Chapter 5 Symmetries and Conservation Laws

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 5 Symmetries and Conservation Laws Chapter 5 Symmetries and Conservation Laws At a very fundamental level, physics is about identifying patterns of order in Na- ture. The inception of the field starts arguably with Tycho Brahe (1546-1601) — the first modern experimental physicist — and Johannes Kepler (1571-1630) — the first modern theoretical physicist. In the 16th century, Brahe collected large amounts of astrophysical data about the location of planets and stars with groundbreaking accuracy — using an impressive telescope set up in a castle. Kepler pondered for years over Brahe’s long tables of numbers until he could finally iden- tify a pattern underlying planetary dynamics: this was summarized in the three laws of Kepler. Later on, Isaac Newton (1643-1727) referred to these achieve- ments through his famous quote: “If I have seen further it is only by standing on the shoulders of giants”. Since then, physics has always been about identifying patterns in numbers, in measurements. And a pattern is simply an indication of a repeating rule, a constant attribute in complexity, an underlying symmetry. In 1918, Emmy Noether (1882–1935) published a seminal work that clarified the deep relations between symmetries and conserved quantities in Nature. In a sense, this work organizes physics in a clear diagram and gives one a bird’s eye view of the myriad of branches of the field. Noether’s theorem, as it is called, can change the way one thinks about physics in general. It is profound yet simple. While one can study Noether’s theorem in the Newtonian formulation of me- chanics through cumbersome methods, the subject is an excellent demonstration of the power of the new formalism we developed, Lagrangian mechanics. In this sec- tion, we use the variational principle to develop a statement of Noether’s theorem. We then prove it and demonstrate its importance through examples. 189 190 CHAPTER 5. SYMMETRIES AND CONSERVATION LAWS Figure 5.1: Two particles on a rail. EXAMPLE 5-1: A simple example Let us start with a simple mechanics problem from an example we saw in Section 4.1. We have two particles, with masses m1 and m2, confined to moving along a horizontal frictionless rail as depicted in Figure 5. The action for the system is 1 1 S = dt m q˙2 + m q˙2 V (q q ) (5.1) 2 1 1 2 2 2 − 1 − 2 Z ✓ ◆ where we are considering some interaction between the particles described by a potential V (q q ) that depends only on the distance between the particles. 1 − 2 Let us consider a simple transformation of the coordinates given by q10 = q1 +C ,q20 = q2 + C (5.2) where C is some arbitrary constant. We then have q˙10 =˙q1 , q˙20 =˙q2 . (5.3) Hence, the kinetic terms in the action are unchanged under this transformation. Furthermore, we also have q0 q0 = q q (5.4) 1 − 2 1 − 2 5.1. INFINITESIMAL TRANSFORMATIONS 191 implying that the potential term is also unchanged. The action then preserves its overall structural form under the transformation 1 2 1 2 S dt m q˙ + m q˙ V (q0 q0 ) . (5.5) ! 2 1 01 2 2 02 − 1 − 2 Z ✓ ◆ This means that the equations of motion, written in the primed transformed co- ordinates, are identical to the ones written in the unprimed original coordinates. We then say that the transformation given by (5.2) is a symmetry of our system. Physically, we are simply saying that — since the interaction between the particles depends only on the distance between them — a constant shift of both coordinates leaves the dynamics una↵ected. It is also useful to consider an infinitesimal version of such a transformation. Let us assume that the constant C is small, C ✏; and we write ! q0 q δq = ✏ (5.6) k − k ⌘ k for k =1, 2. We then say that δqk = ✏ is a symmetry of our system. To make these ideas more useful, we want to extend this example by considering a general class of interesting transformations that we may want to consider. 5.1 Infinitesimal transformations There are two useful types of infinitesimal transformations: direct and indirect ones. 5.1.1 Direct transformations A direct transformation deforms the degrees of freedom of a setup directly: δq (t)=q0 (t) q (t) ∆q (t, q) . (5.7) k k − k ⌘ k We use the notation ∆to distinguish a direct transformation. Note that ∆qk(t, q) is possibly a function of time and all of the degrees of freedom in the problem. In the previous example, we had the special case where ∆qk(t, q) = constant. But it need not be so. Figure 5.2(a) depicts a direct transformation: it is an arbitrary shift in the qk’s. Note also that ∆qk(t, q) is a small deformation. 192 CHAPTER 5. SYMMETRIES AND CONSERVATION LAWS Figure 5.2: The two types of transformations considered: direct on the left, indirect on the right. 5.1.2 Indirect transformations In contrast, an indirect transformation a↵ects the degrees of freedom indirectly — through the transformation of the time coordinate: δt(t, q) t0 t. (5.8) ⌘ − Note again that the shift in time can depend on time and the degrees of freedom as well! It is again assumed to be small. This is however not the end of the story since the degrees of freedom depend on time and will get a↵ected as well — indirectly dqk qk(t)=qk(t0 δt) qk(t0) δt qk(t0) q˙kδt (5.9) − ' − dt0 ' − where we have used a Taylor expansion in δt to linear order since δt is small. We also have used dqk/dt0 = dqk/dt since this term already multiplies a power of δt: to linear order in δt, δtdq /dt = δtdq /dt δtq˙ . We then see that shifting time k 0 k ⌘ k results in a shift in the degrees of freedom δq = q (t0) q (t)=q ˙δt(t, q) . (5.10) k k − k Figure 5.2(b) shows how you can think of this e↵ect graphically. 5.1. INFINITESIMAL TRANSFORMATIONS 193 5.1.3 Combined transformations In general, we want to consider a transformation that may include both direct and indirect pieces. We would write δq = q0 (t0) q (t)=q0 (t0)+[ q (t0)+q (t0)] q (t) k k − k k − k k − k =[q0 (t0) q (t0)] + [q (t0) q (t)] = ∆q (t0,q)+q ˙δt(t, q) k − k k − k k =∆qk(t, q)+q ˙δt(t, q) . (5.11) where in the last line we have equated t and t0 since the first term is already linear in the small parameters. To specify a particular transformation, we would then need to provide a set of functions δt(t, q) and δqk(t, q) . (5.12) Equation (5.11) then determines ∆qk(t, q). For N degrees of freedom, that’s N +1 functions of time and the qk’s. Let us look at a few examples. EXAMPLE 5-2: Translations Consider a single particle in three dimensions, described by the three Cartesian coordinates x1 = x, x2 = y, and x3 = z. We also have the time coordinate x0 = ct. An infinitesimal spatial translation can be realized by δxi(t, x)=✏i ,δt(t, x)=0 ∆xi(t, x)=✏i (5.13) ) where i =1, 2, 3, and the ✏i’s are three small constants. A translation in space is then defined by δt(t, x)=0,δxi(t, x)=✏i (5.14) { } A translation in time on the other hand would be given by δxi(t, x)=0 ,δt(t, x)=" ∆xi(t, x)= x˙ i" (5.15) ) − for constant ". Notice that we require that the total shifts in the xi’s — the δxi(t, x)’s — vanish. This then generates direct shifts, the ∆xi’s, to compensate for the indirect e↵ect on the spatial coordinates from the shifting of the time. A translation in time is then defined by δt(t, x)=✏,δxi(t, x)=0 (5.16) { } 194 CHAPTER 5. SYMMETRIES AND CONSERVATION LAWS EXAMPLE 5-3: Rotations To describe rotations, let us consider a particle in two dimensions for simplicity. We use the coordinates x1 = x and x2 = y. We start by specifying δt(t, x)=0 δxi(t, x)=∆xi(t, x) . (5.17) ) Next, we look at an arbitrary rotation angle ✓ using (2.21) x10 cos ✓ sin ✓ x1 = . (5.18) x20 sin ✓ cos ✓ x2 ✓ ◆ ✓ − ◆✓ ◆ We however need to focus on an infinitesimal version of this transformation: i.e. we need to consider small angle ✓. Using cos ✓ 1 and sin ✓ ✓ to second order ⇠ ⇠ in ✓, we then write x10 1 ✓ x1 = . (5.19) x20 ✓ 1 x2 ✓ ◆ ✓ − ◆✓ ◆ This gives δx1(t, x)=x10 (t) x1(t)=✓x2(t)=∆x1(t, x) − δx2(t, x)=x20 (t) x2(t)= ✓x1(t)=∆x2(t, x) . (5.20) − − We see we have a more non-trivial transformation. Rotations can then be defined through δt(t, x)=0,δxi(t, x)=✓"ijxj(t) (5.21) { } where j is summed over 1 and 2. We have also introduced a useful shorthand: "ij. It is called the totally antisymmetric matrix in two dimensions and defined as: "11 = "22 =0 ,"12 = "21 =1. (5.22) − It allows us to write the transformation in a more compact notation. EXAMPLE 5-4: Lorentz transformations To find the infinitesimal form of Lorentz transformations, we can start with the general transformation equations (2.15) and take β small. We need to be careful however to keep the leading order terms in β in all expansions. Given our previous example, it is easier to map the problem onto a rotation with hyperbolic trig functions using (2.27).
Recommended publications
  • Quantum Mechanics and Relativity): 21St Century Analysis
    Noname manuscript No. (will be inserted by the editor) Joule's 19th Century Energy Conservation Meta-Law and the 20th Century Physics (Quantum Mechanics and Relativity): 21st Century Analysis Vladik Kreinovich · Olga Kosheleva Received: December 22, 2019 / Accepted: date Abstract Joule's Energy Conservation Law was the first \meta-law": a gen- eral principle that all physical equations must satisfy. It has led to many important and useful physical discoveries. However, a recent analysis seems to indicate that this meta-law is inconsistent with other principles { such as the existence of free will. We show that this conclusion about inconsistency is based on a seemingly reasonable { but simplified { analysis of the situation. We also show that a more detailed mathematical and physical analysis of the situation reveals that not only Joule's principle remains true { it is actually strengthened: it is no longer a principle that all physical theories should satisfy { it is a principle that all physical theories do satisfy. Keywords Joule · Energy Conservation Law · Free will · General Relativity · Planck's constant 1 Introduction Joule's Energy Conservation Law: historically the first meta-law. Throughout the centuries, physicists have been trying to come up with equa- tions and laws that describe different physical phenomenon. Before Joule, how- ever, there were no general principles that restricted such equations. James Joule showed, in [13{15], that different physical phenomena are inter-related, and that there is a general principle covering all the phenom- This work was supported in part by the US National Science Foundation grants 1623190 (A Model of Change for Preparing a New Generation for Professional Practice in Computer Science) and HRD-1242122 (Cyber-ShARE Center of Excellence).
    [Show full text]
  • Noether's Theorems and Gauge Symmetries
    Noether’s Theorems and Gauge Symmetries Katherine Brading St. Hugh’s College, Oxford, OX2 6LE [email protected] and Harvey R. Brown Sub-faculty of Philosophy, University of Oxford, 10 Merton Street, Oxford OX1 4JJ [email protected] August 2000 Consideration of the Noether variational problem for any theory whose action is invariant under global and/or local gauge transformations leads to three distinct the- orems. These include the familiar Noether theorem, but also two equally important but much less well-known results. We present, in a general form, all the main results relating to the Noether variational problem for gauge theories, and we show the rela- tionships between them. These results hold for both Abelian and non-Abelian gauge theories. 1 Introduction There is widespread confusion over the role of Noether’s theorem in the case of local gauge symmetries,1 as pointed out in this journal by Karatas and Kowalski (1990), and Al-Kuwari and Taha (1991).2 In our opinion, the main reason for the confusion is failure to appreciate that Noether offered two theorems in her 1918 work. One theorem applies to symmetries associated with finite dimensional arXiv:hep-th/0009058v1 8 Sep 2000 Lie groups (global symmetries), and the other to symmetries associated with infinite dimensional Lie groups (local symmetries); the latter theorem has been widely forgotten. Knowledge of Noether’s ‘second theorem’ helps to clarify the significance of the results offered by Al-Kuwari and Taha for local gauge sym- metries, along with other important and related work such as that of Bergmann (1949), Trautman (1962), Utiyama (1956, 1959), and Weyl (1918, 1928/9).
    [Show full text]
  • Oil and Gas Conservation Law in Review
    The Pennsylvania Oil and Gas Conservation Law: A Summary of the Statutory Provisions 58 P.S. §§ 401-419 (March 2009) Prepared by Anna M. Clovis, Research Assistant Under the Supervision of Ross H. Pifer, J.D., LL.M., Director The Agricultural Law Resource and Reference Center Penn State Dickinson School of Law http://www.dsl.psu.edu/centers/aglaw.cfm The Oil and Gas Conservation Law (OGC Law) is one of the primary laws that regulate the extraction of natural gas in Pennsylvania. The OGC Law has been summarized in a section- by-section manner in this research publication, which provides an overview of the issues regulated by the OGC Law. To obtain detailed information on the individual statutory sections, the full text of the OGC Law should be reviewed. Declaration of Policy The OGC Law reflects a policy “to foster, encourage, and promote the development, production, and utilization” of Pennsylvania’s oil and natural gas resources. Development activities should occur in a manner to prevent the waste of oil and natural gas. Oil and natural gas operations should be regulated to ensure that “the Commonwealth shall realize and enjoy the maximum benefit of these natural resources.” Since strata above the Onondaga horizon have been extensively developed since 1850, it would be impractical to include shallow wells within the regulations of the OGC Law. Section 1: Short Title (58 P.S. § 401) This statute shall be known as the “Oil and Gas Conservation Law.” Section 2: Definitions (58 P.S. § 402) This section provides statutory definitions for various oil and gas terms used in the OGC Law.
    [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]
  • Physics 325: General Relativity Spring 2019 Problem Set 2
    Physics 325: General Relativity Spring 2019 Problem Set 2 Due: Fri 8 Feb 2019. Reading: Please skim Chapter 3 in Hartle. Much of this should be review, but probably not all of it|be sure to read Box 3.2 on Mach's principle. Then start on Chapter 6. Problems: 1. Spacetime interval. Hartle Problem 4.13. 2. Four-vectors. Hartle Problem 5.1. 3. Lorentz transformations and hyperbolic geometry. In class, we saw that a Lorentz α0 α β transformation in 2D can be written as a = L β(#)a , that is, 0 ! ! ! a0 cosh # − sinh # a0 = ; (1) a10 − sinh # cosh # a1 where a is spacetime vector. Here, the rapidity # is given by tanh # = β; cosh # = γ; sinh # = γβ; (2) where v = βc is the velocity of frame S0 relative to frame S. (a) Show that two successive Lorentz boosts of rapidity #1 and #2 are equivalent to a single α γ α Lorentz boost of rapidity #1 +#2. In other words, check that L γ(#1)L(#2) β = L β(#1 +#2), α where L β(#) is the matrix in Eq. (1). You will need the following hyperbolic trigonometry identities: cosh(#1 + #2) = cosh #1 cosh #2 + sinh #1 sinh #2; (3) sinh(#1 + #2) = sinh #1 cosh #2 + cosh #1 sinh #2: (b) From Eq. (3), deduce the formula for tanh(#1 + #2) in terms of tanh #1 and tanh #2. For the appropriate choice of #1 and #2, use this formula to derive the special relativistic velocity tranformation rule V − v V 0 = : (4) 1 − vV=c2 Physics 325, Spring 2019: Problem Set 2 p.
    [Show full text]
  • Application of Axiomatic Formal Theory to the Abraham--Minkowski
    Application of axiomatic formal theory to the Abraham{Minkowski controversy Michael E. Crenshaw∗ Charles M. Bowden Research Laboratory, US Army Combat Capabilities Development Command (DEVCOM) - Aviation and Missile Center, Redstone Arsenal, AL 35898, USA (Dated: August 23, 2019) A transparent linear dielectric in free space that is illuminated by a finite quasimonochromatic field is a thermodynamically closed system. The energy{momentum tensor that is derived from Maxwellian continuum electrodynamics for this closed system is inconsistent with conservation of momentum; that is the foundational issue of the century-old Abraham{Minkowski controversy. The long-standing resolution of the Abraham{Minkowski controversy is to view continuum electrodynam- ics as a subsystem and write the total energy{momentum tensor as the sum of an electromagnetic energy{momentum tensor and a phenomenological material energy{momentum tensor. We prove that if we add a material energy{momentum tensor to the electromagnetic energy{momentum ten- sor, in the prescribed manner, then the total energy, the total linear momentum, and the total angular momentum are constant in time, but other aspects of the conservation laws are violated. Specifically, the four-divergence of the total (electromagnetic plus material) energy{momentum ten- sor is self-inconsistent and violates Poynting's theorem. Then, the widely accepted resolution of the Abraham{Minkowski dilemma is proven to be manifestly false. The fundamental physical principles of electrodynamics, conservation laws, and special relativity are intrinsic to the vacuum; the ex- tant dielectric versions of physical principles are non-fundamental extrapolations from the vacuum theory. Because the resolution of the Abraham{Minkowski momentum controversy is proven false, the extant theoretical treatments of macroscopic continuum electrodynamics, dielectric special rel- ativity, and energy{momentum conservation in a simple linear dielectric are mutually inconsistent for macroscopic fields in matter.
    [Show full text]
  • Noether: the More Things Change, the More Stay the Same
    Noether: The More Things Change, the More Stay the Same Grzegorz G luch∗ R¨udigerUrbanke EPFL EPFL Abstract Symmetries have proven to be important ingredients in the analysis of neural networks. So far their use has mostly been implicit or seemingly coincidental. We undertake a systematic study of the role that symmetry plays. In particular, we clarify how symmetry interacts with the learning algorithm. The key ingredient in our study is played by Noether's celebrated theorem which, informally speaking, states that symmetry leads to conserved quantities (e.g., conservation of energy or conservation of momentum). In the realm of neural networks under gradient descent, model symmetries imply restrictions on the gradient path. E.g., we show that symmetry of activation functions leads to boundedness of weight matrices, for the specific case of linear activations it leads to balance equations of consecutive layers, data augmentation leads to gradient paths that have \momentum"-type restrictions, and time symmetry leads to a version of the Neural Tangent Kernel. Symmetry alone does not specify the optimization path, but the more symmetries are contained in the model the more restrictions are imposed on the path. Since symmetry also implies over-parametrization, this in effect implies that some part of this over-parametrization is cancelled out by the existence of the conserved quantities. Symmetry can therefore be thought of as one further important tool in understanding the performance of neural networks under gradient descent. 1 Introduction There is a large body of work dedicated to understanding what makes neural networks (NNs) perform so well under versions of gradient descent (GD).
    [Show full text]
  • Integrals of Motion Key Point #15 from Bachelor Course
    KeyKey12-10-07 see PointPoint http://www.strw.leidenuniv.nl/˜ #15#10 fromfrom franx/college/ BachelorBachelor mf-sts-07-c4-5 Course:Course:12-10-07 see http://www.strw.leidenuniv.nl/˜ IntegralsIntegrals ofof franx/college/ MotionMotion mf-sts-07-c4-6 4.2 Constants and Integrals of motion Integrals (BT 3.1 p 110-113) Much harder to define. E.g.: 1 2 Energy (all static potentials): E(⃗x,⃗v)= 2 v + Φ First, we define the 6 dimensional “phase space” coor- dinates (⃗x,⃗v). They are conveniently used to describe Lz (axisymmetric potentials) the motions of stars. Now we introduce: L⃗ (spherical potentials) • Integrals constrain geometry of orbits. • Constant of motion: a function C(⃗x,⃗v, t) which is constant along any orbit: examples: C(⃗x(t1),⃗v(t1),t1)=C(⃗x(t2),⃗v(t2),t2) • 1. Spherical potentials: E,L ,L ,L are integrals of motion, but also E,|L| C is a function of ⃗x, ⃗v, and time t. x y z and the direction of L⃗ (given by the unit vector ⃗n, Integral of motion which is defined by two independent numbers). ⃗n de- • : a function I(x, v) which is fines the plane in which ⃗x and ⃗v must lie. Define coor- constant along any orbit: dinate system with z axis along ⃗n I[⃗x(t1),⃗v(t1)] = I[⃗x(t2),⃗v(t2)] ⃗x =(x1,x2, 0) I is not a function of time ! Thus: integrals of motion are constants of motion, ⃗v =(v1,v2, 0) but constants of motion are not always integrals of motion! → ⃗x and ⃗v constrained to 4D region of the 6D phase space.
    [Show full text]
  • STUDENT HIGHLIGHT – Hassan Najafi Alishah - “Symplectic Geometry, Poisson Geometry and KAM Theory” (Mathematics)
    SEPTEMBER OCTOBER // 2010 NUMBER// 28 STUDENT HIGHLIGHT – Hassan Najafi Alishah - “Symplectic Geometry, Poisson Geometry and KAM theory” (Mathematics) Hassan Najafi Alishah, a CoLab program student specializing in math- The Hamiltonian function (or energy function) is constant along the ematics, is doing advanced work in the Mathematics Departments of flow. In other words, the Hamiltonian function is a constant of motion, Instituto Superior Técnico, Lisbon, and the University of Texas at Austin. a principle that is sometimes called the energy conservation law. A This article provides an overview of his research topics. harmonic spring, a pendulum, and a one-dimensional compressible fluid provide examples of Hamiltonian Symplectic and Poisson geometries underlie all me- systems. In the former, the phase spaces are symplec- chanical systems including the solar system, the motion tic manifolds, but for the later phase space one needs of a rigid body, or the interaction of small molecules. the more general concept of a Poisson manifold. The origins of symplectic and Poisson structures go back to the works of the French mathematicians Joseph-Louis A Hamiltonian system with enough suitable constants Lagrange (1736-1813) and Simeon Denis Poisson (1781- of motion can be integrated explicitly and its orbits can 1840). Hermann Weyl (1885-1955) first used “symplectic” be described in detail. These are called completely in- with its modern mathematical meaning in his famous tegrable Hamiltonian systems. Under appropriate as- monograph “The Classical groups.” (The word “sym- sumptions (e.g., in a compact symplectic manifold) plectic” itself is derived from a Greek word that means the motions of a completely integrable Hamiltonian complex.) Lagrange introduced the concept of a system are quasi-periodic.
    [Show full text]
  • Conservation Laws in Arbitrary Space-Times Annales De L’I
    ANNALES DE L’I. H. P., SECTION A I. M. BENN Conservation laws in arbitrary space-times Annales de l’I. H. P., section A, tome 37, no 1 (1982), p. 67-91 <http://www.numdam.org/item?id=AIHPA_1982__37_1_67_0> © Gauthier-Villars, 1982, tous droits réservés. L’accès aux archives de la revue « Annales de l’I. H. P., section A » implique l’accord avec les conditions générales d’utilisation (http://www.numdam. org/conditions). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright. Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques http://www.numdam.org/ Ann. Henri Poincaré, Section A : Vol. XXXVII, n° 1. 1982, 67 Physique theorique.’ Conservation laws in arbitrary space-times I. M. BENN Department of Physics, University of Lancaster, Lancaster, U. K. ABSTRACT. - Definitions of energy and angular momentum are dis- cussed within the context of field theories in arbitrary space-times. It is shown that such definitions rely upon the existence of symmetries of the space-time geometry. The treatment of that geometry as a background is contrasted with Einstein’s theory of dynamical geometry : general rela- tivity. It is shown that in the former case symmetries lead to the existence of closed 3-forms, whereas in the latter they lead to the existence of 2-forms which are closed in source-free regions, thus allowing mass and angular momentum to be defined as de Rham periods. INTRODUCTION The concepts of energy and angular momentum are perhaps two of the most deeply rooted concepts of physics, yet in general relativity their exact role is still open to question [7]] [2 ].
    [Show full text]
  • Lecture 8 Symmetries, Conserved Quantities, and the Labeling of States Angular Momentum
    Lecture 8 Symmetries, conserved quantities, and the labeling of states Angular Momentum Today’s Program: 1. Symmetries and conserved quantities – labeling of states 2. Ehrenfest Theorem – the greatest theorem of all times (in Prof. Anikeeva’s opinion) 3. Angular momentum in QM ˆ2 ˆ 4. Finding the eigenfunctions of L and Lz - the spherical harmonics. Questions you will by able to answer by the end of today’s lecture 1. How are constants of motion used to label states? 2. How to use Ehrenfest theorem to determine whether the physical quantity is a constant of motion (i.e. does not change in time)? 3. What is the connection between symmetries and constant of motion? 4. What are the properties of “conservative systems”? 5. What is the dispersion relation for a free particle, why are E and k used as labels? 6. How to derive the orbital angular momentum observable? 7. How to check if a given vector observable is in fact an angular momentum? References: 1. Introductory Quantum and Statistical Mechanics, Hagelstein, Senturia, Orlando. 2. Principles of Quantum Mechanics, Shankar. 3. http://mathworld.wolfram.com/SphericalHarmonic.html 1 Symmetries, conserved quantities and constants of motion – how do we identify and label states (good quantum numbers) The connection between symmetries and conserved quantities: In the previous section we showed that the Hamiltonian function plays a major role in our understanding of quantum mechanics using it we could find both the eigenfunctions of the Hamiltonian and the time evolution of the system. What do we mean by when we say an object is symmetric?�What we mean is that if we take the object perform a particular operation on it and then compare the result to the initial situation they are indistinguishable.�When one speaks of a symmetry it is critical to state symmetric with respect to which operation.
    [Show full text]
  • Conservation of Total Momentum TM1
    L3:1 Conservation of total momentum TM1 We will here prove that conservation of total momentum is a consequence Taylor: 268-269 of translational invariance. Consider N particles with positions given by . If we translate the whole system with some distance we get The potential energy must be una ected by the translation. (We move the "whole world", no relative position change.) Also, clearly all velocities are independent of the translation Let's consider to be in the x-direction, then Using Lagrange's equation, we can rewrite each derivative as I.e., total momentum is conserved! (Clearly there is nothing special about the x-direction.) L3:2 Generalized coordinates, velocities, momenta and forces Gen:1 Taylor: 241-242 With we have i.e., di erentiating w.r.t. gives us the momentum in -direction. generalized velocity In analogy we de✁ ne the generalized momentum and we say that and are conjugated variables. Note: Def. of gen. force We also de✁ ne the generalized force from Taylor 7.15 di ers from def in HUB I.9.17. such that generalized rate of change of force generalized momentum Note: (generalized coordinate) need not have dimension length (generalized velocity) velocity momentum (generalized momentum) (generalized force) force Example: For a system which rotates around the z-axis we have for V=0 s.t. The angular momentum is thus conjugated variable to length dimensionless length/time 1/time mass*length/time mass*(length)^2/time (torque) energy Cyclic coordinates = ignorable coordinates and Noether's theorem L3:3 Cyclic:1 We have de ned the generalized momentum Taylor: From EL we have 266-267 i.e.
    [Show full text]