Continuous Symmetries and Conservation Laws. Noether's

Total Page:16

File Type:pdf, Size:1020Kb

Continuous Symmetries and Conservation Laws. Noether's Nature always creates the best of all options. Aristoteles (384 BC–322 BC) 8 Continuous Symmetries and Conservation Laws. Noether’s Theorem In many physical systems, the action is invariant under some continuous set of transformations. In such systems, there exist local and global conservation laws analogous to current and charge conservation in electrodynamics. The analogs of the charges can be used to generate the symmetry transformation, from which they were derived, with the help of Poisson brackets, or after quantization, with the help of commutators. 8.1 Point Mechanics Consider a simple mechanical system with a generic action tb A = dt L(q(t), q˙(t), t). (8.1) Zta 8.1.1 Continuous Symmetries and Conservation Law Suppose A is invariant under a continuous set of transformations of the dynamical variables: q(t) → q′(t)= f(q(t), q˙(t)), (8.2) where f(q(t), q˙(t)) is some functional of q(t). Such transformations are called sym- metry transformations. Thereby it is important that the equations of motion are not used when establishing the invariance of the action under (8.2). If the action is subjected successively to two symmetry transformations, the re- sult is again a symmetry transformation. Thus, symmetry transformations form a group called the symmetry group of the system. For infinitesimal symmetry trans- formations (8.2), the difference ′ δsq(t) ≡ q (t) − q(t) (8.3) will be called a symmetry variation. It has the general form δsq(t)= ǫ∆(q(t), q˙(t), t). (8.4) 619 620 8 Continuous Symmetries and Conservation Laws. Noether’s Theorem Symmetry variations must not be confused with ordinary variations δq(t) used in Section 1.1 to derive the Euler-Lagrange equations (1.8). While the ordinary vari- ations δq(t) vanish at initial and final times, δq(tb) = δq(ta) = 0 [recall (1.4)], the symmetry variations δsq(t) are usually nonzero at the ends. Let us calculate the change of the action under a symmetry variation (8.4). Using the chain rule of differentiation and an integration by parts, we obtain tb tb ∂L ∂L ∂L δsA = dt − ∂t δsq(t)+ δsq(t) . (8.5) ta "∂q(t) ∂q˙(t)# ∂q˙(t) Z ta For orbits q(t) that satisfy the Euler-Lagrange equations (1.8), only boundary terms survive, and we are left with t ∂L a δsA = ǫ ∆(q, q,˙ t) . (8.6) ∂q˙ tb Under the symmetry assumption, δsA vanishes for any orbit q(t), implying that the quantity ∂L Q(t) ≡ ∆(q, q,˙ t) (8.7) ∂q˙ is the same at times t = ta and t = tb. Since tb is arbitrary, Q(t) is independent of the time t, i.e., it satisfies Q(t) ≡ Q. (8.8) It is a conserved quantity, a constant of motion. The expression on the right-hand side of (8.7) is called Noether charge. The statement can be generalized to transformations δsq(t) for which the action is not directly invariant but its symmetry variation is equal to an arbitrary boundary term: tb δsA = ǫ Λ(q, q,˙ t) . (8.9) ta In this case, ∂L Q(t)= ∆(q, q,˙ t) − Λ(q, q,˙ t) (8.10) ∂q˙ is a conserved Noether charge. It is also possible to derive the constant of motion (8.10) without invoking the action, but starting from the Lagrangian. For it we evaluate the symmetry variation as follows: ∂L ∂L d ∂L δsL≡L (q+δsq, q˙+δsq˙) − L(q, q˙)= −∂t δsq(t)+ δsq(t) . (8.11) "∂q(t) ∂q˙(t)# dt "∂q˙(t) # 8.1 Point Mechanics 621 On account of the Euler-Lagrange equations (1.8), the first term on the right-hand side vanishes as before, and only the last term survives. The assumption of invariance of the action up to a possible surface term in Eq. (8.9) is equivalent to assuming that the symmetry variation of the Lagrangian is a total time derivative of some function Λ(q, q,˙ t): d δ L(q, q,˙ t)= ǫ Λ(q, q,˙ t). (8.12) s dt Inserting this into the left-hand side of (8.11), we find d ∂L ǫ ∆(q, q,˙ t) − Λ(q, q,˙ t) =0, (8.13) dt " ∂q˙ # thus recovering again the conserved Noether charge (8.8). The existence of a conserved quantity for every continuous symmetry is the content of Noether’s theorem [1]. 8.1.2 Alternative Derivation Let us do the substantial variation in Eq. (8.5) explicitly, and change a classical orbit qc(t), that extremizes the action, by an arbitrary variation δaq(t). If this does not vanish at the boundaries, the action changes by a pure boundary term that follows directly from (8.5): t ∂L b δaA = δaq . (8.14) ∂q˙ ta From this equation we can derive Noether’s theorem in yet another way. Suppose we subject a classical orbit to a new type of symmetry variation, to be called local symmetry transformations, which generalizes the previous symmetry variations (8.4) by making the parameter ǫ time-dependent: t δsq(t)= ǫ(t)∆(q(t), q˙(t), t). (8.15) t The superscript t of δsq(t) indicates the new time dependence in the parameter ǫ(t). These variations may be considered as a special set of the general variations δaq(t) t introduced above. Thus also δsA must be a pure boundary term of the type (8.14). For the subsequent discussion it is useful to introduce the infinitesimally transformed orbit ǫ t q (t) ≡ q(t)+ δsq(t)= q(t)+ ǫ(t)∆(q(t), q˙(t), t), (8.16) and the associated Lagrangian: Lǫ ≡ L(qǫ(t), q˙ǫ(t)). (8.17) Using the time-dependent parameter ǫ(t), the local symmetry variation of the action can be written as tb tb ǫ ǫ ǫ t ∂L d ∂L d ∂L δsA = dt − ǫ(t)+ ǫ(t) . (8.18) ta "∂ǫ(t) dt ∂ǫ˙(t)# dt " ∂ǫ˙ # Z ta 622 8 Continuous Symmetries and Conservation Laws. Noether’s Theorem Along the classical orbits, the action is extremal and satisfies the equation δA =0, (8.19) δǫ(t) which translates for a local action to an Euler-Lagrange type of equation: ∂Lǫ d ∂Lǫ − =0. (8.20) ∂ǫ(t) dt ∂ǫ˙(t) This can also be checked explicitly by differentiating (8.17) according to the chain rule of differentiation: ∂Lǫ ∂Lǫ ∂Lǫ = ∆(q, q,˙ t)+ ∆(˙ q, q,˙ t); (8.21) ∂ǫ(t) ∂q(t) ∂q˙(t) ∂Lǫ ∂Lǫ = ∆(q, q,˙ t), (8.22) ∂ǫ˙(t) ∂q˙(t) and inserting on the right-hand side the ordinary Euler-Lagrange equations (1.8). We now invoke the symmetry assumption that the action is a pure surface term under the time-independent transformations (8.15). This implies that ∂Lǫ d = Λ. (8.23) ∂ǫ dt Combining this with (8.20), we derive a conservation law for the charge: ∂Lǫ Q = − Λ. (8.24) ∂ǫ˙ Inserting here Eq. (8.22), we find that this is the same charge as that derived by the previous method. 8.2 Displacement and Energy Conservation As a simple but physically important example consider the case that the Lagrangian does not depend explicitly on time, i.e., that L(q, q,˙ t) ≡ L(q, q˙). Let us perform a time translation on the coordinate frame: t′ = t − ǫ. (8.25) In the new coordinate frame, the same orbit has the new description q˙(t′)= q(t), (8.26) i.e., the orbitq ˙(t) at the translated time t′ is precisely the same as the orbit q(t) at the original time t. If we replace the argument ofq ˙(t) in (8.26) by t′, we describe a 8.2 Displacement and Energy Conservation 623 time-translated orbit in terms of the original coordinates. This implies the symmetry variation of the form (8.4): ′ ′ δsq(t) = q (t) − q(t)= q(t + ǫ) − q(t) = q(t′)+ ǫq˙(t′) − q(t)= ǫq˙(t). (8.27) The symmetry variation of the Lagrangian is in general ∂L ∂L δ L = L(q′(t), q˙′(t)) − L(q(t), q˙(t)) = δ q(t)+ δ q˙(t). (8.28) s ∂q s ∂q˙ s Inserting δsq(t) from (8.27) we find, without using the Euler-Lagrange equation, ∂L ∂L d δsL = ǫ q˙ + q¨ = ǫ L. (8.29) ∂q˙ ∂q˙ ! dt This has precisely the form of Eq. (8.12), with Λ = L as expected, since time translations are symmetry transformations. Here the function Λ in (8.12) happens to coincide with the Lagrangian. According to Eq. (8.10), we find the Noether charge ∂L Q = q˙ − L(q, q˙) (8.30) ∂q˙ to be a constant of motion. This is recognized as the Legendre transform of the Lagrangian which is, of course, the Hamiltonian of the system. s Let us briefly check how this Noether charge is obtained from the alternative formula (8.10). The time-dependent symmetry variation is here t δsq(t) = ǫ(t)q ˙(t), (8.31) under which the Lagrangian is changed by ∂L ∂L ∂Lǫ ∂Lǫ δtL = ǫq˙ + (˙ǫq˙ + ǫq¨)= ǫ + ǫ,˙ (8.32) s ∂q ∂q˙ ∂ǫ˙ ∂ǫ˙ with ∂Lǫ ∂L = q˙ (8.33) ∂ǫ˙ ∂q˙ and ∂Lǫ ∂L ∂L d = q˙ + ǫq¨ = L. (8.34) ∂ǫ ∂q ∂q˙ dt This shows that time translations fulfill the symmetry condition (8.23), and that the Noether charge (8.24) coincides with the Hamiltonian found in Eq. (8.10). 624 8 Continuous Symmetries and Conservation Laws. Noether’s Theorem 8.3 Momentum and Angular Momentum While the conservation law of energy follows from the symmetry of the action under time translations, conservation laws of momentum and angular momentum are found if the action is invariant under translations and rotations.
Recommended publications
  • Newtonian Gravity and Special Relativity 12.1 Newtonian Gravity
    Physics 411 Lecture 12 Newtonian Gravity and Special Relativity Lecture 12 Physics 411 Classical Mechanics II Monday, September 24th, 2007 It is interesting to note that under Lorentz transformation, while electric and magnetic fields get mixed together, the force on a particle is identical in magnitude and direction in the two frames related by the transformation. Indeed, that was the motivation for looking at the manifestly relativistic structure of Maxwell's equations. The idea was that Maxwell's equations and the Lorentz force law are automatically in accord with the notion that observations made in inertial frames are physically equivalent, even though observers may disagree on the names of these forces (electric or magnetic). Today, we will look at a force (Newtonian gravity) that does not have the property that different inertial frames agree on the physics. That will lead us to an obvious correction that is, qualitatively, a prediction of (linearized) general relativity. 12.1 Newtonian Gravity We start with the experimental observation that for a particle of mass M and another of mass m, the force of gravitational attraction between them, according to Newton, is (see Figure 12.1): G M m F = − RR^ ≡ r − r 0: (12.1) r 2 From the force, we can, by analogy with electrostatics, construct the New- tonian gravitational field and its associated point potential: GM GM G = − R^ = −∇ − : (12.2) r 2 r | {z } ≡φ 1 of 7 12.2. LINES OF MASS Lecture 12 zˆ m !r M !r ! yˆ xˆ Figure 12.1: Two particles interacting via the Newtonian gravitational force.
    [Show full text]
  • 1-Crystal Symmetry and Classification-1.Pdf
    R. I. Badran Solid State Physics Fundamental types of lattices and crystal symmetry Crystal symmetry: What is a symmetry operation? It is a physical operation that changes the positions of the lattice points at exactly the same places after and before the operation. In other words, it is an operation when applied to an object leaves it apparently unchanged. e.g. A translational symmetry is occurred, for example, when the function sin x has a translation through an interval x = 2 leaves it apparently unchanged. Otherwise a non-symmetric operation can be foreseen by the rotation of a rectangle through /2. There are two groups of symmetry operations represented by: a) The point groups. b) The space groups (these are a combination of point groups with translation symmetry elements). There are 230 space groups exhibited by crystals. a) When the symmetry operations in crystal lattice are applied about a lattice point, the point groups must be used. b) When the symmetry operations are performed about a point or a line in addition to symmetry operations performed by translations, these are called space group symmetry operations. Types of symmetry operations: There are five types of symmetry operations and their corresponding elements. 85 R. I. Badran Solid State Physics Note: The operation and its corresponding element are denoted by the same symbol. 1) The identity E: It consists of doing nothing. 2) Rotation Cn: It is the rotation about an axis of symmetry (which is called "element"). If the rotation through 2/n (where n is integer), and the lattice remains unchanged by this rotation, then it has an n-fold axis.
    [Show full text]
  • 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]
  • Chapter 5 the Relativistic Point Particle
    Chapter 5 The Relativistic Point Particle To formulate the dynamics of a system we can write either the equations of motion, or alternatively, an action. In the case of the relativistic point par- ticle, it is rather easy to write the equations of motion. But the action is so physical and geometrical that it is worth pursuing in its own right. More importantly, while it is difficult to guess the equations of motion for the rela- tivistic string, the action is a natural generalization of the relativistic particle action that we will study in this chapter. We conclude with a discussion of the charged relativistic particle. 5.1 Action for a relativistic point particle How can we find the action S that governs the dynamics of a free relativis- tic particle? To get started we first think about units. The action is the Lagrangian integrated over time, so the units of action are just the units of the Lagrangian multiplied by the units of time. The Lagrangian has units of energy, so the units of action are L2 ML2 [S]=M T = . (5.1.1) T 2 T Recall that the action Snr for a free non-relativistic particle is given by the time integral of the kinetic energy: 1 dx S = mv2(t) dt , v2 ≡ v · v, v = . (5.1.2) nr 2 dt 105 106 CHAPTER 5. THE RELATIVISTIC POINT PARTICLE The equation of motion following by Hamilton’s principle is dv =0. (5.1.3) dt The free particle moves with constant velocity and that is the end of the story.
    [Show full text]
  • Arxiv:1910.10745V1 [Cond-Mat.Str-El] 23 Oct 2019 2.2 Symmetry-Protected Time Crystals
    A Brief History of Time Crystals Vedika Khemania,b,∗, Roderich Moessnerc, S. L. Sondhid aDepartment of Physics, Harvard University, Cambridge, Massachusetts 02138, USA bDepartment of Physics, Stanford University, Stanford, California 94305, USA cMax-Planck-Institut f¨urPhysik komplexer Systeme, 01187 Dresden, Germany dDepartment of Physics, Princeton University, Princeton, New Jersey 08544, USA Abstract The idea of breaking time-translation symmetry has fascinated humanity at least since ancient proposals of the per- petuum mobile. Unlike the breaking of other symmetries, such as spatial translation in a crystal or spin rotation in a magnet, time translation symmetry breaking (TTSB) has been tantalisingly elusive. We review this history up to recent developments which have shown that discrete TTSB does takes place in periodically driven (Floquet) systems in the presence of many-body localization (MBL). Such Floquet time-crystals represent a new paradigm in quantum statistical mechanics — that of an intrinsically out-of-equilibrium many-body phase of matter with no equilibrium counterpart. We include a compendium of the necessary background on the statistical mechanics of phase structure in many- body systems, before specializing to a detailed discussion of the nature, and diagnostics, of TTSB. In particular, we provide precise definitions that formalize the notion of a time-crystal as a stable, macroscopic, conservative clock — explaining both the need for a many-body system in the infinite volume limit, and for a lack of net energy absorption or dissipation. Our discussion emphasizes that TTSB in a time-crystal is accompanied by the breaking of a spatial symmetry — so that time-crystals exhibit a novel form of spatiotemporal order.
    [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]
  • Higher Dimensional Conundra
    Higher Dimensional Conundra Steven G. Krantz1 Abstract: In recent years, especially in the subject of harmonic analysis, there has been interest in geometric phenomena of RN as N → +∞. In the present paper we examine several spe- cific geometric phenomena in Euclidean space and calculate the asymptotics as the dimension gets large. 0 Introduction Typically when we do geometry we concentrate on a specific venue in a particular space. Often the context is Euclidean space, and often the work is done in R2 or R3. But in modern work there are many aspects of analysis that are linked to concrete aspects of geometry. And there is often interest in rendering the ideas in Hilbert space or some other infinite dimensional setting. Thus we want to see how the finite-dimensional result in RN changes as N → +∞. In the present paper we study some particular aspects of the geometry of RN and their asymptotic behavior as N →∞. We choose these particular examples because the results are surprising or especially interesting. We may hope that they will lead to further studies. It is a pleasure to thank Richard W. Cottle for a careful reading of an early draft of this paper and for useful comments. 1 Volume in RN Let us begin by calculating the volume of the unit ball in RN and the surface area of its bounding unit sphere. We let ΩN denote the former and ωN−1 denote the latter. In addition, we let Γ(x) be the celebrated Gamma function of L. Euler. It is a helpful intuition (which is literally true when x is an 1We are happy to thank the American Institute of Mathematics for its hospitality and support during this work.
    [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]
  • Chapter 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [Show full text]
  • Electron-Positron Pairs in Physics and Astrophysics
    Electron-positron pairs in physics and astrophysics: from heavy nuclei to black holes Remo Ruffini1,2,3, Gregory Vereshchagin1 and She-Sheng Xue1 1 ICRANet and ICRA, p.le della Repubblica 10, 65100 Pescara, Italy, 2 Dip. di Fisica, Universit`adi Roma “La Sapienza”, Piazzale Aldo Moro 5, I-00185 Roma, Italy, 3 ICRANet, Universit´ede Nice Sophia Antipolis, Grand Chˆateau, BP 2135, 28, avenue de Valrose, 06103 NICE CEDEX 2, France. Abstract Due to the interaction of physics and astrophysics we are witnessing in these years a splendid synthesis of theoretical, experimental and observational results originating from three fundamental physical processes. They were originally proposed by Dirac, by Breit and Wheeler and by Sauter, Heisenberg, Euler and Schwinger. For almost seventy years they have all three been followed by a continued effort of experimental verification on Earth-based experiments. The Dirac process, e+e 2γ, has been by − → far the most successful. It has obtained extremely accurate experimental verification and has led as well to an enormous number of new physics in possibly one of the most fruitful experimental avenues by introduction of storage rings in Frascati and followed by the largest accelerators worldwide: DESY, SLAC etc. The Breit–Wheeler process, 2γ e+e , although conceptually simple, being the inverse process of the Dirac one, → − has been by far one of the most difficult to be verified experimentally. Only recently, through the technology based on free electron X-ray laser and its numerous applications in Earth-based experiments, some first indications of its possible verification have been reached. The vacuum polarization process in strong electromagnetic field, pioneered by Sauter, Heisenberg, Euler and Schwinger, introduced the concept of critical electric 2 3 field Ec = mec /(e ).
    [Show full text]
  • Aspects of Symplectic Dynamics and Topology: Gauge Anomalies, Chiral Kinetic Theory and Transfer Matrices
    ASPECTS OF SYMPLECTIC DYNAMICS AND TOPOLOGY: GAUGE ANOMALIES, CHIRAL KINETIC THEORY AND TRANSFER MATRICES BY VATSAL DWIVEDI DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Physics in the Graduate College of the University of Illinois at Urbana-Champaign, 2017 Urbana, Illinois Doctoral Committee: Associate Professor Shinsei Ryu, Chair Professor Michael Stone, Director of Research Professor James N. Eckstein Professor Jared Bronski Abstract This thesis presents some work on two quite disparate kinds of dynamical systems described by Hamiltonian dynamics. The first part describes a computation of gauge anomalies and their macroscopic effects in a semiclassical picture. The geometric (symplectic) formulation of classical mechanics is used to describe the dynamics of Weyl fermions in even spacetime dimensions, the only quantum input to the symplectic form being the Berry curvature that encodes the spin-momentum locking. The (semi-)classical equations of motion are used in a kinetic theory setup to compute the gauge and singlet currents, whose conservation laws reproduce the nonabelian gauge and singlet anomalies. Anomalous contributions to the hydrodynamic currents for a gas of Weyl fermions at a finite temperature and chemical potential are also calculated, and are in agreement with similar results in literature which were obtained using thermodynamic and/or quantum field theoretical arguments. The second part describes a generalized transfer matrix formalism for noninteracting tight-binding models. The formalism is used to study the bulk and edge spectra, both of which are encoded in the spectrum of the transfer matrices, for some of the common tight-binding models for noninteracting electronic topological phases of matter.
    [Show full text]