Basic Hamiltonian Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Basic Hamiltonian Mechanics BASIC HAMILTONIAN MECHANICS B.W. Montague 1 . INTRODUCTION At the time of Newton, mechanics was considered mainly in terms of forces, masses and velocities, since these quantities were directly tangible in terms of everyday experience. However, the subsequent evolution of celestial mechanics called for more compact and general methods of handling dynamical systems, and led to the increasing use of potentials as the relevant quantities in the mathematical treatment. Another important development of this period was to free classical mechanics from the constraints of specific co-ordinate systems and to facilitate the transformation from one system to another. This is not only a matter of convenience but a powerful tool for finding invariants of the motion, and a fundamental feature of the Hamiltonian formulation. Joseph—Louis Lagrange was one of the outstanding pioneers of this development;his . Méchanique Analytique laid the foundations of the analytic, in contrast with the geometrical, approach to generalised dynamics. Nevertheless he suggested, apparently as a concession to tradition, that mechanics (with the time dimension included), might be considered as the geometry of a four-dimensional space, which was in a sense a precursor of the structure of special relativity. Generalised classical mechanics has developed considerably since the time of Lagrange and remains not only a broad and fundamental part of classical physics as a whole but has become an important part of the framework on which quantum mechanics has been formulated. The range of topics is so large that even in the restricted field of particle accelerators our coverage here of Lagrangian and Hamiltonian dynamics can only be rather limited. More detailed presentations of the subjects can be found in the Bibliography and are cited in the text wherever the need is specially manifest. The Lagrange equations of motion can be presented in a number of different versions, depending on the nature of the problem and the form of the dynamical constraints. A detailed discussion of these topics is given by Goldstein [1] in Chapter 1; for the application to accelerator physics and the Hamiltonian formulation it is sufficient to consider a restricted set of conditions, namely the motion of charged particles in electromagnetic fields, a domain which is thoroughly covered by Jackson [2] in Chapter 12. Furthermore we shall be dealing with the motion of single particles, taking no account of the forces due to space charge. In the Lagrangian formulation the dynamical behaviour of a system with lc degrees of freedom is characterised by a set of generalised "position" co—ordinates qk, generalised "velocity" co-ordinates qk = dqk / dt and the independent variable t, which is often, but not necessarily, the time variable. The "Lagrangian" or Lagrange function L of the form: L(qkvqk>t) is then a function of 2k dynamical variables. 2 . OUTLINE OF LAGRANGIAN AND HAMILTONIAN FORMALISM In the simplest, non-relativistic case where the forces can be derived from a scalar potential V, independent of velocity, the Lagrangian takes the specific form: OCR Output L(q,¢?»¢) = T(<1.<?J) — V(q,r). where T is the kinetic energy, V is the potential energy and the index k is implicit. In the presence of electromagnetic fields, which can be time-dependent, a generalised potential, and consequently a Lagrangian, can be formulated by combining the Lorentz equation for the force with Maxwell’s equations for the electromagnetic fields. The Lorentz equation for the force F on a particle of charge e moving with velocity v in an electromagnetic field is given by F=e[E+v><B], (3) where E and B are respectively the electric and magnetic fields. Now in the case of static fields the part of the force F arising from E can be derived from a scalar potential tp, but the presence of a magnetic field and time-variation of either, requires the introduction of a vector potential A, which contributes to both electric and magnetic forces. Since we are not taking account here of space-charge forces, only two of Maxwell's equations are required, namely: curlE+dB/dz=O (4) div B = O (5) From Eq. (5) it follows that one can write: B = curl A , (6) where A is the magnetic vector potential. Substituting Eq. (6) into Eq. (4) expanding and rearranging terms, it is then straightforward but somewhat lengthy to show that the non relativistic Lagrangian for tirne—dependent electromagnetic fields is: L(q»¤?,r) = T(¢1»¢?»¢) - U (cmr) (7) where the scalar potential V of Eq. (2) has been replaced by a generalised potential: U = ed) — A · v The relativistic Lagrangian is not just the difference between kinetic and potential energies, and is more complicated to derive formally. A full discussion is given in Chapter 7 of Goldstein [1]; here we simply present the form appropriate to accelerator dynamics, vrz: 2 ]/2 L=—mcc—v·v-e>+eA·v[]q (9) We note that in the non-relativistic limit, v << c, this reduces almost to the form of Eq. (7), apart from a constant- mcz, which vanishes on subsequent differentiations. In the framework of Hamiltonian theory the importance of the Lagrangian lies in the formulation of Hamilton's Principle of Stationary Action (sometimes called "least action" which is usually but not always the case). This principle states that the action integral defined by: OCR Output S= |L(q,q,r)dt (10) is an extremum for the dynamically true path of the time trajectory between tl and tz, i.e. 5S= 5jLdt=O to first order The evaluation of this by the calculus of variations, which is very clearly explained in Vol. II of the Feynmann Lectures [3], results in the Lagrangian equations of motion for a conservative system: .¢i(g)_a dt Qqk Qqk (12, The quantity (QL/ Qqk) is known as the canonical momentum; in the simplest cases where there is no vector potential it is the same as the mechanical momentum, but the presence of an A · v term gives rise to a so—called magnetic momentum. One then recognises that Eq. (12) is Newtor1's Second Law in a more modern guise. Summarising, for a system of k degrees of freedom the Lagrangian is a function of 2k dynamical variables (generalised co-ordinates and velocities) and the "time" t, k of these variables (qk) being the "time" derivatives of the other k variables (qk). The corresponding Lagrange equations consist of a set of k second-order differential equations describing the motion of the system. The Hamiltonian formulation of mechanics describes a system in terms of generalised co ordinates (qk) and generalised momenta (pk the same as the canonical momenta we identified in Eq. (12) above. The change of basis from the set (qk,qk,z) to the set (qk, pk,t) is obtained through a Legendre transformation, defined by the function: H(<1k1¤k¢)=Zpk clk —L(q,r2,r) (13) We can take the differential of H on the left-hand side as QH QH QH (14) dH=—dt+** ———dq q}; + —dp 8g;<9Pk ak gl k or altematively, from the right-hand side of (13) as QL QL . QL . dH=ZPk dqk+`L<1k dP1t‘E‘d<lk’Z"T" dQk‘—’d’ (15) OCR Output Designating the canonical momenta of Eq. (12) as 8L Pk = -7- 1 ( 16) 8% the first and fourth summations cancel and the remaining terms can be identified with the corresponding terms in (14) 8H 8L (17) 8t 8: 8H k <2Pk 8H 3% The function H(q,p,t) is the Hamiltonian and Eqs. (18) and (19) are the Hamilton equations of motion. They are first order, 2k in number for k degrees of freedom, and show a remarkable symmetry of form between the generalised position co-ordinates qk and their conjugate momenta pk. This symmetry leads to very flexible transformation properties between sets of dynamical variables. For non-relativistic motion the Hamiltonian is often, though not necessarily, the sum of potential and kinetic energies H(q,p,t) =T +U (20) A relativistic Hamiltonian for a single particle in an electromagnetic field can be derived from the Lagrangian of Eq. (9). In Cartesian co-ordinates, k = x, y, z, the canonical momenta given by (16) are pk =m¤1’ vt +eAk (21) and differ from the component moyvk of the mechanical momentum by the contribution eAk, the electromagnetic momentum. The resulting Hamiltonian is easily shown to be 2 2 2 1/2 H(q,z>,r)=e¢+¢I(p-6A) +m1>¢ l » (22) where the pk have been expressed in vector form. 3 . SOME PROPERTIES OF THE HAMILTONIAN From the Hamiltonian H (qk,p k,t) the Hamilton equations of motion are obtained by . d 8H qk 2 i 2 _ (18) dt 8pk pk 2 FE1; 2 - EI. (19) OCR Output dt k The Poisson bracket of any two dynamical variables f (qk, pk;) and g(qk,pk,t) is defined by {Lg} (9 3 k aflk 6'Pk 6'Pk aqk (23) One sees that {f,f} E 0 {af} = -{f1s} The time derivative off d . é = %+ Z{%Qk + pk} k qk lc (24) + {,*,11} (from (18), (19) and (23) at In particular, iff = H we have dH BH (25) dt Qt If H does not depend explicitly on time it is a constant of the motion. Any invariant of the motion not containing t explicitly, has a vanishing Poisson bracket with H. 4 . CANONICAL TRANSFORMATIONS Canonical transformations are of considerable utility in simplifying problems by an advantageous choice of co—ordinate system. In particular they can be used to reduce the number of degrees of freedom of a system by exposing invariant quantities, or quantities that are "almost invariant" apart from a small parameter, permitting perturbation theory to be applied.
Recommended publications
  • Chapter 7. the Principle of Least Action
    Chapter 7. The Principle of Least Action 7.1 Force Methods vs. Energy Methods We have so far studied two distinct ways of analyzing physics problems: force methods, basically consisting of the application of Newton’s Laws, and energy methods, consisting of the application of the principle of conservation of energy (the conservations of linear and angular momenta can also be considered as part of this). Both have their advantages and disadvantages. More precisely, energy methods often involve scalar quantities (e.g., work, kinetic energy, and potential energy) and are thus easier to handle than forces, which are vectorial in nature. However, forces tell us more. The simple example of a particle subjected to the earth’s gravitation field will clearly illustrate this. We know that when a particle moves from position y0 to y in the earth’s gravitational field, the conservation of energy tells us that ΔK + ΔUgrav = 0 (7.1) 1 2 2 m v − v0 + mg(y − y0 ) = 0, 2 ( ) which implies that the final speed of the particle, of mass m , is 2 v = v0 + 2g(y0 − y). (7.2) Although we know the final speed v of the particle, given its initial speed v0 and the initial and final positions y0 and y , we do not know how its position and velocity (and its speed) evolve with time. On the other hand, if we apply Newton’s Second Law, then we write −mgey = maey dv (7.3) = m e . dt y This equation is easily manipulated to yield t v = dv ∫t0 t = − gdτ + v0 (7.4) ∫t0 = −g(t − t0 ) + v0 , - 138 - where v0 is the initial velocity (it appears in equation (7.4) as a constant of integration).
    [Show full text]
  • Classical Mechanics
    Classical Mechanics Hyoungsoon Choi Spring, 2014 Contents 1 Introduction4 1.1 Kinematics and Kinetics . .5 1.2 Kinematics: Watching Wallace and Gromit ............6 1.3 Inertia and Inertial Frame . .8 2 Newton's Laws of Motion 10 2.1 The First Law: The Law of Inertia . 10 2.2 The Second Law: The Equation of Motion . 11 2.3 The Third Law: The Law of Action and Reaction . 12 3 Laws of Conservation 14 3.1 Conservation of Momentum . 14 3.2 Conservation of Angular Momentum . 15 3.3 Conservation of Energy . 17 3.3.1 Kinetic energy . 17 3.3.2 Potential energy . 18 3.3.3 Mechanical energy conservation . 19 4 Solving Equation of Motions 20 4.1 Force-Free Motion . 21 4.2 Constant Force Motion . 22 4.2.1 Constant force motion in one dimension . 22 4.2.2 Constant force motion in two dimensions . 23 4.3 Varying Force Motion . 25 4.3.1 Drag force . 25 4.3.2 Harmonic oscillator . 29 5 Lagrangian Mechanics 30 5.1 Configuration Space . 30 5.2 Lagrangian Equations of Motion . 32 5.3 Generalized Coordinates . 34 5.4 Lagrangian Mechanics . 36 5.5 D'Alembert's Principle . 37 5.6 Conjugate Variables . 39 1 CONTENTS 2 6 Hamiltonian Mechanics 40 6.1 Legendre Transformation: From Lagrangian to Hamiltonian . 40 6.2 Hamilton's Equations . 41 6.3 Configuration Space and Phase Space . 43 6.4 Hamiltonian and Energy . 45 7 Central Force Motion 47 7.1 Conservation Laws in Central Force Field . 47 7.2 The Path Equation .
    [Show full text]
  • Dynamics of the Elastic Pendulum Qisong Xiao; Shenghao Xia ; Corey Zammit; Nirantha Balagopal; Zijun Li Agenda
    Dynamics of the Elastic Pendulum Qisong Xiao; Shenghao Xia ; Corey Zammit; Nirantha Balagopal; Zijun Li Agenda • Introduction to the elastic pendulum problem • Derivations of the equations of motion • Real-life examples of an elastic pendulum • Trivial cases & equilibrium states • MATLAB models The Elastic Problem (Simple Harmonic Motion) 푑2푥 푑2푥 푘 • 퐹 = 푚 = −푘푥 = − 푥 푛푒푡 푑푡2 푑푡2 푚 • Solve this differential equation to find 푥 푡 = 푐1 cos 휔푡 + 푐2 sin 휔푡 = 퐴푐표푠(휔푡 − 휑) • With velocity and acceleration 푣 푡 = −퐴휔 sin 휔푡 + 휑 푎 푡 = −퐴휔2cos(휔푡 + 휑) • Total energy of the system 퐸 = 퐾 푡 + 푈 푡 1 1 1 = 푚푣푡2 + 푘푥2 = 푘퐴2 2 2 2 The Pendulum Problem (with some assumptions) • With position vector of point mass 푥 = 푙 푠푖푛휃푖 − 푐표푠휃푗 , define 푟 such that 푥 = 푙푟 and 휃 = 푐표푠휃푖 + 푠푖푛휃푗 • Find the first and second derivatives of the position vector: 푑푥 푑휃 = 푙 휃 푑푡 푑푡 2 푑2푥 푑2휃 푑휃 = 푙 휃 − 푙 푟 푑푡2 푑푡2 푑푡 • From Newton’s Law, (neglecting frictional force) 푑2푥 푚 = 퐹 + 퐹 푑푡2 푔 푡 The Pendulum Problem (with some assumptions) Defining force of gravity as 퐹푔 = −푚푔푗 = 푚푔푐표푠휃푟 − 푚푔푠푖푛휃휃 and tension of the string as 퐹푡 = −푇푟 : 2 푑휃 −푚푙 = 푚푔푐표푠휃 − 푇 푑푡 푑2휃 푚푙 = −푚푔푠푖푛휃 푑푡2 Define 휔0 = 푔/푙 to find the solution: 푑2휃 푔 = − 푠푖푛휃 = −휔2푠푖푛휃 푑푡2 푙 0 Derivation of Equations of Motion • m = pendulum mass • mspring = spring mass • l = unstreatched spring length • k = spring constant • g = acceleration due to gravity • Ft = pre-tension of spring 푚푔−퐹 • r = static spring stretch, 푟 = 푡 s 푠 푘 • rd = dynamic spring stretch • r = total spring stretch 푟푠 + 푟푑 Derivation of Equations of Motion
    [Show full text]
  • The Celestial Mechanics of Newton
    GENERAL I ARTICLE The Celestial Mechanics of Newton Dipankar Bhattacharya Newton's law of universal gravitation laid the physical foundation of celestial mechanics. This article reviews the steps towards the law of gravi­ tation, and highlights some applications to celes­ tial mechanics found in Newton's Principia. 1. Introduction Newton's Principia consists of three books; the third Dipankar Bhattacharya is at the Astrophysics Group dealing with the The System of the World puts forth of the Raman Research Newton's views on celestial mechanics. This third book Institute. His research is indeed the heart of Newton's "natural philosophy" interests cover all types of which draws heavily on the mathematical results derived cosmic explosions and in the first two books. Here he systematises his math­ their remnants. ematical findings and confronts them against a variety of observed phenomena culminating in a powerful and compelling development of the universal law of gravita­ tion. Newton lived in an era of exciting developments in Nat­ ural Philosophy. Some three decades before his birth J 0- hannes Kepler had announced his first two laws of plan­ etary motion (AD 1609), to be followed by the third law after a decade (AD 1619). These were empirical laws derived from accurate astronomical observations, and stirred the imagination of philosophers regarding their underlying cause. Mechanics of terrestrial bodies was also being developed around this time. Galileo's experiments were conducted in the early 17th century leading to the discovery of the Keywords laws of free fall and projectile motion. Galileo's Dialogue Celestial mechanics, astronomy, about the system of the world was published in 1632.
    [Show full text]
  • Measuring Earth's Gravitational Constant with a Pendulum
    Measuring Earth’s Gravitational Constant with a Pendulum Philippe Lewalle, Tony Dimino PHY 141 Lab TA, Fall 2014, Prof. Frank Wolfs University of Rochester November 30, 2014 Abstract In this lab we aim to calculate Earth’s gravitational constant by measuring the period of a pendulum. We obtain a value of 9.79 ± 0.02m/s2. This measurement agrees with the accepted value of g = 9.81m/s2 to within the precision limits of our procedure. Limitations of the techniques and assumptions used to calculate these values are discussed. The pedagogical context of this example report for PHY 141 is also discussed in the final remarks. 1 Theory The gravitational acceleration g near the surface of the Earth is known to be approximately constant, disregarding small effects due to geological variations and altitude shifts. We aim to measure the value of that acceleration in our lab, by observing the motion of a pendulum, whose motion depends both on g and the length L of the pendulum. It is a well known result that a pendulum consisting of a point mass and attached to a massless rod of length L obeys the relationship shown in eq. (1), d2θ g = − sin θ, (1) dt2 L where θ is the angle from the vertical, as showng in Fig. 1. For small displacements (ie small θ), we make a small angle approximation such that sin θ ≈ θ, which yields eq. (2). d2θ g = − θ (2) dt2 L Eq. (2) admits sinusoidal solutions with an angular frequency ω. We relate this to the period T of oscillation, to obtain an expression for g in terms of the period and length of the pendulum, shown in eq.
    [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]
  • TOPICS in CELESTIAL MECHANICS 1. the Newtonian N-Body Problem
    TOPICS IN CELESTIAL MECHANICS RICHARD MOECKEL 1. The Newtonian n-body Problem Celestial mechanics can be defined as the study of the solution of Newton's differ- ential equations formulated by Isaac Newton in 1686 in his Philosophiae Naturalis Principia Mathematica. The setting for celestial mechanics is three-dimensional space: 3 R = fq = (x; y; z): x; y; z 2 Rg with the Euclidean norm: p jqj = x2 + y2 + z2: A point particle is characterized by a position q 2 R3 and a mass m 2 R+. A motion of such a particle is described by a curve q(t) where t runs over some interval in R; the mass is assumed to be constant. Some remarks will be made below about why it is reasonable to model a celestial body by a point particle. For every motion of a point particle one can define: velocity: v(t) =q _(t) momentum: p(t) = mv(t): Newton formulated the following laws of motion: Lex.I. Corpus omne perservare in statu suo quiescendi vel movendi uniformiter in directum, nisi quatenus a viribus impressis cogitur statum illum mutare 1 Lex.II. Mutationem motus proportionem esse vi motrici impressae et fieri secundem lineam qua vis illa imprimitur. 2 Lex.III Actioni contrarium semper et aequalem esse reactionem: sive corporum duorum actiones in se mutuo semper esse aequales et in partes contrarias dirigi. 3 The first law is statement of the principle of inertia. The second law asserts the existence of a force function F : R4 ! R3 such that: p_ = F (q; t) or mq¨ = F (q; t): In celestial mechanics, the dependence of F (q; t) on t is usually indirect; the force on one body depends on the positions of the other massive bodies which in turn depend on t.
    [Show full text]
  • The Development of the Action Principle a Didactic History from Euler-Lagrange to Schwinger Series: Springerbriefs in Physics
    springer.com Walter Dittrich The Development of the Action Principle A Didactic History from Euler-Lagrange to Schwinger Series: SpringerBriefs in Physics Provides a philosophical interpretation of the development of physics Contains many worked examples Shows the path of the action principle - from Euler's first contributions to present topics This book describes the historical development of the principle of stationary action from the 17th to the 20th centuries. Reference is made to the most important contributors to this topic, in particular Bernoullis, Leibniz, Euler, Lagrange and Laplace. The leading theme is how the action principle is applied to problems in classical physics such as hydrodynamics, electrodynamics and gravity, extending also to the modern formulation of quantum mechanics 1st ed. 2021, XV, 135 p. 23 illus., 1 illus. and quantum field theory, especially quantum electrodynamics. A critical analysis of operator in color. versus c-number field theory is given. The book contains many worked examples. In particular, the term "vacuum" is scrutinized. The book is aimed primarily at actively working researchers, Printed book graduate students and historians interested in the philosophical interpretation and evolution of Softcover physics; in particular, in understanding the action principle and its application to a wide range 54,99 € | £49.99 | $69.99 of natural phenomena. [1]58,84 € (D) | 60,49 € (A) | CHF 65,00 eBook 46,00 € | £39.99 | $54.99 [2]46,00 € (D) | 46,00 € (A) | CHF 52,00 Available from your library or springer.com/shop MyCopy [3] Printed eBook for just € | $ 24.99 springer.com/mycopy Order online at springer.com / or for the Americas call (toll free) 1-800-SPRINGER / or email us at: [email protected].
    [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]
  • Perturbation Theory in Celestial Mechanics
    Perturbation Theory in Celestial Mechanics Alessandra Celletti Dipartimento di Matematica Universit`adi Roma Tor Vergata Via della Ricerca Scientifica 1, I-00133 Roma (Italy) ([email protected]) December 8, 2007 Contents 1 Glossary 2 2 Definition 2 3 Introduction 2 4 Classical perturbation theory 4 4.1 The classical theory . 4 4.2 The precession of the perihelion of Mercury . 6 4.2.1 Delaunay action–angle variables . 6 4.2.2 The restricted, planar, circular, three–body problem . 7 4.2.3 Expansion of the perturbing function . 7 4.2.4 Computation of the precession of the perihelion . 8 5 Resonant perturbation theory 9 5.1 The resonant theory . 9 5.2 Three–body resonance . 10 5.3 Degenerate perturbation theory . 11 5.4 The precession of the equinoxes . 12 6 Invariant tori 14 6.1 Invariant KAM surfaces . 14 6.2 Rotational tori for the spin–orbit problem . 15 6.3 Librational tori for the spin–orbit problem . 16 6.4 Rotational tori for the restricted three–body problem . 17 6.5 Planetary problem . 18 7 Periodic orbits 18 7.1 Construction of periodic orbits . 18 7.2 The libration in longitude of the Moon . 20 1 8 Future directions 20 9 Bibliography 21 9.1 Books and Reviews . 21 9.2 Primary Literature . 22 1 Glossary KAM theory: it provides the persistence of quasi–periodic motions under a small perturbation of an integrable system. KAM theory can be applied under quite general assumptions, i.e. a non– degeneracy of the integrable system and a diophantine condition of the frequency of motion.
    [Show full text]
  • The Celestial Mechanics Approach: Theoretical Foundations
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by RERO DOC Digital Library J Geod (2010) 84:605–624 DOI 10.1007/s00190-010-0401-7 ORIGINAL ARTICLE The celestial mechanics approach: theoretical foundations Gerhard Beutler · Adrian Jäggi · Leoš Mervart · Ulrich Meyer Received: 31 October 2009 / Accepted: 29 July 2010 / Published online: 24 August 2010 © Springer-Verlag 2010 Abstract Gravity field determination using the measure- of the CMA, in particular to the GRACE mission, may be ments of Global Positioning receivers onboard low Earth found in Beutler et al. (2010) and Jäggi et al. (2010b). orbiters and inter-satellite measurements in a constellation of The determination of the Earth’s global gravity field using satellites is a generalized orbit determination problem involv- the data of space missions is nowadays either based on ing all satellites of the constellation. The celestial mechanics approach (CMA) is comprehensive in the sense that it encom- 1. the observations of spaceborne Global Positioning (GPS) passes many different methods currently in use, in particular receivers onboard low Earth orbiters (LEOs) (see so-called short-arc methods, reduced-dynamic methods, and Reigber et al. 2004), pure dynamic methods. The method is very flexible because 2. or precise inter-satellite distance monitoring of a close the actual solution type may be selected just prior to the com- satellite constellation using microwave links (combined bination of the satellite-, arc- and technique-specific normal with the measurements of the GPS receivers on all space- equation systems. It is thus possible to generate ensembles crafts involved) (see Tapley et al.
    [Show full text]