PHYS 705: Classical Mechanics Hamilton-Jacobi Equation 2

Total Page:16

File Type:pdf, Size:1020Kb

PHYS 705: Classical Mechanics Hamilton-Jacobi Equation 2 1 PHYS 705: Classical Mechanics Hamilton-Jacobi Equation 2 Hamilton-Jacobi Equation There is also a very elegant relation between the Hamiltonian Formulation of Mechanics and Quantum Mechanics. To do that, we need to derive the Hamilton-Jacobi equation. The Hamilton-Jacobi equation also represents a very general method in solving mechanical problems. To start… Let say we are able to find a canonical transformation taking our 2n phase q, p space variables i i directly to 2n constants of motion, i , i i.e., Qi i Pi i 3 Hamilton-Jacobi Equation One sufficient condition to ensure that our new variables are constant in time is that the transformed Hamiltonian K shall be identically zero. If that is the case, the equations of motion will be, K Qi 0 Pi K Pi 0 Qi Recall that we have the new and old Hamiltonian, K and H, relating through the generating function F by: F K H t 4 Hamilton-Jacobi Equation With K identically equals to zero, we can write: F H qpt, , 0 (*) t For convenience in later calculations, we shall take F to be of Type 2 so that its independent variables are the old coordinates q i and the new momenta P i . In the same notations as previously, we can write FqQPt ,,, FqPt2 (,,) QPi i Now, in order to write H in terms of this chosen set of variables qPt ,, , we will replace all the p i in H using the transformation equation: F2 qPt,, pi qi 5 Hamilton-Jacobi Equation F F Then, one can formally rewrite Eq. (*) as: note that 2 since t t FqPt ,, FqPt2 (,,) QPi i F2 F 2 F 2 Hqq1, ,n ; , , ; t 0 q1 qn t Fq ,, t p 2 This PDE is known as the Hamilton-Jacobi Equation. i qi Notes: -Since P i i are constants, the HJ equation constitutes a partial differential equation of (n+1) independent variables: q1,,, qtn - We used the fact that the new P’s and Q’s are constants but we have not specified in how to determine them yet. - It is customary to denote the solution F 2 by S and called it the Hamilton’s Principal Function. 6 Hamilton-Jacobi Equation Suppose we are able to find a solution to this 1st order partial differential equation in (n+1) variables… FSSqq2 1,,,n 1,, n ; t where we have explicitly write out the constant new momenta: Pi i Recall that for a Type 2 generating function, we have the following partial derivatives describing the canonical transformation: Sq ,, t S pi pi ( T 1) qi qi Sq,, t S Q Qi i ( T 2) i Pi i 7 Hamilton-Jacobi Equation After explicitly taking the partial derivative on the RHS of Eq. (T1) Sq ,, t pi qi and evaluating them at the initial time t 0 , we will have n equations that one can invert to solve for the n unknown constants i in terms of the initial conditions qp 0, 0 , t 0 , i.e., i i qp0, 0, t 0 Similarly, by explicitly evaluating the partial derivatives on the RHS of Eq. (T2) at time t 0 , we obtain the other n constants of motion i Sq ,, t Qi i i qq0, tt 0 8 Hamilton-Jacobi Equation With all 2n constants of motion i , i solved, we can now use Eq. (T2) again to solve for q i in terms of the i , i at a later time t. Sq ,, t qqi i ,, t i i p Then, with ii , , and q i known, we can use Eq. (T1) again to evaluate i in terms of i , i at a later time t. Sq ,,,, t t p pp,, t i q i i i The two boxed equations constitute the desired complete solutions of the Hamilton equations of motion. 9 Hamilton Principal Function and Action Now, let consider the total time derivative of the Hamilton Principal function S, dSq ,, t SS qi dt qi t Using Eq. (T1) and the Hamilton-Jacobi equation itself, we can replace the two partial derivatives on the right by p i and H, S S dS pi H 0 t pqi i H L qi dt So, the Hamilton Principal Function is differed at most from the action by a constant, i.e., S Ldt constant 10 Example: Solving Harmonic Oscillator with HJ m Recall, we have: 1 2 0 x H pmqE2 2 2 2 k m kq2 2m f( q ) kq U( q ) 2 The Hamilton-Jacobi equation gives: 2 1 SS SS m2 2 q 2 0 Hq, ; t 0 q t 2mq t Recall that under this scheme, we have: K 0 Q and P 11 Example: Solving Harmonic Oscillator with HJ m 1 H pmqE2 2 2 2 2 k m 2m 0 x kq2 f( q ) kq U( q ) When H does not explicitly depend on time, 2 we can generally try this trial solution for S: SqtWq ,,, t (q and t are separable) With this trial solution, we have S t and the HJ equation becomes: 2 2 1 SS 2 2 2 1 W 2 2 2 recall: LHS H m q 0 m q 2mq t 2m q so H 12 Example: Solving Harmonic Oscillator with HJ m The previous equation can be immediately integrated: 2 2 1 W W 2 2 2 m2 2 q 2 mq 2 m q 2m q 2 W 2 2 2 2m m q q 2 2 2 W 2 m m2 2 q 2 dq and S 2 m m q dq t 13 Example: Solving Harmonic Oscillator with HJ m We don’t need to integrate yet since we only need SS to use , in the transformation eqs: q Sq ,, t S 2 2 2 #1 eq: i Q 2 m m q dq t i 1 dq m dq 2m t t 2 2m m2 2 q 2 2 1 m 2 q 2 2 m Letting x q , we can rewrite the integral as: 2 1 dx 1 1 m t tarcsin x arcsin q 2 1 x 2 14 Example: Solving Harmonic Oscillator with HJ m 1 1 m tarcsin x arcsin q 2 Inverting the arcsin, we have 2 qt sin t ' with ' m 2 Sq ,,,, t t nd p The 2 transformation equation gives, i q i S #2 eq: p2 m m2 2 q 2 dq t 2m m2 2 q 2 q q 15 Example: Solving Harmonic Oscillator with HJ Substituting , we have, m q t m 2 2 pt2 m 1 sin2 t ' 2 2 m pt2 m 1sin 2 t ' with ' pt2 m cos2 t ' pt 2 m cos t ' 16 Example: Solving Harmonic Oscillator with HJ 2 m qt sin t ' m 2 pt 2 m cos t ' Completing the calculations, we still need to link the two constants of motion , to initial conditions: @ t0, q q0 , p p 0 Recall that the constant of motion is H=E of the system so that we can calculate this by squaring q and p at time t 0 : 1 2 2 2 2 p0 m q 0 2m q 1 And, we can calculate at time t 0 by dividing the 2 eqs: 0 tan ' p0 m 17 Another example in using the Hamilton-Jacobi Method (t-dependent H) (G 10.8) Suppose we are given a Hamiltonian to a system as, p2 H mAtq where A is a constant 2m Our task is to solve for the equation of motion by means of the Hamilton’s principle function S with the initial conditions t0, q0 0, pmv 0 0 The Hamilton-Jacobi equation for S is: SS Hq; ; t 0 q t 18 Example in using the Hamilton-Jacobi Method Recall that the Hamilton’s principle function S is a Type 2 generating function with independent variables qPt,, S SqPt ,, with the condition that the canonically transformed variables are constants, i.e., Qand P For a Type 2 generating function, we have the following partial derivatives for S: Sq ,, t Sq ,, t p ( T 1) Q ( T 2) q 19 Example in using the Hamilton-Jacobi Method Writing H out explicitly in the Hamilton-Jacobi equation, we have 2 1 S S mAtq 0 (*) 2mq t S Hq; ; t q Here, we assume a solution to be separatable of the form: Sq ,;, t f tqgt Substituting this into the above Eq (*), we have, 1 2 f, t mAtq f ' , tqgt ' 0 ( ' is time 2m derivative here) 20 Example in using the Hamilton-Jacobi Method 1 2 f, t mAtq f ',t qg' t 0 2m Concentrating on the q dependent terms, they have to be independently add up to zero… So, we have, mAt 2 f', t mAt ft, f 2 0 And, requiring the remaining two terms adding up to zero also, we have 1 gt' ft2 2m Substituting f (t) from above, we have 22 2 2 4 1mAt 1 mAt 2 2 gt' f0 mAtff 0 0 2m 2 2 m 4 21 Example in using the Hamilton-Jacobi Method Integrating wrt time on both sides, we then have, mA2 t 5 Af f 2 g t 0 t3 0 t g 40 6 2m 0 Putting both f (t) and g(t) back into the Hamilton’s principle function, we have, 2 2 5 2 mAt mAt Af0 3 f 0 S q, ; t f0 q t t g 0 2 40 6 2m Since the Hamilton-Jacobi Equation only involves partial derivatives of S, g 0 can be taken to be zero without affect the dynamics and for simplicity, we will take the integration constant f 0 to be simply , i.e., f0 2 2 5 2 mAt mAt A3 Sqt, ; q tt 2 40 6 2m 22 Example in using the Hamilton-Jacobi Method mAt2 mA2t 5 A 2 We can then solve for using Eq.
Recommended publications
  • Relativistic Dynamics
    Chapter 4 Relativistic dynamics We have seen in the previous lectures that our relativity postulates suggest that the most efficient (lazy but smart) approach to relativistic physics is in terms of 4-vectors, and that velocities never exceed c in magnitude. In this chapter we will see how this 4-vector approach works for dynamics, i.e., for the interplay between motion and forces. A particle subject to forces will undergo non-inertial motion. According to Newton, there is a simple (3-vector) relation between force and acceleration, f~ = m~a; (4.0.1) where acceleration is the second time derivative of position, d~v d2~x ~a = = : (4.0.2) dt dt2 There is just one problem with these relations | they are wrong! Newtonian dynamics is a good approximation when velocities are very small compared to c, but outside of this regime the relation (4.0.1) is simply incorrect. In particular, these relations are inconsistent with our relativity postu- lates. To see this, it is sufficient to note that Newton's equations (4.0.1) and (4.0.2) predict that a particle subject to a constant force (and initially at rest) will acquire a velocity which can become arbitrarily large, Z t ~ d~v 0 f ~v(t) = 0 dt = t ! 1 as t ! 1 . (4.0.3) 0 dt m This flatly contradicts the prediction of special relativity (and causality) that no signal can propagate faster than c. Our task is to understand how to formulate the dynamics of non-inertial particles in a manner which is consistent with our relativity postulates (and then verify that it matches observation, including in the non-relativistic regime).
    [Show full text]
  • OCC D 5 Gen5d Eee 1305 1A E
    this cover and their final version of the extended essay to is are not is chose to write about applications of differential calculus because she found a great interest in it during her IB Math class. She wishes she had time to complete a deeper analysis of her topic; however, her busy schedule made it difficult so she is somewhat disappointed with the outcome of her essay. It was a pleasure meeting with when she was able to and her understanding of her topic was evident during our viva voce. I, too, wish she had more time to complete a more thorough investigation. Overall, however, I believe she did well and am satisfied with her essay. must not use Examiner 1 Examiner 2 Examiner 3 A research 2 2 D B introduction 2 2 c 4 4 D 4 4 E reasoned 4 4 D F and evaluation 4 4 G use of 4 4 D H conclusion 2 2 formal 4 4 abstract 2 2 holistic 4 4 Mathematics Extended Essay An Investigation of the Various Practical Uses of Differential Calculus in Geometry, Biology, Economics, and Physics Candidate Number: 2031 Words 1 Abstract Calculus is a field of math dedicated to analyzing and interpreting behavioral changes in terms of a dependent variable in respect to changes in an independent variable. The versatility of differential calculus and the derivative function is discussed and highlighted in regards to its applications to various other fields such as geometry, biology, economics, and physics. First, a background on derivatives is provided in regards to their origin and evolution, especially as apparent in the transformation of their notations so as to include various individuals and ways of denoting derivative properties.
    [Show full text]
  • Time-Derivative Models of Pavlovian Reinforcement Richard S
    Approximately as appeared in: Learning and Computational Neuroscience: Foundations of Adaptive Networks, M. Gabriel and J. Moore, Eds., pp. 497–537. MIT Press, 1990. Chapter 12 Time-Derivative Models of Pavlovian Reinforcement Richard S. Sutton Andrew G. Barto This chapter presents a model of classical conditioning called the temporal- difference (TD) model. The TD model was originally developed as a neuron- like unit for use in adaptive networks (Sutton and Barto 1987; Sutton 1984; Barto, Sutton and Anderson 1983). In this paper, however, we analyze it from the point of view of animal learning theory. Our intended audience is both animal learning researchers interested in computational theories of behavior and machine learning researchers interested in how their learning algorithms relate to, and may be constrained by, animal learning studies. For an exposition of the TD model from an engineering point of view, see Chapter 13 of this volume. We focus on what we see as the primary theoretical contribution to animal learning theory of the TD and related models: the hypothesis that reinforcement in classical conditioning is the time derivative of a compos- ite association combining innate (US) and acquired (CS) associations. We call models based on some variant of this hypothesis time-derivative mod- els, examples of which are the models by Klopf (1988), Sutton and Barto (1981a), Moore et al (1986), Hawkins and Kandel (1984), Gelperin, Hop- field and Tank (1985), Tesauro (1987), and Kosko (1986); we examine several of these models in relation to the TD model. We also briefly ex- plore relationships with animal learning theories of reinforcement, including Mowrer’s drive-induction theory (Mowrer 1960) and the Rescorla-Wagner model (Rescorla and Wagner 1972).
    [Show full text]
  • 1 the Basic Set-Up 2 Poisson Brackets
    MATHEMATICS 7302 (Analytical Dynamics) YEAR 2016–2017, TERM 2 HANDOUT #12: THE HAMILTONIAN APPROACH TO MECHANICS These notes are intended to be read as a supplement to the handout from Gregory, Classical Mechanics, Chapter 14. 1 The basic set-up I assume that you have already studied Gregory, Sections 14.1–14.4. The following is intended only as a succinct summary. We are considering a system whose equations of motion are written in Hamiltonian form. This means that: 1. The phase space of the system is parametrized by canonical coordinates q =(q1,...,qn) and p =(p1,...,pn). 2. We are given a Hamiltonian function H(q, p, t). 3. The dynamics of the system is given by Hamilton’s equations of motion ∂H q˙i = (1a) ∂pi ∂H p˙i = − (1b) ∂qi for i =1,...,n. In these notes we will consider some deeper aspects of Hamiltonian dynamics. 2 Poisson brackets Let us start by considering an arbitrary function f(q, p, t). Then its time evolution is given by n df ∂f ∂f ∂f = q˙ + p˙ + (2a) dt ∂q i ∂p i ∂t i=1 i i X n ∂f ∂H ∂f ∂H ∂f = − + (2b) ∂q ∂p ∂p ∂q ∂t i=1 i i i i X 1 where the first equality used the definition of total time derivative together with the chain rule, and the second equality used Hamilton’s equations of motion. The formula (2b) suggests that we make a more general definition. Let f(q, p, t) and g(q, p, t) be any two functions; we then define their Poisson bracket {f,g} to be n def ∂f ∂g ∂f ∂g {f,g} = − .
    [Show full text]
  • Time-Dependent Hamiltonian Mechanics on a Locally Conformal
    Time-dependent Hamiltonian mechanics on a locally conformal symplectic manifold Orlando Ragnisco†, Cristina Sardón∗, Marcin Zając∗∗ Department of Mathematics and Physics†, Universita degli studi Roma Tre, Largo S. Leonardo Murialdo, 1, 00146 , Rome, Italy. ragnisco@fis.uniroma3.it Department of Applied Mathematics∗, Universidad Polit´ecnica de Madrid. C/ Jos´eGuti´errez Abascal, 2, 28006, Madrid. Spain. [email protected] Department of Mathematical Methods in Physics∗∗, Faculty of Physics. University of Warsaw, ul. Pasteura 5, 02-093 Warsaw, Poland. [email protected] Abstract In this paper we aim at presenting a concise but also comprehensive study of time-dependent (t- dependent) Hamiltonian dynamics on a locally conformal symplectic (lcs) manifold. We present a generalized geometric theory of canonical transformations and formulate a time-dependent geometric Hamilton-Jacobi theory on lcs manifolds. In contrast to previous papers concerning locally conformal symplectic manifolds, here the introduction of the time dependency brings out interesting geometric properties, as it is the introduction of contact geometry in locally symplectic patches. To conclude, we show examples of the applications of our formalism, in particular, we present systems of differential equations with time-dependent parameters, which admit different physical interpretations as we shall point out. arXiv:2104.02636v1 [math-ph] 6 Apr 2021 Contents 1 Introduction 2 2 Fundamentals on time-dependent Hamiltonian systems 5 2.1 Time-dependentsystems. ....... 5 2.2 Canonicaltransformations . ......... 6 2.3 Generating functions of canonical transformations . ................ 8 1 3 Geometry of locally conformal symplectic manifolds 8 3.1 Basics on locally conformal symplectic manifolds . ............... 8 3.2 Locally conformal symplectic structures on cotangent bundles............
    [Show full text]
  • Lie Time Derivative £V(F) of a Spatial Field F
    Lie time derivative $v(f) of a spatial field f • A way to obtain an objective rate of a spatial tensor field • Can be used to derive objective Constitutive Equations on rate form D −1 Definition: $v(f) = χ? Dt (χ? (f)) Procedure in 3 steps: 1. Pull-back of the spatial tensor field,f, to the Reference configuration to obtain the corresponding material tensor field, F. 2. Take the material time derivative on the corresponding material tensor field, F, to obtain F_ . _ 3. Push-forward of F to the Current configuration to obtain $v(f). D −1 Important!|Note that the material time derivative, i.e. Dt (χ? (f)) is executed in the Reference configuration (rotation neutralized). Recall that D D χ−1 (f) = (F) = F_ = D F Dt ?(2) Dt v d D F = F(X + v) v d and hence, $v(f) = χ? (DvF) Thus, the Lie time derivative of a spatial tensor field is the push-forward of the directional derivative of the corresponding material tensor field in the direction of v (velocity vector). More comments on the Lie time derivative $v() • Rate constitutive equations must be formulated based on objective rates of stresses and strains to ensure material frame-indifference. • Rates of material tensor fields are by definition objective, since they are associated with a frame in a fixed linear space. • A spatial tensor field is said to transform objectively under superposed rigid body motions if it transforms according to standard rules of tensor analysis, e.g. A+ = QAQT (preserves distances under rigid body rotations).
    [Show full text]
  • Canonical Transformations (Lecture 4)
    Canonical transformations (Lecture 4) January 26, 2016 61/441 Lecture outline We will introduce and discuss canonical transformations that conserve the Hamiltonian structure of equations of motion. Poisson brackets are used to verify that a given transformation is canonical. A practical way to devise canonical transformation is based on usage of generation functions. The motivation behind this study is to understand the freedom which we have in the choice of various sets of coordinates and momenta. Later we will use this freedom to select a convenient set of coordinates for description of partilcle's motion in an accelerator. 62/441 Introduction Within the Lagrangian approach we can choose the generalized coordinates as we please. We can start with a set of coordinates qi and then introduce generalized momenta pi according to Eqs. @L(qk ; q_k ; t) pi = ; i = 1;:::; n ; @q_i and form the Hamiltonian ! H = pi q_i - L(qk ; q_k ; t) : i X Or, we can chose another set of generalized coordinates Qi = Qi (qk ; t), express the Lagrangian as a function of Qi , and obtain a different set of momenta Pi and a different Hamiltonian 0 H (Qi ; Pi ; t). This type of transformation is called a point transformation. The two representations are physically equivalent and they describe the same dynamics of our physical system. 63/441 Introduction A more general approach to the problem of using various variables in Hamiltonian formulation of equations of motion is the following. Let us assume that we have canonical variables qi , pi and the corresponding Hamiltonian H(qi ; pi ; t) and then make a transformation to new variables Qi = Qi (qk ; pk ; t) ; Pi = Pi (qk ; pk ; t) : i = 1 ::: n: (4.1) 0 Can we find a new Hamiltonian H (Qi ; Pi ; t) such that the system motion in new variables satisfies Hamiltonian equations with H 0? What are requirements on the transformation (4.1) for such a Hamiltonian to exist? These questions lead us to canonical transformations.
    [Show full text]
  • Classical Mechanics Examples (Canonical Transformation)
    Classical Mechanics Examples (Canonical Transformation) Dipan Kumar Ghosh Centre for Excellence in Basic Sciences Kalina, Mumbai 400098 November 10, 2016 1 Introduction In classical mechanics, there is no unique prescription for one to choose the generalized coordinates for a problem. As long as the coordinates and the corresponding momenta span the entire phase space, it becomes an acceptable set. However, it turns out in prac- tice that some choices are better than some others as they make a given problem simpler while still preserving the form of Hamilton's equations. Going over from one set of chosen coordinates and momenta to another set which satisfy Hamilton's equations is done by canonical transformation. For instance, if we consider the central force problem in two dimensions and choose Carte- k sian coordinate, the potential is − . However, if we choose (r; θ) coordinates, px2 + y2 k the potential is − and θ is cyclic, which greatly simplifies the problem. The number of r cyclic coordinates in a problem may depend on the choice of generalized coordinates. A cyclic coordinate results in a constant of motion which is its conjugate momentum. The example above where we replace one set of coordinates by another set is known as point transformation. In the Hamiltonian formalism, the coordinates and momenta are given equal status and the dynamics occurs in what we know as the phase space. In dealing with phase space dynamics, we need point transformations in the phase space where the new coordinates (Qi) and the new momenta (Pi) are functions of old coordinates (pi) and old momenta (pi): Qi = Qi(fqjg; fpjg; t) Pi = Pi(fqjg; fpjg; t) 1 c D.
    [Show full text]
  • PHYS 705: Classical Mechanics Canonical Transformation 2
    1 PHYS 705: Classical Mechanics Canonical Transformation 2 Canonical Variables and Hamiltonian Formalism As we have seen, in the Hamiltonian Formulation of Mechanics, qj, p j are independent variables in phase space on equal footing The Hamilton’s Equation for q j , p j are “symmetric” (symplectic, later) H H qj and p j pj q j This elegant formal structure of mechanics affords us the freedom in selecting other appropriate canonical variables as our phase space “coordinates” and “momenta” - As long as the new variables formally satisfy this abstract structure (the form of the Hamilton’s Equations. 3 Canonical Transformation Recall (from hw) that the Euler-Lagrange Equation is invariant for a point transformation: Qj Qqt j ( ,) L d L i.e., if we have, 0, qj dt q j L d L then, 0, Qj dt Q j Now, the idea is to find a generalized (canonical) transformation in phase space (not config. space) such that the Hamilton’s Equations are invariant ! Qj Qqpt j (, ,) (In general, we look for transformations which Pj Pqpt j (, ,) are invertible.) 4 Invariance of EL equation for Point Transformation First look at the situation in config. space first: dL L Given: 0, and a point transformation: Q Qqt( ,) j j dt q j q j dL L 0 Need to show: dt Q j Q j L Lqi L q i Formally, calculate: (chain rule) Qji qQ ij i qQ ij L L q L q i i Q j i qiQ j i q i Q j From the inverse point transformation equation q i qQt i ( ,) , we have, qi q q q q 0 and i i qi Q i Q i k j Q j Qj k Qk t 5 Invariance of EL equation
    [Show full text]
  • Solutions to Assignments 05
    Solutions to Assignments 05 1. It is useful to first recall how this works in the case of classical mechanics (i.e. a 0+1 dimensional “field theory”). Consider a Lagrangian L(q; q_; t) that is a total time-derivative, i.e. d @F @F L(q; q_; t) = F (q; t) = + q_(t) : (1) dt @t @q(t) Then one has @L @2F @2F = + q_(t) (2) @q(t) @q(t)@t @q(t)2 and @L @F d @L @2F @2F = ) = + q_(t) (3) @q_(t) @q(t) dt @q_(t) @t@q(t) @q(t)2 Therefore one has @L d @L = identically (4) @q(t) dt @q_(t) Now we consider the field theory case. We define d @W α @W α @φ(x) L := W α(φ; x) = + dxα @xα @φ(x) @xα α α = @αW + @φW @αφ (5) to compute the Euler-Lagrange equations for it. One gets @L = @ @ W α + @2W α@ φ (6) @φ φ α φ α d @L d α β β = β @φW δα dx @(@βφ) dx β 2 β = @β@φW + @φW @βφ : (7) Thus (since partial derivatives commute) the Euler-Lagrange equations are satis- fied identically. Remark: One can also show the converse: if a Lagrangian L gives rise to Euler- Lagrange equations that are identically satisfied then (locally) the Lagrangian is a total derivative. The proof is simple. Assume that L(q; q_; t) satisfies @L d @L ≡ (8) @q dt @q_ identically. The left-hand side does evidently not depend on the acceleration q¨.
    [Show full text]
  • Hamilton-Jacobi Theory for Hamiltonian and Non-Hamiltonian Systems
    JOURNAL OF GEOMETRIC MECHANICS doi:10.3934/jgm.2020024 c American Institute of Mathematical Sciences Volume 12, Number 4, December 2020 pp. 563{583 HAMILTON-JACOBI THEORY FOR HAMILTONIAN AND NON-HAMILTONIAN SYSTEMS Sergey Rashkovskiy Ishlinsky Institute for Problems in Mechanics of the Russian Academy of Sciences Vernadskogo Ave., 101/1 Moscow, 119526, Russia (Communicated by David Mart´ınde Diego) Abstract. A generalization of the Hamilton-Jacobi theory to arbitrary dy- namical systems, including non-Hamiltonian ones, is considered. The general- ized Hamilton-Jacobi theory is constructed as a theory of ensemble of identi- cal systems moving in the configuration space and described by the continual equation of motion and the continuity equation. For Hamiltonian systems, the usual Hamilton-Jacobi equations naturally follow from this theory. The pro- posed formulation of the Hamilton-Jacobi theory, as the theory of ensemble, allows interpreting in a natural way the transition from quantum mechanics in the Schr¨odingerform to classical mechanics. 1. Introduction. In modern physics, the Hamilton-Jacobi theory occupies a spe- cial place. On the one hand, it is one of the formulations of classical mechanics [8,5, 12], while on the other hand it is the Hamilton-Jacobi theory that is the clas- sical limit into which quantum (wave) mechanics passes at ~ ! 0. In particular, the Schr¨odingerequation describing the motion of quantum particles, in the limit ~ ! 0, splits into two real equations: the classical Hamilton-Jacobi equation and the continuity equation for some ensemble of classical particles [9]. Thus, it is the Hamilton-Jacobi theory that is the bridge that connects the classical mechanics of point particles with the wave mechanics of quantum objects.
    [Show full text]
  • An Essay on Lagrangian and Eulerian Kinematics of Fluid Flow
    Lagrangian and Eulerian Representations of Fluid Flow: Kinematics and the Equations of Motion James F. Price Woods Hole Oceanographic Institution, Woods Hole, MA, 02543 [email protected], http://www.whoi.edu/science/PO/people/jprice June 7, 2006 Summary: This essay introduces the two methods that are widely used to observe and analyze fluid flows, either by observing the trajectories of specific fluid parcels, which yields what is commonly termed a Lagrangian representation, or by observing the fluid velocity at fixed positions, which yields an Eulerian representation. Lagrangian methods are often the most efficient way to sample a fluid flow and the physical conservation laws are inherently Lagrangian since they apply to moving fluid volumes rather than to the fluid that happens to be present at some fixed point in space. Nevertheless, the Lagrangian equations of motion applied to a three-dimensional continuum are quite difficult in most applications, and thus almost all of the theory (forward calculation) in fluid mechanics is developed within the Eulerian system. Lagrangian and Eulerian concepts and methods are thus used side-by-side in many investigations, and the premise of this essay is that an understanding of both systems and the relationships between them can help form the framework for a study of fluid mechanics. 1 The transformation of the conservation laws from a Lagrangian to an Eulerian system can be envisaged in three steps. (1) The first is dubbed the Fundamental Principle of Kinematics; the fluid velocity at a given time and fixed position (the Eulerian velocity) is equal to the velocity of the fluid parcel (the Lagrangian velocity) that is present at that position at that instant.
    [Show full text]