The Representation of Time and Change in Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

The Representation of Time and Change in Mechanics GORDON BELOT THE REPRESENTATION OF TIME AND CHANGE IN MECHANICS ABSTRACT: This chapter is concerned with the representation of time and change in classical (i.e., non-quantum) physical theories. One of the main goals of the chapter is to attempt to clarify the nature and scope of the so-called problem of time: a knot of technical and interpretative problems that appear to stand in the way of attempts to quantize general relativity, and which have their roots in the general covariance of that theory. The most natural approach to these questions is via a consideration of more clear cases. So much of the chapter is given over to a discussion of the representation of time and change in other, better understood theories, starting with the most straightforward cases and proceeding through a consideration of cases that lead up to the features of general relativity that are responsible for the problem of time. Keywords: Classical Mechanics; General Relativity; Symmetry; Time Draft version 2.0. Forthcoming in J. Butterfield and J. Earman (eds.), Handbook of the Philosophy of Physics. North-Holland. CONTENTS 1 INTRODUCTION . 1 2 HAMILTONIAN AND LAGRANGIAN MECHANICS . 7 2.1 The n-Body Problem . 7 2.2 The Hamiltonian Approach . 8 2.3 The Lagrangian Approach . 11 3 SYMPLECTIC MATTERS . 14 3.1 Preliminaries . 14 3.2 Symplectic Manifolds . 17 3.3 Presymplectic Manifolds . 19 3.4 Symplectic Structures and Quantization . 21 4 LAGRANGIAN FIELD THEORY . 22 4.1 The Lagrangian Approach . 25 4.2 The Structure of the Space of Solutions . 28 4.3 Symmetries and Conserved Quantities . 29 5 TIME AND CHANGE IN WELL-BEHAVED FIELD THEORIES . 30 5.1 The Lagrangian Picture . 31 5.2 The Hamiltonian Picture . 33 5.3 Relation between the Lagrangian and Hamiltonian Pictures . 36 5.4 Time and Change . 37 5.5 Overview . 38 6 COMPLICATIONS . 39 6.1 Singular Dynamics . 40 6.2 Gauge Freedom . 45 6.3 Time-Dependent Systems . 59 7 THE PROBLEM OF TIME IN GENERAL RELATIVITY . 64 7.1 The General Covariance of General Relativity . 64 7.2 The Problem of Time . 77 7.3 Finding Time in General Relativity . 80 ACKNOWLEDGEMENTS . 89 BIBLIOGRAPHY . 89 THE REPRESENTATION OF TIME AND CHANGE IN MECHANICS 1 If time is objective the physicist must have discovered that fact, if there is Be- coming the physicist must know it; but if time is merely subjective and Being is timeless, the physicist must have been able to ignore time in his construction of reality and describe the world without the help of time. If there is a solution to the philosophical problem of time, it is written down in the equations of mathe- matical physics. Perhaps it would be better to say that the solution is to be read between the lines of the physicist’s writings. Physical equations formulate specific laws . but philosophical analysis is concerned with statements about the equations rather than with the content of the equations themselves. —Reichenbach.1 For many years I have been tormented by the certainty that the most extraor- dinary discoveries await us in the sphere of Time. We know less about time than about anything else. —Tarkovsky.2 1 INTRODUCTION This chapter is concerned with the representation of time and change in classical (i.e., non-quantum) physical theories. One of the main goals of the chapter is to attempt to clarify the nature and scope of the so-called problem of time: a knot of technical and interpretative problems that appear to stand in the way of attempts to quantize general relativity, and which have their roots in the general covariance of that theory. The most natural approach to these questions is via consideration of more clear cases. So much of the chapter is given over to a discussion of the representation of time and change in other, better understood theories, starting with the most straightforward cases and proceeding through a consideration of cases that pre- pare one, in one sense or another, for the features of general relativity that are responsible for the problem of time. Let me begin by saying a bit about what sort of thing I have in mind in speak- ing of the representation of time and change in physical theories, grounding the discussion in the most tractable case of all, Newtonian physics. As a perfectly general matter, many questions and claims about the content of a physical theory admit of two construals—as questions about structural features of solutions to the equations of motion of the theory, or as questions about structural features of these equations. For instance, on the one hand time appears as an aspect of the spacetimes in which physics unfolds—that is, as an aspect of the background in which the solutions to the equations of the theory are set. On the other, time is represented via its role in the laws of physics—in particular, in its 1[1991, pp. 16 f.]. 2[1991, p. 53]. 2 GORDON BELOT role in the differential equations encoding these laws. So questions and claims about the nature of time in physical theories will admit of two sorts of reading. Consider, for instance, the claim that time is homogeneous in Newtonian physics (or, as Newton would put it, that time flows equably). There are two sorts of fact that we might look to as grounding this claim. 1. There is a sense in which time is a separable aspect of the spacetime of Newtonian physics and there is a sense in which time, so considered, is homogenous.3 2. The laws of the fundamental-looking theories of classical mechanics (e.g., Newton’s theory of gravity) are time translation invariant—the differential equations of these theories do not change their form when the origin of the temporal coordinate is changed—so the laws of such theories are indifferent to the identity of the instants of time. In the Newtonian setting, these two sorts of considerations mesh nicely and pro- vide mutual support: there is a consilience between the symmetries of the laws and the symmetries of spacetime. But in principle, the two sorts of considera- tion need not lead to the same sort of answer: one might consider a system in Newtonian spacetime that is subject to time-dependent forces; or one could set the Newtonian n-body problem in a spacetime which a featured a preferred instant, but otherwise had the structure of Newtonian spacetime. And as one moves away from the familiar setting of Newtonian physics, it becomes even more important to distinguish the two approaches: in general relativity, the laws have an enormous (indeed, infinite-dimensional) group of symmetries while generic solutions have no symmetries whatsoever. In discussing the representation of time and change, this chapter will focus on structural features of the laws of physical theories rather than on features of par- ticular solutions. To emphasize this point, I will say that I am interested in the structure of this or that theory as a dynamical theory. I will approach my topics via the Lagrangian and Hamiltonian approaches to classical theories, two great over-arching—and intimately related—frameworks in which such topics are naturally addressed.4 Roughly speaking, in each of these approaches the content of the equations of a theory is encoded in certain structures 3(Neo)Newtonian spacetime is partitioned in a natural way by instants of absolute simultaneity, and time can be identified with the structure that the set of these instants inherits from the structure of spacetime: time then has the structure of an affine space modelled on the real numbers—so for any two instants, there is a temporal symmetry which maps one to the other. 4Why pursue our question within the realm of Lagrangian and Hamiltonian mechanics rather than working directly with the differential equations of theories? Because the benefits are large: these over- arching approaches provide powerful mathematical frameworks in which to compare theories. And because the costs are minimal: almost every theory of interest can be put into Lagrangian or Hamil- tonian form, without any obvious change of content. And because it leads us where we want to go: current attempts to understand the content of classical physical theories are necessarily shaped by ef- forts to construct or understand deeper, quantum theories; and it appears that a classical theory must be placed in Lagrangian or Hamiltonian form in order to be quantized. THE REPRESENTATION OF TIME AND CHANGE IN MECHANICS 3 on a space of possibilities associated with the theory.5 In the Lagrangian approach the featured space is the space of solutions to the equations of the theory, which for heuristic purposes we can identify with the space of possible worlds allowed by the theory.6 On the Hamiltonian side, the featured space is the space of initial data for the equations of the theory, which we can in the same spirit identify with the space of possible instantaneous states allowed by the theory. In Newtonian mechanics, the reflection within the Lagrangian framework of the time translation invariance of the laws is that the space of solutions is itself invari- ant under time translations: given a set of particle trajectories in spacetime obeying Newton’s laws of motion, we can construct the set of particle trajectories that re- sult if all events are translated in time by amount t; the latter set is a solution (i.e., is permitted by the laws of motion) if and only if the former set is; furthermore, the map that carries us from a solution to its time translate preserves the structure on the space of solutions that encodes the dynamics of the theory.
Recommended publications
  • An Introduction to Quantum Field Theory
    AN INTRODUCTION TO QUANTUM FIELD THEORY By Dr M Dasgupta University of Manchester Lecture presented at the School for Experimental High Energy Physics Students Somerville College, Oxford, September 2009 - 1 - - 2 - Contents 0 Prologue....................................................................................................... 5 1 Introduction ................................................................................................ 6 1.1 Lagrangian formalism in classical mechanics......................................... 6 1.2 Quantum mechanics................................................................................... 8 1.3 The Schrödinger picture........................................................................... 10 1.4 The Heisenberg picture............................................................................ 11 1.5 The quantum mechanical harmonic oscillator ..................................... 12 Problems .............................................................................................................. 13 2 Classical Field Theory............................................................................. 14 2.1 From N-point mechanics to field theory ............................................... 14 2.2 Relativistic field theory ............................................................................ 15 2.3 Action for a scalar field ............................................................................ 15 2.4 Plane wave solution to the Klein-Gordon equation ...........................
    [Show full text]
  • Structure” of Physics: a Case Study∗ (Journal of Philosophy 106 (2009): 57–88)
    The “Structure” of Physics: A Case Study∗ (Journal of Philosophy 106 (2009): 57–88) Jill North We are used to talking about the “structure” posited by a given theory of physics. We say that relativity is a theory about spacetime structure. Special relativity posits one spacetime structure; different models of general relativity posit different spacetime structures. We also talk of the “existence” of these structures. Special relativity says that the world’s spacetime structure is Minkowskian: it posits that this spacetime structure exists. Understanding structure in this sense seems important for understand- ing what physics is telling us about the world. But it is not immediately obvious just what this structure is, or what we mean by the existence of one structure, rather than another. The idea of mathematical structure is relatively straightforward. There is geometric structure, topological structure, algebraic structure, and so forth. Mathematical structure tells us how abstract mathematical objects t together to form different types of mathematical spaces. Insofar as we understand mathematical objects, we can understand mathematical structure. Of course, what to say about the nature of mathematical objects is not easy. But there seems to be no further problem for understanding mathematical structure. ∗For comments and discussion, I am extremely grateful to David Albert, Frank Arntzenius, Gordon Belot, Josh Brown, Adam Elga, Branden Fitelson, Peter Forrest, Hans Halvorson, Oliver Davis Johns, James Ladyman, David Malament, Oliver Pooley, Brad Skow, TedSider, Rich Thomason, Jason Turner, Dmitri Tymoczko, the philosophy faculty at Yale, audience members at The University of Michigan in fall 2006, and in 2007 at the Paci c APA, the Joint Session of the Aristotelian Society and Mind Association, and the Bellingham Summer Philosophy Conference.
    [Show full text]
  • On the Time Evolution of Wavefunctions in Quantum Mechanics C
    On the Time Evolution of Wavefunctions in Quantum Mechanics C. David Sherrill School of Chemistry and Biochemistry Georgia Institute of Technology October 1999 1 Introduction The purpose of these notes is to help you appreciate the connection between eigenfunctions of the Hamiltonian and classical normal modes, and to help you understand the time propagator. 2 The Classical Coupled Mass Problem Here we will review the results of the coupled mass problem, Example 1.8.6 from Shankar. This is an example from classical physics which nevertheless demonstrates some of the essential features of coupled degrees of freedom in quantum mechanical problems and a general approach for removing such coupling. The problem involves two objects of equal mass, connected to two different walls andalsotoeachotherbysprings.UsingF = ma and Hooke’s Law( F = −kx) for the springs, and denoting the displacements of the two masses as x1 and x2, it is straightforward to deduce equations for the acceleration (second derivative in time,x ¨1 andx ¨2): k k x −2 x x ¨1 = m 1 + m 2 (1) k k x x − 2 x . ¨2 = m 1 m 2 (2) The goal of the problem is to solve these second-order differential equations to obtain the functions x1(t)andx2(t) describing the motion of the two masses at any given time. Since they are second-order differential equations, we need two initial conditions for each variable, i.e., x1(0), x˙ 1(0),x2(0), andx ˙ 2(0). Our two differential equations are clearly coupled,since¨x1 depends not only on x1, but also on x2 (and likewise forx ¨2).
    [Show full text]
  • An S-Matrix for Massless Particles Arxiv:1911.06821V2 [Hep-Th]
    An S-Matrix for Massless Particles Holmfridur Hannesdottir and Matthew D. Schwartz Department of Physics, Harvard University, Cambridge, MA 02138, USA Abstract The traditional S-matrix does not exist for theories with massless particles, such as quantum electrodynamics. The difficulty in isolating asymptotic states manifests itself as infrared divergences at each order in perturbation theory. Building on insights from the literature on coherent states and factorization, we construct an S-matrix that is free of singularities order-by-order in perturbation theory. Factorization guarantees that the asymptotic evolution in gauge theories is universal, i.e. independent of the hard process. Although the hard S-matrix element is computed between well-defined few particle Fock states, dressed/coherent states can be seen to form as intermediate states in the calculation of hard S-matrix elements. We present a framework for the perturbative calculation of hard S-matrix elements combining Lorentz-covariant Feyn- man rules for the dressed-state scattering with time-ordered perturbation theory for the asymptotic evolution. With hard cutoffs on the asymptotic Hamiltonian, the cancella- tion of divergences can be seen explicitly. In dimensional regularization, where the hard cutoffs are replaced by a renormalization scale, the contribution from the asymptotic evolution produces scaleless integrals that vanish. A number of illustrative examples are given in QED, QCD, and N = 4 super Yang-Mills theory. arXiv:1911.06821v2 [hep-th] 24 Aug 2020 Contents 1 Introduction1 2 The hard S-matrix6 2.1 SH and dressed states . .9 2.2 Computing observables using SH ........................... 11 2.3 Soft Wilson lines . 14 3 Computing the hard S-matrix 17 3.1 Asymptotic region Feynman rules .
    [Show full text]
  • Advanced Quantum Mechanics
    Advanced Quantum Mechanics Rajdeep Sensarma [email protected] Quantum Dynamics Lecture #9 Schrodinger and Heisenberg Picture Time Independent Hamiltonian Schrodinger Picture: A time evolving state in the Hilbert space with time independent operators iHtˆ i@t (t) = Hˆ (t) (t) = e− (0) | i | i | i | i i✏nt Eigenbasis of Hamiltonian: Hˆ n = ✏ n (t) = cn(t) n c (t)=c (0)e− n | i | i n n | i | i n X iHtˆ iHtˆ Operators and Expectation: A(t)= (t) Aˆ (t) = (0) e Aeˆ − (0) = (0) Aˆ(t) (0) h | | i h | | i h | | i Heisenberg Picture: A static initial state and time dependent operators iHtˆ iHtˆ Aˆ(t)=e Aeˆ − Schrodinger Picture Heisenberg Picture (0) (t) (0) (0) | i!| i | i!| i Equivalent description iHtˆ iHtˆ Aˆ Aˆ Aˆ Aˆ(t)=e Aeˆ − ! ! of a quantum system i@ (t) = Hˆ (t) i@ Aˆ(t)=[Aˆ(t), Hˆ ] t| i | i t Time Evolution and Propagator iHtˆ (t) = e− (0) Time Evolution Operator | i | i (x, t)= x (t) = x Uˆ(t) (0) = dx0 x Uˆ(t) x0 x0 (0) h | i h | | i h | | ih | i Z Propagator: Example : Free Particle The propagator satisfies and hence is often called the Green’s function Retarded and Advanced Propagator The following propagators are useful in different contexts Retarded or Causal Propagator: This propagates states forward in time for t > t’ Advanced or Anti-Causal Propagator: This propagates states backward in time for t < t’ Both the retarded and the advanced propagator satisfies the same diff. eqn., but with different boundary conditions Propagators in Frequency Space Energy Eigenbasis |n> The integral is ill defined due to oscillatory nature
    [Show full text]
  • The Liouville Equation in Atmospheric Predictability
    The Liouville Equation in Atmospheric Predictability Martin Ehrendorfer Institut fur¨ Meteorologie und Geophysik, Universitat¨ Innsbruck Innrain 52, A–6020 Innsbruck, Austria [email protected] 1 Introduction and Motivation It is widely recognized that weather forecasts made with dynamical models of the atmosphere are in- herently uncertain. Such uncertainty of forecasts produced with numerical weather prediction (NWP) models arises primarily from two sources: namely, from imperfect knowledge of the initial model condi- tions and from imperfections in the model formulation itself. The recognition of the potential importance of accurate initial model conditions and an accurate model formulation dates back to times even prior to operational NWP (Bjerknes 1904; Thompson 1957). In the context of NWP, the importance of these error sources in degrading the quality of forecasts was demonstrated to arise because errors introduced in atmospheric models, are, in general, growing (Lorenz 1982; Lorenz 1963; Lorenz 1993), which at the same time implies that the predictability of the atmosphere is subject to limitations (see, Errico et al. 2002). An example of the amplification of small errors in the initial conditions, or, equivalently, the di- vergence of initially nearby trajectories is given in Fig. 1, for the system discussed by Lorenz (1984). The uncertainty introduced into forecasts through uncertain initial model conditions, and uncertainties in model formulations, has been the subject of numerous studies carried out in parallel to the continuous development of NWP models (e.g., Leith 1974; Epstein 1969; Palmer 2000). In addition to studying the intrinsic predictability of the atmosphere (e.g., Lorenz 1969a; Lorenz 1969b; Thompson 1985a; Thompson 1985b), efforts have been directed at the quantification or predic- tion of forecast uncertainty that arises due to the sources of uncertainty mentioned above (see the review papers by Ehrendorfer 1997 and Palmer 2000, and Ehrendorfer 1999).
    [Show full text]
  • Branched Hamiltonians and Supersymmetry
    Branched Hamiltonians and Supersymmetry Thomas Curtright, University of Miami Wigner 111 seminar, 12 November 2013 Some examples of branched Hamiltonians are explored, as recently advo- cated by Shapere and Wilczek. These are actually cases of switchback poten- tials, albeit in momentum space, as previously analyzed for quasi-Hamiltonian dynamical systems in a classical context. A basic model, with a pair of Hamiltonian branches related by supersymmetry, is considered as an inter- esting illustration, and as stimulation. “It is quite possible ... we may discover that in nature the relation of past and future is so intimate ... that no simple representation of a present may exist.” – R P Feynman Based on work with Cosmas Zachos, Argonne National Laboratory Introduction to the problem In quantum mechanics H = p2 + V (x) (1) is neither more nor less difficult than H = x2 + V (p) (2) by reason of x, p duality, i.e. the Fourier transform: ψ (x) φ (p) ⎫ ⎧ x ⎪ ⎪ +i∂/∂p ⎪ ⇐⇒ ⎪ ⎬⎪ ⎨⎪ i∂/∂x p − ⎪ ⎪ ⎪ ⎪ ⎭⎪ ⎩⎪ This equivalence of (1) and (2) is manifest in the QMPS formalism, as initiated by Wigner (1932), 1 2ipy/ f (x, p)= dy x + y ρ x y e− π | | − 1 = dk p + k ρ p k e2ixk/ π | | − where x and p are on an equal footing, and where even more general H (x, p) can be considered. See CZ to follow, and other talks at this conference. Or even better, in addition to the excellent books cited at the conclusion of Professor Schleich’s talk yesterday morning, please see our new book on the subject ... Even in classical Hamiltonian mechanics, (1) and (2) are equivalent under a classical canonical transformation on phase space: (x, p) (p, x) ⇐⇒ − But upon transitioning to Lagrangian mechanics, the equivalence between the two theories becomes obscure.
    [Show full text]
  • BRST APPROACH to HAMILTONIAN SYSTEMS Our Starting Point Is the Partition Function
    BRST APPROACH TO HAMILTONIAN SYSTEMS A.K.Aringazin1, V.V.Arkhipov2, and A.S.Kudusov3 Department of Theoretical Physics Karaganda State University Karaganda 470074 Kazakstan Preprint KSU-DTP-10/96 Abstract BRST formulation of cohomological Hamiltonian mechanics is presented. In the path integral approach, we use the BRST gauge fixing procedure for the partition func- tion with trivial underlying Lagrangian to fix symplectic diffeomorphism invariance. Resulting Lagrangian is BRST and anti-BRST exact and the Liouvillian of classical mechanics is reproduced in the ghost-free sector. The theory can be thought of as a topological phase of Hamiltonian mechanics and is considered as one-dimensional cohomological field theory with the target space a symplectic manifold. Twisted (anti- )BRST symmetry is related to global N = 2 supersymmetry, which is identified with an exterior algebra. Landau-Ginzburg formulation of the associated d = 1, N = 2 model is presented and Slavnov identity is analyzed. We study deformations and per- turbations of the theory. Physical states of the theory and correlation functions of the BRST invariant observables are studied. This approach provides a powerful tool to investigate the properties of Hamiltonian systems. PACS number(s): 02.40.+m, 03.40.-t,03.65.Db, 11.10.Ef, 11.30.Pb. arXiv:hep-th/9811026v1 2 Nov 1998 [email protected] [email protected] [email protected] 1 1 INTRODUCTION Recently, path integral approach to classical mechanics has been developed by Gozzi, Reuter and Thacker in a series of papers[1]-[10]. They used a delta function constraint on phase space variables to satisfy Hamilton’s equation and a sort of Faddeev-Popov representation.
    [Show full text]
  • Antibrackets and Supersymmetric Mechanics
    Preprint JINR E2-93-225 Antibrackets and Supersymmetric Mechanics Armen Nersessian 1 Laboratory of Theoretical Physics, JINR Dubna, Head Post Office, P.O.Box 79, 101 000 Moscow, Russia Abstract Using odd symplectic structure constructed over tangent bundle of the symplec- tic manifold, we construct the simple supergeneralization of an arbitrary Hamilto- arXiv:hep-th/9306111v2 29 Jul 1993 nian mechanics on it. In the case, if the initial mechanics defines Killing vector of some Riemannian metric, corresponding supersymmetric mechanics can be refor- mulated in the terms of even symplectic structure on the supermanifold. 1E-MAIL:[email protected] 1 Introduction It is well-known that on supermanifolds M the Poisson brackets of two types can be defined – even and odd ones, in correspondence with their Grassmannian grading 1. That is defined by the expression ∂ f ∂ g {f,g} = r ΩAB(z) l , (1.1) κ ∂zA κ ∂zB which satisfies the conditions p({f,g}κ)= p(f)+ p(g) + 1 (grading condition), (p(f)+κ)(p(g)+κ) {f,g}κ = −(−1) {g, f}κ (”antisymmetrisity”), (1.2) (p(f)+κ)(p(h)+κ) (−1) {f, {g, h}1}1 + cycl.perm.(f, g, h) = 0 (Jacobi id.), (1.3) l A ∂r ∂ where z are the local coordinates on M, ∂zA and ∂zA denote correspondingly the right and the left derivatives, κ = 0, 1 denote correspondingly the even and the odd Poisson brackets. Obviously, the even Poisson brackets can be nondegenerate only if dimM = (2N.M), and the odd one if dimM =(N.N). With nondegenerate Poisson bracket one can associate the symplectic structure A B Ωκ = dz Ω(κ)ABdz , (1.4) BC C where Ω(κ)AB Ωκ = δA .
    [Show full text]
  • Classical Hamiltonian Field Theory
    Classical Hamiltonian Field Theory Alex Nelson∗ Email: [email protected] August 26, 2009 Abstract This is a set of notes introducing Hamiltonian field theory, with focus on the scalar field. Contents 1 Lagrangian Field Theory1 2 Hamiltonian Field Theory2 1 Lagrangian Field Theory As with Hamiltonian mechanics, wherein one begins by taking the Legendre transform of the Lagrangian, in Hamiltonian field theory we \transform" the Lagrangian field treatment. So lets review the calculations in Lagrangian field theory. Consider the classical fields φa(t; x¯). We use the index a to indicate which field we are talking about. We should think of x¯ as another index, except it is continuous. We will use the confusing short hand notation φ for the column vector φ1; : : : ; φn. Consider the Lagrangian Z 3 L(φ) = L(φ, @µφ)d x¯ (1.1) all space where L is the Lagrangian density. Hamilton's principle of stationary action is still used to determine the equations of motion from the action Z S[φ] = L(φ, @µφ)dt (1.2) where we find the Euler-Lagrange equations of motion for the field d @L @L µ a = a (1.3) dx @(@µφ ) @φ where we note these are evil second order partial differential equations. We also note that we are using Einstein summation convention, so there is an implicit sum over µ but not over a. So that means there are n independent second order partial differential equations we need to solve. But how do we really know these are the correct equations? How do we really know these are the Euler-Lagrange equations for classical fields? We can obtain it directly from the action S by functional differentiation with respect to the field.
    [Show full text]
  • Hamiltonian Systems and Noether's Theorem
    HAMILTONIAN SYSTEMS AND NOETHER'S THEOREM DANIEL SPIEGEL Abstract. This paper uses the machinery of symplectic geometry to make rigorous the mathematical framework of Hamiltonian mechanics. This frame- work is then shown to imply Newton's laws and conservation of energy, thus validating it as a physical theory. We look at symmetries of physical systems in the form of Lie groups, and show that the Hamiltonian framework grants us the insight that the existence of a symmetry corresponds to the conserva- tion of a physical quantity, i.e. Noether's Theorem. Throughout the paper we pay heed to the correspondence between mathematical definitions and physical concepts, and supplement these definitions with examples. Contents 1. Introduction 1 2. Hamiltonian Systems 2 3. Symmetries and Lie Groups 5 4. Noether's Theorem 9 Acknowledgments 11 References 11 1. Introduction In 1834, the Irish mathematician William Hamilton reformulated classical New- tonian mechanics into what is now known as Hamiltonian mechanics. While New- ton's laws are valuable for their clarity of meaning and utility in Cartesian coordi- nates, Hamiltonian mechanics provides ease of computation for many other choices of coordinates (as did the Lagrangian reformulation in 1788), tools for handling sta- tistical systems with large numbers of particles, and a more natural transition into quantum mechanics. We refer the reader to chapter 13 of John Taylor's Classical Mechanics [3] for an introductory discussion of these uses. This paper, however, will focus on the insights that Hamilton's formalism provides with regards to the conservation of physical quantities. For all these advantages, it is clear why Hamil- tonian mechanics is an important area of study for any physics student.
    [Show full text]
  • On the Complete Symmetry Group of the Classical Kepler System J
    On the complete symmetry group of the classical Kepler system J. Krause Citation: Journal of Mathematical Physics 35, 5734 (1994); doi: 10.1063/1.530708 View online: https://doi.org/10.1063/1.530708 View Table of Contents: http://aip.scitation.org/toc/jmp/35/11 Published by the American Institute of Physics Articles you may be interested in Symmetry transformations of the classical Kepler problem Journal of Mathematical Physics 14, 1125 (1973); 10.1063/1.1666448 Laplace-Runge-Lenz vector in quantum mechanics in noncommutative space Journal of Mathematical Physics 54, 122106 (2013); 10.1063/1.4835615 The complete Kepler group can be derived by Lie group analysis Journal of Mathematical Physics 37, 1772 (1996); 10.1063/1.531496 Illustrating dynamical symmetries in classical mechanics: The Laplace–Runge–Lenz vector revisited American Journal of Physics 71, 243 (2003); 10.1119/1.1524165 Prehistory of the ’’Runge–Lenz’’ vector American Journal of Physics 43, 737 (1975); 10.1119/1.9745 More on the prehistory of the Laplace or Runge–Lenz vector American Journal of Physics 44, 1123 (1976); 10.1119/1.10202 On the complete symmetry group of the classical Kepler system J. Krause Facultad de Fisica, Pontijcia Vniversidad Cato’lica de Chile, Casilla 306, Santiago 22, Chile (Received 18 April 1994; accepted for publication 17 May 1994) A rather strong concept of symmetry is introduced in classical mechanics, in the sense that some mechanical systems can be completely characterized by the sym- metry laws they obey. Accordingly, a “complete symmetry group” realization in mechanics must be endowed with the following two features: (1) the group acts freely and transitively on the manifold of all allowed motions of the system; (2) the given equations of motion are the only ordinary differential equations that remain invariant under the specified action of the group.
    [Show full text]