Hamiltonian Field Theory

Total Page:16

File Type:pdf, Size:1020Kb

Hamiltonian Field Theory Hamiltonian Field Theory August 31, 2016 1 Introduction So far we have treated classical field theory using Lagrangian and an action principle for Lagrangian. This approach is called Lagrangian field theory and is best suited for quantizing a field theory using the path- integral (aka functional integral) approach. However, a quicker route to quantizing a classical field theory is thru the \canonical quantization" program which is what you will find in most introductory treatments on quantum field theory. So for us, the Lagrangian is just a crutch to extract the Hamiltonian. However there is a price to pay when we switch from a Lagrangian approach to a Hamiltonian approach, that is manifest Lorentz invariance. Recall that a field theory is a system where we have a degree of freedom located at each point in space. If this degree of freedom is a Lorentz scalar, we will denote the degree of freedom located at position x by Φ(x) and its conjugate momenta, Π(x). These degrees of freedom of course fluctuate in time, so when we include this time-dependence, the notation becomes, Φ(x; t) and Π(x; t): The job of Hamiltonian field theory is describe time-evolution of Π and Φ in equations as follows, Π_ = :::; Φ_ = ::: The dot is time derivative, i.e. we have to choose a time axis or equivalently preferred reference frame which breaks manifest Lorentz invariance. Recall that earlier, in the Lagrangian/action approach we did not have to choose a time-frame because the equation of motion looked Lorentz invariant, e.g. the Euler-Lagrange equation for scalar field, @L @L @µ − = 0: @ (@µΦ) @Φ Here all Lorentz indices are contracted and the only quantities that appear are Lagrange densityL and Φ which are both Lorentz invariants/scalars - so the whole equation is manifestly Lorentz invariant. Similarly the action functional approach leads to a equation of motion of the schematic form, δI = 0; δΦ(x) which is also manifestly Lorentz invariant as both the numerator and denominator are Lorentz invariants themselves. In this note we review Hamiltonian field theory with an eye towards canonical quantization of classical fields. Since in most texts in classical mechanics, the dependence on time for the canonical pair of variables p and q are omitted from most formulas, we too shall choose not to display the time dependence, i.e. instead of Φ(x; t) we will use, Φ(x) and instead of Π(x; t) we will use, Π(x). Just as in case of q; p it would be implicitly understood that Φ(x); Π(x) are functions of time. 1 2 Hamilton's principle and Hamilton's equation for field theory Hamilton's principle for classical mechanics states that the equations of motion of a physical system described by the generalized coordinate and momenta, (q; p), can be obtain by extremizing (i.e. setting the first order variation to zero) of the following functional, I = dt [pq_ − H(p; q)] : (1) ˆ The function, H(p; q) is called the Hamiltonian function. One can easily check by varying the above action that the equation of motion for this system are, @H q_ = ; @p @H p_ = − : (2) @q These equations are the well known Hamilton's equation. Note that in contrast with the Lagrangian ap- proach where the Euler-Lagrange equations involve second order time-derivatives in time, the equations of motion in the Hamilton's approach are first order differential equations in time. Of course the price to pay is that we have twice the number of equations, one set for q and one set for p. One can easily generalize Hamilton's framework for classical mechanics for a single degree of freedom to field theory i.e. infinite degrees of freedom. To accomplish that first we write down the action for N number of degrees of freedom, (qi; pi); i = 1; 2; : : : ; N: The action and Hamilton's equation for this case involving N degrees of freedom are: " N ! # X I = dt p q_ − H(p ; q ) (3) ˆ i i i i i=1 and, @H q_i = ; @pi @H p_i = − : (4) @qi Now in order to go to a continuum limit, we set N ! 1, then the discrete label, i turns into a contin- uum label such as the position coordinates, x = (x; y; z) (which are real number/continuous valued not discrete/integer valued), qi ! Φ(x); pi ! Π(x): The summation of course is turned into an integration over the continuum position coordinate, N X ! d3x: ˆ i=1 2 Finally one should also realize that the Hamiltonian function would change into a (global) functional of the fields, Φ(x); Π(x). H(p ; q ) ! H [Π(x); Φ(x)] = d3x H(Π(x); Φ(x)): i i ˆ The local functional H is quite naturally called the Hamiltonian density since its volume integral is the Hamiltonian. Making these changes to the Eq.s (3, 4), we get the the action and Hamilton's equation for a field theory, I [Φ(x; t); Π(x; t)] = dt d3x Π(x) Φ(_ x) − H [Π(x); Φ(x)] ; (5) ˆ ˆ and, δH Φ(_ x) = ; δΠ(x) δH Π(_ x) = − : (6) δΦ(x) However the rules of functional derivatives are different from what we had in Lorentz invariant action functional approach. The new rules of functional integration can be easily deduced by generalizing the following formula for N−degrees of freedom to continuum limit, @qi @pi = δij; = δij: @qj @pj Using the prescribed rules, we now replace the i; j labels by position coordinates, x; y. To wit, δΦ(x) δΠ(x) = δ3(x − y); = δ3(x − y): (7) δΦ(y) δΠ(y) 3 Hamiltonian from the Lagrangian Starting with a Lagrangian for a physical system, one can extract the Hamiltonian thru a Legendre transform. Two functions, f(x) and g(x) are said to be Legendre transforms of each other if their first derivatives are functional inverse of each other, i.e. 0 0 df(y) x = f (g (x)) = dy y=g0(x) where primes denote first derivative of the function wrt their respective arguments. Solving this condition one gets, x g0(x) = f(y) So the Hamiltonian is defined to be the Legendre transform of the Lagrangian, H(p; q) = p q_(p) − L(q; q_(p)): Now hereq _(p) means we have inverted the relation, @L(q; q_) p = @q_ 3 to express,q _ as a function of p (and more generally a function of both p and q). For a system with N- degrees of freedom, the definition is, ! X H = pi q_i − L(qi; q_i) i Now we can follow the prescription mentioned in the previous section to take the continuum limit, N ! 1 whereby we obtain the Lagrangian for a field theory, H = d3x Π(x) Φ(_ x) − L[Φ(x); Φ(_ x)]; ˆ Now since the Lagrangian itself can be expressed as a volume integral of Lagrangian density, L[Φ(x); Φ(_ x)] = d3xL; L = L(Φ(x); Φ(_ x)); ˆ we have, H = d3x Π(x) Φ(_ x) − L ; ˆ where first we need to express, Φ(_ x) in terms of Π(x) by inverting, δL Π(x) = : δΦ(_ x) 3.1 Example: Real (free) scalar field theory The real scalar field theory is defined by the action, 1 1 I[Φ(x)] = d4x L; L = (@ Φ) (@µΦ) − m2Φ2: ˆ 2 µ 2 So the Lagrangian is, 1 1 L = d3x (@ Φ) (@µΦ) − m2Φ2 ; ˆ 2 µ 2 1 1 1 = d3x Φ_ 2 − rΦ · rΦ − m2Φ2 : ˆ 2 2 2 The momentum conjugate to Φ(x) is, δL Π(x) = δΦ(_ x) δ 3 1 2 1 1 2 2 = d y So Φ_ (y) − ryΦ(y) · ryΦ(y) − m Φ (y) δΦ(_ x) ˆ 2 2 2 δ 1 = d3y Φ_ 2(y) ˆ δΦ(_ x) 2 δΦ(_ y) = d3y Φ(_ y) ˆ δΦ(_ x) = d3y Φ(_ y) δ3(x − y) ˆ = Φ(_ x): 4 We can use this to eliminate Φ_ in terms of Π. The Lagrangian then becomes, 1 1 1 L = d3x Π2 − rΦ · rΦ − m2Φ2 ; ˆ 2 2 2 and the Hamiltonian is, H = d3x Π(x) Φ(_ x) − L ˆ = d3x Π2(x) − L ˆ 1 1 1 = d3x Π2(x) − d3x Π2 − rΦ · rΦ − m2Φ2 ˆ ˆ 2 2 2 1 1 1 = d3x Π2 + rΦ · rΦ + m2Φ2 : (8) ˆ 2 2 2 From this we can extract the Hamilton's equations. The first Hamilton's equation doesn't give us anything but the old relation, δH Φ(_ x) = δΠ(x) = Π(x): (9) However the second equation is non-trivial,, δH −Π(_ x) = δΦ(x) δ 1 1 1 = d3y Π2 + r Φ · r Φ + m2Φ2 δΦ(x) ˆ 2 2 y y 2 δ δΦ(y) = d3y r Φ · r Φ(y) + m2Φ(y) ˆ y y δΦ(x) δΦ(x) = d3y r Φ · r δ3(x − y) + m2Φ(y) δ3(x − y) ˆ y y = d3y −∇2 Φ + m2Φ δ3(x − y) ˆ y = − r2 − m2 Φ(x): (10) So equations are, Φ(_ x) = Π(x); Π(_ x) = r2 − m2 Φ(x): Replacing, Π = Φ_ in the second equation we get the second equation to look like, Φ¨ = r2 − m2 Φ; 2 2 2 @t − r + m Φ = 0: This is nothing but the well-familiar Klein-Gordon equation, @2 + m2 Φ = 0: Thus we have recovered the correct equation of motion using the field Hamiltonian, Eq.
Recommended publications
  • Covariant Hamiltonian Field Theory 3
    December 16, 2020 2:58 WSPC/INSTRUCTION FILE kfte COVARIANT HAMILTONIAN FIELD THEORY JURGEN¨ STRUCKMEIER and ANDREAS REDELBACH GSI Helmholtzzentrum f¨ur Schwerionenforschung GmbH Planckstr. 1, 64291 Darmstadt, Germany and Johann Wolfgang Goethe-Universit¨at Frankfurt am Main Max-von-Laue-Str. 1, 60438 Frankfurt am Main, Germany [email protected] Received 18 July 2007 Revised 14 December 2020 A consistent, local coordinate formulation of covariant Hamiltonian field theory is pre- sented. Whereas the covariant canonical field equations are equivalent to the Euler- Lagrange field equations, the covariant canonical transformation theory offers more gen- eral means for defining mappings that preserve the form of the field equations than the usual Lagrangian description. It is proved that Poisson brackets, Lagrange brackets, and canonical 2-forms exist that are invariant under canonical transformations of the fields. The technique to derive transformation rules for the fields from generating functions is demonstrated by means of various examples. In particular, it is shown that the infinites- imal canonical transformation furnishes the most general form of Noether’s theorem. We furthermore specify the generating function of an infinitesimal space-time step that conforms to the field equations. Keywords: Field theory; Hamiltonian density; covariant. PACS numbers: 11.10.Ef, 11.15Kc arXiv:0811.0508v6 [math-ph] 15 Dec 2020 1. Introduction Relativistic field theories and gauge theories are commonly formulated on the basis of a Lagrangian density L1,2,3,4. The space-time evolution of the fields is obtained by integrating the Euler-Lagrange field equations that follow from the four-dimensional representation of Hamilton’s action principle.
    [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]
  • Hamilton's Equations. Conservation Laws. Reduction. Poisson Brackets
    Hamiltonian Formalism: Hamilton's equations. Conservation laws. Reduction. Poisson Brackets. Physics 6010, Fall 2016 Hamiltonian Formalism: Hamilton's equations. Conservation laws. Reduction. Poisson Brackets. Relevant Sections in Text: 8.1 { 8.3, 9.5 The Hamiltonian Formalism We now return to formal developments: a study of the Hamiltonian formulation of mechanics. This formulation of mechanics is in many ways more powerful than the La- grangian formulation. Among the advantages of Hamiltonian mechanics we note that: it leads to powerful geometric techniques for studying the properties of dynamical systems, it allows for a beautiful expression of the relation between symmetries and conservation laws, and it leads to many structures that can be viewed as the macroscopic (\classical") imprint of quantum mechanics. Although the Hamiltonian form of mechanics is logically independent of the Lagrangian formulation, it is convenient and instructive to introduce the Hamiltonian formalism via transition from the Lagrangian formalism, since we have already developed the latter. (Later I will indicate how to give an ab initio development of the Hamiltonian formal- ism.) The most basic change we encounter when passing from Lagrangian to Hamiltonian methods is that the \arena" we use to describe the equations of motion is no longer the configuration space, or even the velocity phase space, but rather the momentum phase space. Recall that the Lagrangian formalism is defined once one specifies a configuration space Q (coordinates qi) and then the velocity phase space Ω (coordinates (qi; q_i)). The mechanical system is defined by a choice of Lagrangian, L, which is a function on Ω (and possible the time): L = L(qi; q_i; t): Curves in the configuration space Q { or in the velocity phase space Ω { satisfying the Euler-Lagrange (EL) equations, @L d @L − = 0; @qi dt @q_i define the dynamical behavior of the system.
    [Show full text]
  • Poisson Structures and Integrability
    Poisson Structures and Integrability Peter J. Olver University of Minnesota http://www.math.umn.edu/ olver ∼ Hamiltonian Systems M — phase space; dim M = 2n Local coordinates: z = (p, q) = (p1, . , pn, q1, . , qn) Canonical Hamiltonian system: dz O I = J H J = − dt ∇ ! I O " Equivalently: dpi ∂H dqi ∂H = = dt − ∂qi dt ∂pi Lagrange Bracket (1808): n ∂pi ∂qi ∂qi ∂pi [ u , v ] = ∂u ∂v − ∂u ∂v i#= 1 (Canonical) Poisson Bracket (1809): n ∂u ∂v ∂u ∂v u , v = { } ∂pi ∂qi − ∂qi ∂pi i#= 1 Given functions u , . , u , the (2n) (2n) matrices with 1 2n × respective entries [ u , u ] u , u i, j = 1, . , 2n i j { i j } are mutually inverse. Canonical Poisson Bracket n ∂F ∂H ∂F ∂H F, H = F T J H = { } ∇ ∇ ∂pi ∂qi − ∂qi ∂pi i#= 1 = Poisson (1809) ⇒ Hamiltonian flow: dz = z, H = J H dt { } ∇ = Hamilton (1834) ⇒ First integral: dF F, H = 0 = 0 F (z(t)) = const. { } ⇐⇒ dt ⇐⇒ Poisson Brackets , : C∞(M, R) C∞(M, R) C∞(M, R) { · · } × −→ Bilinear: a F + b G, H = a F, H + b G, H { } { } { } F, a G + b H = a F, G + b F, H { } { } { } Skew Symmetric: F, H = H, F { } − { } Jacobi Identity: F, G, H + H, F, G + G, H, F = 0 { { } } { { } } { { } } Derivation: F, G H = F, G H + G F, H { } { } { } F, G, H C∞(M, R), a, b R. ∈ ∈ In coordinates z = (z1, . , zm), F, H = F T J(z) H { } ∇ ∇ where J(z)T = J(z) is a skew symmetric matrix. − The Jacobi identity imposes a system of quadratically nonlinear partial differential equations on its entries: ∂J jk ∂J ki ∂J ij J il + J jl + J kl = 0 ! ∂zl ∂zl ∂zl " #l Given a Poisson structure, the Hamiltonian flow corresponding to H C∞(M, R) is the system of ordinary differential equati∈ons dz = z, H = J(z) H dt { } ∇ Lie’s Theory of Function Groups Used for integration of partial differential equations: F , F = G (F , .
    [Show full text]
  • Advanced Quantum Theory AMATH473/673, PHYS454
    Advanced Quantum Theory AMATH473/673, PHYS454 Achim Kempf Department of Applied Mathematics University of Waterloo Canada c Achim Kempf, September 2017 (Please do not copy: textbook in progress) 2 Contents 1 A brief history of quantum theory 7 1.1 The classical period . 7 1.2 Planck and the \Ultraviolet Catastrophe" . 7 1.3 Discovery of h ................................ 8 1.4 Mounting evidence for the fundamental importance of h . 9 1.5 The discovery of quantum theory . 9 1.6 Relativistic quantum mechanics . 11 1.7 Quantum field theory . 12 1.8 Beyond quantum field theory? . 14 1.9 Experiment and theory . 16 2 Classical mechanics in Hamiltonian form 19 2.1 Newton's laws for classical mechanics cannot be upgraded . 19 2.2 Levels of abstraction . 20 2.3 Classical mechanics in Hamiltonian formulation . 21 2.3.1 The energy function H contains all information . 21 2.3.2 The Poisson bracket . 23 2.3.3 The Hamilton equations . 25 2.3.4 Symmetries and Conservation laws . 27 2.3.5 A representation of the Poisson bracket . 29 2.4 Summary: The laws of classical mechanics . 30 2.5 Classical field theory . 31 3 Quantum mechanics in Hamiltonian form 33 3.1 Reconsidering the nature of observables . 34 3.2 The canonical commutation relations . 35 3.3 From the Hamiltonian to the equations of motion . 38 3.4 From the Hamiltonian to predictions of numbers . 42 3.4.1 Linear maps . 42 3.4.2 Choices of representation . 43 3.4.3 A matrix representation . 44 3 4 CONTENTS 3.4.4 Example: Solving the equations of motion for a free particle with matrix-valued functions .
    [Show full text]
  • Description of Bound States and Scattering Processes Using Hamiltonian Renormalization Group Procedure for Effective Particles in Quantum field Theory
    Description of bound states and scattering processes using Hamiltonian renormalization group procedure for effective particles in quantum field theory Marek Wi˛eckowski June 2005 Doctoral thesis Advisor: Professor Stanisław D. Głazek Warsaw University Department of Physics Institute of Theoretical Physics Warsaw 2005 i Acknowledgments I am most grateful to my advisor, Professor Stanisław Głazek, for his invaluable advice and assistance during my work on this thesis. I would also like to thank Tomasz Masłowski for allowing me to use parts of our paper here. My heartfelt thanks go to Dr. Nick Ukiah, who supported and encouraged me during the writing of this thesis and who helped me in editing the final manuscript in English. Finally, I would like to thank my family, especially my grandmother, who has been a great support to me during the last few years. MAREK WI ˛ECKOWSKI,DESCRIPTION OF BOUND STATES AND SCATTERING... ii MAREK WI ˛ECKOWSKI,DESCRIPTION OF BOUND STATES AND SCATTERING... iii Abstract This thesis presents examples of a perturbative construction of Hamiltonians Hλ for effective particles in quantum field theory (QFT) on the light front. These Hamiltonians (1) have a well-defined (ultraviolet-finite) eigenvalue problem for bound states of relativistic constituent fermions, and (2) lead (in a scalar theory with asymptotic freedom in perturbation theory in third and partly fourth order) to an ultraviolet-finite and covariant scattering matrix, as the Feynman diagrams do. λ is a parameter of the renormalization group for Hamiltonians and qualitatively means the inverse of the size of the effective particles. The same procedure of calculating the operator Hλ applies in description of bound states and scattering.
    [Show full text]
  • Turbulence, Entropy and Dynamics
    TURBULENCE, ENTROPY AND DYNAMICS Lecture Notes, UPC 2014 Jose M. Redondo Contents 1 Turbulence 1 1.1 Features ................................................ 2 1.2 Examples of turbulence ........................................ 3 1.3 Heat and momentum transfer ..................................... 4 1.4 Kolmogorov’s theory of 1941 ..................................... 4 1.5 See also ................................................ 6 1.6 References and notes ......................................... 6 1.7 Further reading ............................................ 7 1.7.1 General ............................................ 7 1.7.2 Original scientific research papers and classic monographs .................. 7 1.8 External links ............................................. 7 2 Turbulence modeling 8 2.1 Closure problem ............................................ 8 2.2 Eddy viscosity ............................................. 8 2.3 Prandtl’s mixing-length concept .................................... 8 2.4 Smagorinsky model for the sub-grid scale eddy viscosity ....................... 8 2.5 Spalart–Allmaras, k–ε and k–ω models ................................ 9 2.6 Common models ........................................... 9 2.7 References ............................................... 9 2.7.1 Notes ............................................. 9 2.7.2 Other ............................................. 9 3 Reynolds stress equation model 10 3.1 Production term ............................................ 10 3.2 Pressure-strain interactions
    [Show full text]
  • Advances in Quantum Field Theory
    ADVANCES IN QUANTUM FIELD THEORY Edited by Sergey Ketov Advances in Quantum Field Theory Edited by Sergey Ketov Published by InTech Janeza Trdine 9, 51000 Rijeka, Croatia Copyright © 2012 InTech All chapters are Open Access distributed under the Creative Commons Attribution 3.0 license, which allows users to download, copy and build upon published articles even for commercial purposes, as long as the author and publisher are properly credited, which ensures maximum dissemination and a wider impact of our publications. After this work has been published by InTech, authors have the right to republish it, in whole or part, in any publication of which they are the author, and to make other personal use of the work. Any republication, referencing or personal use of the work must explicitly identify the original source. As for readers, this license allows users to download, copy and build upon published chapters even for commercial purposes, as long as the author and publisher are properly credited, which ensures maximum dissemination and a wider impact of our publications. Notice Statements and opinions expressed in the chapters are these of the individual contributors and not necessarily those of the editors or publisher. No responsibility is accepted for the accuracy of information contained in the published chapters. The publisher assumes no responsibility for any damage or injury to persons or property arising out of the use of any materials, instructions, methods or ideas contained in the book. Publishing Process Manager Romana Vukelic Technical Editor Teodora Smiljanic Cover Designer InTech Design Team First published February, 2012 Printed in Croatia A free online edition of this book is available at www.intechopen.com Additional hard copies can be obtained from [email protected] Advances in Quantum Field Theory, Edited by Sergey Ketov p.
    [Show full text]
  • Hamilton-Poisson Formulation for the Rotational Motion of a Rigid Body In
    Hamilton-Poisson formulation for the rotational motion of a rigid body in the presence of an axisymmetric force field and a gyroscopic torque Petre Birtea,∗ Ioan Ca¸su, Dan Com˘anescu Department of Mathematics, West University of Timi¸soara Bd. V. Pˆarvan, No 4, 300223 Timi¸soara, Romˆania E-mail addresses: [email protected]; [email protected]; [email protected] Keywords: Poisson bracket, Casimir function, rigid body, gyroscopic torque. Abstract We give sufficient conditions for the rigid body in the presence of an axisymmetric force field and a gyroscopic torque to admit a Hamilton-Poisson formulation. Even if by adding a gyroscopic torque we initially lose one of the conserved Casimirs, we recover another conservation law as a Casimir function for a modified Poisson structure. We apply this frame to several well known results in the literature. 1 Introduction The generalized Euler-Poisson system, which is a dynamical system for the rotational motion of a rigid body in the presence of an axisymmetric force field, admits a Hamilton-Poisson formulation, where the bracket is the ”−” Kirillov-Kostant-Souriau bracket on the dual Lie algebra e(3)∗. The Hamiltonian function is of the type kinetic energy plus potential energy. The Hamiltonian function and the two Casimir functions for the K-K-S Poisson structure are conservation laws for the dynamic. Adding a certain type of gyroscopic torque the Hamiltonian and one of the Casimirs remain conserved along the solutions of the new dynamical system. In this paper we give a technique for finding a third conservation law in the presence of gyroscopic torque.
    [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]
  • Higher Order Geometric Theory of Information and Heat Based On
    Article Higher Order Geometric Theory of Information and Heat Based on Poly-Symplectic Geometry of Souriau Lie Groups Thermodynamics and Their Contextures: The Bedrock for Lie Group Machine Learning Frédéric Barbaresco Department of Advanced Radar Concepts, Thales Land Air Systems, Voie Pierre-Gilles de Gennes, 91470 Limours, France; [email protected] Received: 9 August 2018; Accepted: 9 October; Published: 2 November 2018 Abstract: We introduce poly-symplectic extension of Souriau Lie groups thermodynamics based on higher-order model of statistical physics introduced by Ingarden. This extended model could be used for small data analytics and machine learning on Lie groups. Souriau geometric theory of heat is well adapted to describe density of probability (maximum entropy Gibbs density) of data living on groups or on homogeneous manifolds. For small data analytics (rarified gases, sparse statistical surveys, …), the density of maximum entropy should consider higher order moments constraints (Gibbs density is not only defined by first moment but fluctuations request 2nd order and higher moments) as introduced by Ingarden. We use a poly-sympletic model introduced by Christian Günther, replacing the symplectic form by a vector-valued form. The poly-symplectic approach generalizes the Noether theorem, the existence of moment mappings, the Lie algebra structure of the space of currents, the (non-)equivariant cohomology and the classification of G-homogeneous systems. The formalism is covariant, i.e., no special coordinates or coordinate systems on the parameter space are used to construct the Hamiltonian equations. We underline the contextures of these models, and the process to build these generic structures. We also introduce a more synthetic Koszul definition of Fisher Metric, based on the Souriau model, that we name Souriau-Fisher metric.
    [Show full text]
  • An Introduction to Free Quantum Field Theory Through Klein-Gordon Theory
    AN INTRODUCTION TO FREE QUANTUM FIELD THEORY THROUGH KLEIN-GORDON THEORY JONATHAN EMBERTON Abstract. We provide an introduction to quantum field theory by observing the methods required to quantize the classical form of Klein-Gordon theory. Contents 1. Introduction and Overview 1 2. Klein-Gordon Theory as a Hamiltonian System 2 3. Hamilton's Equations 4 4. Quantization 5 5. Creation and Annihilation Operators and Wick Ordering 9 6. Fock Space 11 7. Maxwell's Equations and Generalized Free QFT 13 References 15 Acknowledgments 16 1. Introduction and Overview The phase space for classical mechanics is R3 × R3, where the first copy of R3 encodes position of an object and the second copy encodes momentum. Observables on this space are simply functions on R3 × R3. The classical dy- namics of the system are then given by Hamilton's equations x_ = fh; xg k_ = fh; kg; where the brackets indicate the canonical Poisson bracket and jkj2 h(x; k) = + V (x) 2m is the classical Hamiltonian. To quantize this classical field theory, we set our quantized phase space to be L2(R3). Then the observables are self-adjoint operators A on this space, and in a state 2 L2(R3), the mean value of A is given by hAi := hA ; i: The dynamics of a quantum system are then described by the quantized Hamil- ton's equations i i x_ = [H; x]_p = [H; p] h h 1 2 JONATHAN EMBERTON where the brackets indicate the commutator and the Hamiltonian operator H is given by jpj2 2 H = h(x; p) = − + V (x) = − ~ ∆ + V (x): 2m 2m A question that naturally arises from this construction is then, how do we quan- tize other classical field theories, and is this quantization as straight-forward as the canonical construction? For instance, how do Maxwell's Equations behave on the quantum scale? This happens to be somewhat of a difficult problem, with many steps involved.
    [Show full text]