refers to the entire history of a system s motion between two distinct times. To introduce the idea we first define the action for a system of n degrees of freedom by S[q 1 q 2... q n ] = L(q i (t) q i (t) t) where and t 2 are respectively an initial and a final time and L is the Lagrangian for the system which depends on the n generalised coordinates q i (t) the n generalised velocities q i (t) and possibly the time (if there are explicitly time- dependent external fields). Because we integrate over a finite time interval we see that S depends on the entire trajectory of the system in configuration space between and t 2. Because S depends on the entire trajectory rather than just on the values of the coordinates and velocities at one instant we say that S is a functional of the trajectories q i (t). Hamilton s Principle can be stated thus: Amongst all the possible trajectories q i (t) which take the system from initial configuration q i1 = q i ( ) at time to final configuration q i2 = q i (t 2 ) at time t 2 the physical trajectory is the one which makes the action an extremum (usually a minimum but could be a saddle point). Hamilton s Principle is stated in terms of the action S which is a scalar quantity that shares all the invariances of the Lagrangian L and which is independent of any particular choice of generalised coordinates. However we can show that Hamilton s Principle implies that the trajectory which minimizes the action is the one that also satisfies Lagrange s equations so we see that the two formulations are equivalent. For simplicity we consider a system with one degree of freedom so that there is only one generalised coordinate q(t). The system starts at position q 1 = q( ) and finishes at position q 2 = q(t 2 ). We can denote the physical trajectory which joins these initial and final positions by q(t) which satisfies q( ) = q 1 and q(t 2 ) = q 2. To implement Hamilton s Principle we now have to consider all possible sufficiently smooth trajectories q(t) different from q(t) that start and end at q 1 q 2 respectively and then show that S[q(t)] has an extremum at q(t). We can express all possible other trajectories q(t) in terms of their deviation ɛ(t) from the physical trajectory q(t) as q(t) = q(t) + ɛ(t) where at the endpoints we must have ɛ( ) = ɛ(t 2 ) = 0 since all paths must start at q 1 and finish at q 2. To find a minimum or an extremum it is sufficient to consider

nearby paths such that ɛ(t) is small. Thus we can expand the action in powers of ɛ. To do this we first Taylor expand the Lagrangian as which gives for the action S[q] = S[q + ɛ] = L(q q t) = L(q + ɛ q + ɛ t) = L(q q t) + ɛ q + ɛ +... We can rearrange this as L(q + ɛ q + ɛ t) = S[q + ɛ] S[q] = L(q q t) + ɛ q + ɛ +... and then in the second integral perform an interation by parts ɛ [ = ɛ(t) ] t2 ɛ d ɛ q + ɛ +.... ( ). Since ɛ vanishes at the two endpoints and t 2 there is no contribution from these endpoints and we have δs = S[q + ɛ] S[q] = ( ɛ q d ( )) +... This expression for δs which is first order in the small quantity ɛ is called the first variation of S. It plays the role for functionals that a first derivative plays for an ordinary function. In particular we see that if S is to have a minimum or an extremum at q(t) this first variation must vanish identically for any choice of ɛ(t). That in turn will be possible only if for all times t between and t 2 t t 2 the coefficient of ɛ(t) in the integral above vanishes namely that q d ( ) = 0. However this is just Lagrange s equation for q(t) showing that Hamilton s principle predicts the same mechanical trajectory that Lagrange s equations would predict. For more than one degree of freedom the same argument goes through with appropriate subscripts added such as ɛ i (t) etc. HAMILTONIAN MECHANICS In addition to showing us how Lagrange s equations can be derived from the action principle Hamilton showed us an entirely novel way to reformulate the basic laws of mechanics so that ideas of geometry and topology can come to the fore in analyzing mechanical motion. As a by-product he gave us a form of classical mechanics which can be transformed directly into quantum mechanics. To understand the changed point of view which takes us to Hamiltonian mechanics we first recall that in Lagrangian mechanics the positions q i and velocities q i are the fundamental variables and the Lagrangian itself is a function of these variables

and of the time if there are explicitly time dependent fields present L = L(q i q i t). If we make a small change in all these variables then L changes as dl = dq i + d q i + i t or using the definition of generalised momentum p i = / i and the Lagrange equations of motion / = ṗ i dl = p i d q i + t. Recall also that in Lagrangian mechanics the energy E is defined as E = p i q i L and so we would expect it to depend on positions and velocities as well. However if we calculate the change in E analogous to the change in L above we get something unexpected de = p i d q i + q i dp i dl = de = p i d q i + q i dp i ṗ i dq i q i dp i t p i d q i t where two terms have cancelled out. Unlike the expression for dl above which says that L is in fact a function of the independent variables q i and q i the expression for de says that mathematically we should regard E as a function of the independent variables q i p i and t. How can we realize this explicitly? In the Lagrangian point of view p i = / i = p i (q i q i t) so that p i is a secondary quantity which is a function of q i q i t. However if we can invert this equation and find q i as a function of q i p i and t then we can eliminate all velocities from the energy E in favour of positions q i and momenta p i. If we have carried out this transformation we define the resulting energy function as the Hamiltonian H. In other words E(q i q i t) H(q i p i t). We then write the infinitesimal change in the energy as dh(q i p i t) = q i dp i t. From this we can read off (analogous to the way that we use the thermodynamic identity in thermal physics) that q i = H p i ṗ i = H H t = t If there is no explicit dependence on time in L then there will be no explicit dependence on time in H and we have the 2n first order equations q i = H p i ṗ i = H i = 1 2... n..

as the fundamental equations of motion for the system. These are called the Hamilton equations of motion or sometimes the canonical equations of motion. In this point of view the energy function H generates the equations of motion and the variables q i p i appear on an equal footing rather than the momenta being seen as secondary quantities as in the Lagrangian picture. The origin of conservation laws in the Hamiltonian formulation lies in the symmetries of the system just as in the Lagrangian framework. To see this note that if the Hamiltonian does not depend on a particular q i (i.e. q i is a cyclic variable) then from the Hamilton equation ṗ i = H/ = 0 we have that the corresponding momentum is conserved. For the energy itself we calculate dh(q i p i t) = dq i H + dp i H + H p i t and using the Hamilton equations of motion to express q i and ṗ i dh(q i p i t) = [ q i ṗ i + ṗ i q i ] + H t = H t. Thus if there is no explicit time dependence ( H/ t = 0) in the Hamiltonian (i.e. we have time translation symmetry) the energy (which is H) is conserved as well. PHASE SPACE In the Lagrangian picture we talked about configuration space the n-dimensional space whose coordinates are the generalised coordinates q i for i = 1 2... n. In the Hamiltonian picture since both the q i and p i appear on an equal footing we introduce a 2n-dimensional space whose coordinates are the q i and p i. This space we call phase space. A single point in phase space specifies completely the position and momentum of every particle in the system that is a single point completely specifies the microscopic state of the system. If we specify such a point at time t = 0 we can use the associated q i (0) and p i (0) as initial conditions for Hamilton s equations and if we solve these equations we find q i (t) and p i (t) for all later (and earlier!) times. As a function of time the phase point traces out a phase trajectory which describes the entire history of the system. Since each point in phase space is a possible initial state we can say that the phase trajectories fill phase space but no two trajectories can ever intersect because the solution to Hamilton s equations is uniquely specified by the initial conditions. As a result we can consider that under the Hamiltonian time evolution the entire phase space flows like a liquid. Using Hamilton s equations of motion we can prove that the volume of any region of phase space is constant under this evolution (Liouville s theorem) although the shape of the region may distort in an unbelievably complicated way. As an example consider the simple harmonic oscillator with one degree of freedom. The Lagrangian L and the energy E are respectively L = 1 2 m q2 1 2 kq2 E = p q L = 1 2 m q2 + 1 2 kq2

where p = / = m q. To form the Hamiltonian we express q in terms of p q = p/m and substitute in E to get H(q p) = p2 2m + 1 2 kq2 with two Hamilton equations of motion q = H p = p m ṗ = H q = kq. To find the phase trajectories we can either solve Hamilton s equations for q(t) and p(t) and plot the curve parametrically using time as the parameter or we can remember that with no explicit time dependence the Hamiltonian is constant on any phase trajectory and we can get the equation of the trajectory by solving H(q p) = constant for p as a function of q. Either way we find families of ellipses as discussed in the lecture. As a second example of how to set up a Hamiltonian description consider the spherical pendulum discussed in an earlier summary sheet which has two degrees of freedom. The two convenient generalised coordinates are the spherical polar angles θ and φ. The Lagrangian was L = T V = m 2 ( a 2 θ2 + a 2 sin 2 θ φ 2) mga cos θ and the energy E was E = p θ θ + pφ φ L = T + V = m 2 The generalised momenta are ( a 2 θ2 + a 2 sin 2 θ φ 2) + mga cos θ. p θ = θ = ma2 θ pφ = φ = ma2 sin 2 θ φ. If we invert these to express the velocities in terms of the momenta θ = p θ ma 2 φ = p φ ma 2 sin 2 θ and then eliminate the velocities from the energy E we get the Hamiltonian H(θ φ p θ p φ ) = p θ 2 2 2ma 2 + p φ 2ma 2 sin 2 + mga cos θ. θ We observe that φ is a cyclic variable so from the Hamilton equations we have that p φ is conserved. The Hamiltonian then effectively reduces to a function of θ and p θ only and we could draw phase trajectories in the reduced two dimensional phase space with coordinates θ p θ.