5 the Dirac Equation and Spinors

Total Page:16

File Type:pdf, Size:1020Kb

5 the Dirac Equation and Spinors 5 The Dirac Equation and Spinors In this section we develop the appropriate wavefunctions for fundamental fermions and bosons. 5.1 Notation Review The three dimension differential operator is : ∂ ∂ ∂ = , , (5.1) ∂x ∂y ∂z We can generalise this to four dimensions ∂µ: 1 ∂ ∂ ∂ ∂ ∂ = , , , (5.2) µ c ∂t ∂x ∂y ∂z 5.2 The Schr¨odinger Equation First consider a classical non-relativistic particle of mass m in a potential U. The energy-momentum relationship is: p2 E = + U (5.3) 2m we can substitute the differential operators: ∂ Eˆ i pˆ i (5.4) → ∂t →− to obtain the non-relativistic Schr¨odinger Equation (with = 1): ∂ψ 1 i = 2 + U ψ (5.5) ∂t −2m For U = 0, the free particle solutions are: iEt ψ(x, t) e− ψ(x) (5.6) ∝ and the probability density ρ and current j are given by: 2 i ρ = ψ(x) j = ψ∗ ψ ψ ψ∗ (5.7) | | −2m − with conservation of probability giving the continuity equation: ∂ρ + j =0, (5.8) ∂t · Or in Covariant notation: µ µ ∂µj = 0 with j =(ρ,j) (5.9) The Schr¨odinger equation is 1st order in ∂/∂t but second order in ∂/∂x. However, as we are going to be dealing with relativistic particles, space and time should be treated equally. 25 5.3 The Klein-Gordon Equation For a relativistic particle the energy-momentum relationship is: p p = p pµ = E2 p 2 = m2 (5.10) · µ − | | Substituting the equation (5.4), leads to the relativistic Klein-Gordon equation: ∂2 + 2 ψ = m2ψ (5.11) −∂t2 The free particle solutions are plane waves: ip x i(Et p x) ψ e− · = e− − · (5.12) ∝ The Klein-Gordon equation successfully describes spin 0 particles in relativistic quan- tum field theory. There are problems with the interpretation of the positive and negative energy solutions of the Klein-Gordon equation, E = p2 + m2, since the negative energy solutions have negative probability densities ρ.± 5.4 The Dirac Equation The problems with the Klein-Gordon equation led Dirac to search for an alternative relativistic wave equation in 1928, in which the time and space derivatives are first order. The Dirac equation can be thought of in terms of a “square root” of the Klein-Gordon equation. In covariant form it is written: ∂ iγ0 + iγ m ψ =0 (iγµ∂ m) ψ = 0 (5.13) ∂t · − µ − where we have introduced the coefficients γµ =(γ0,γ )=(γ0, γ1, γ2, γ3), which have to be determined. As we will see in equation (5.20), the Dirac equation is simply four coupled differential equations, describing a wavefunction ψ with four components. 5.5 The Gamma Matrices To find what the γµ, µ =0, 1, 2, 3 objects are, we first multiply the Dirac equation by its conjugate equation: 0 ∂ 0 ∂ ψ† iγ iγ m iγ + iγ m ψ = 0 (5.14) − ∂t − · − ∂t · − and demand that this be consistent with the Klein-Gordon equation, (5.11). This leads to the following conditions on the γµ: (γ0)2 =1, (γi)2 = 1 γµγν + γνγµ =0forµ = ν − with i =1, 2, 3, µ, ν =0, 1, 2, 3 (5.15) 26 Equivalently in terms of anticommutation relations and the metric tensor (equation (3.3)): γµ, γν = γµ, γν + γν, γµ =2gµν µ, ν =0, 1, 2, 3 (5.16) { } The simplest solution for the γµ, that satisfies these anticommutation relations, are 4 4 unitary matrices. We will use the following representation for the γ matrices: × I0 0 σi γ0 = γi = (5.17) 0 I σi 0 − − where I denotes a 2 2 identity matrix, 0 denotes a 2 2 null matrix, and the σi are the Pauli spin matrices× : × 01 0 i 10 σ = σ = σ = (5.18) x 10 y i −0 z 0 1 − Let’s write out the gamma matrices in full: 10 0 0 0001 01 0 0 0010 γ0 = γ1 = 00 10 0 100 − − 10 0 1 1000 − − 000 i 0010 00i −0 000 1 γ2 = γ3 = (5.19) 0 i 00 100− 0 − i 00 0 0100 − Please note, despite the µ superscript, the γµ are not four vectors. However they do remain constant under Lorentz transforms. Finally let’s write out the Dirac Equation in full: i ∂ m 0 i ∂ i ∂ + ∂ 1 ∂t − ∂z ∂x ∂y ψ 0 0 i ∂ mi∂ ∂ i ∂ 2 ∂t ∂x ∂y ∂z ψ 0 ∂ ∂ − ∂ ∂ − − 3 = (5.20) i ∂z i ∂x ∂y i ∂t m 0 ψ 0 −∂ ∂ − ∂− − − ∂ 4 i + i 0 i m ψ 0 − ∂x ∂y ∂z − ∂t − 5.6 Spinors The Dirac equation describes the behaviour of spin-1/2 fermions in relativistic quantum field theory. For a free fermion the wavefunction is the product of a plane wave and a Dirac spinor, u(pµ): µ µ ip x ψ(x )=u(p )e− · (5.21) Substituting the fermion wavefunction, ψ, into the Dirac equation: (γµp m)u(p) = 0 (5.22) µ − 27 For a particle at rest, p = 0, we find the following equations: ∂ mI 0 iγ0 m ψ = γ0E m ψ =0 Euˆ = u (5.23) ∂t − − 0 mI − The solutions are four eigenspinors: 1 0 0 0 0 1 0 0 u1 = u2 = u3 = u4 = (5.24) 0 0 1 0 0 0 0 1 and the associated wavefunctions of the fermion is: 1 imt 1 2 imt 2 3 +imt 3 4 +imt 4 ψ = e− u ψ = e− u ψ = e u ψ = e u (5.25) Note that the spinors are 1 4 column matrices, and that there are four possible states. The spinors are, however,× not four-vectors: the four components do not represent t, x, y, z. The four components are a suprise: we would expect only two spin states for a spin-1/2 fermion! Note also the change of sign in the exponents of the plane waves in the states ψ3 and ψ4. The four solutions in equations (5.24) and (5.25) describe two different spin states ( and ) with E = m, and two spin states with E = m. ↑ ↓ − 5.7 Negative Energy Solutions & Antimatter To describe the negative energy states, Dirac postulated that an electron in a positive energy state is produced from the vacuum accompanied by a hole with negative energy. The hole corresponds to a physical antiparticle, the positron, with charge +e. Another interpretation (Feynman-St¨uckelberg) is that the E = m solutions can either describe a negative energy particle which propagates backwards− in time, or a positive energy antiparticle propagating forward in time: i[( E)( t) ( p) ( x)] i[Et p x] e− − − − − · − = e− − · (5.26) 5.8 Spinors for Moving Particles For a moving particle, p = 0 the Dirac equation becomes (using (5.13) and (5.17)): µ E m σ p uA (γ pµ m) uA uB = − − · = 0 (5.27) − σ p E m uB · − − where u and u denote the 1 2 upper and lower components of u respectively. The A B × equations for uA and uB are coupled: σ p σ p u = · u u = · u (5.28) A E m B B E + m A − 28 1 0 The solutions are obtained by successively setting: u = , u = , u = A 0 A 1 B 1 0 , and u = , to give: 0 B 1 1 0 0 1 u1 = u2 = p /(E + m) (p ip )/(E + m) z x − y (p + ip )/(E + m) p /(E + m) x y − z pz/( E + m) ( px + ipy)/( E + m) ( p− ip−)/( E + m) − p /( E +−m) u3 = x y u4 = z (5.29) − − 1 − −0 0 1 The u1 and u2 solutions describe an electron of energy E =+ m2 + p 2, and momen- tum p: 1 µ ip x 2 µ ip x ψ = u (p )e− · ψ = u (p )e− · (5.30) The u3 and u4 of equation (5.29) describe a positron of energy E = m2 + p 2, and momentum p. − It is usual to change to the spinors v2(p) u3( p) and v1(p) u4( p) to describe these positive energy antiparticle states, E =+≡ −m2 + p 2 ≡ − pz/(E + m) 2 µ 3 µ (px + ipy)/(E + m) 2 µ ip x 3 µ i( p) x v (p ) u ( p )= ψ = v (p )e− · = u ( p )e − · ≡ − 1 − 0 (p ip )/(E + m) x − y 1 µ 4 µ pz/(E + m) 1 µ ip x 4 µ i( p) x v (p ) u ( p )= − ψ = v (p )e− · = u ( p )e − · ≡ − 0 − 1 (5.31) The u and v are the solutions of: (iγµp m)u =0 (iγµp + m)v = 0 (5.32) µ − µ 5.9 Spin and Helicity The two different solutions for each of the fermions and antifermions corresponds to two 1 µ ip x possible spin states. For a fermion with momentum p along the z-axis, ψ = u (p )e− · 2 µ ip x describes a spin-up fermion and ψ = u (p )e− · describes a spin-down fermion. For 1 µ ip x an antifermion with momentum p along the z-axis, ψ = v (p )e− · describes a spin-up 2 µ ip x antifermion and ψ = v (p )e− · describes a spin-down antifermion. 29 Figure 5.1: Helicity eigenstates for a particle or antiparticle travelling along the +z axis. 1 2 1 2 The u ,u ,v ,v spinors are only eigenstates of Sˆz for momentum p along the z-axis. There’s nothing special about projecting out the component of spin along the z-axis, that’s just the conventional choice. For our purposes it makes more sense to project the spin along the particle’s direction of flight, this defines the helicity, h of the particle.
Recommended publications
  • 1 Notation for States
    The S-Matrix1 D. E. Soper2 University of Oregon Physics 634, Advanced Quantum Mechanics November 2000 1 Notation for states In these notes we discuss scattering nonrelativistic quantum mechanics. We will use states with the nonrelativistic normalization h~p |~ki = (2π)3δ(~p − ~k). (1) Recall that in a relativistic theory there is an extra factor of 2E on the right hand side of this relation, where E = [~k2 + m2]1/2. We will use states in the “Heisenberg picture,” in which states |ψ(t)i do not depend on time. Often in quantum mechanics one uses the Schr¨odinger picture, with time dependent states |ψ(t)iS. The relation between these is −iHt |ψ(t)iS = e |ψi. (2) Thus these are the same at time zero, and the Schrdinger states obey d i |ψ(t)i = H |ψ(t)i (3) dt S S In the Heisenberg picture, the states do not depend on time but the op- erators do depend on time. A Heisenberg operator O(t) is related to the corresponding Schr¨odinger operator OS by iHt −iHt O(t) = e OS e (4) Thus hψ|O(t)|ψi = Shψ(t)|OS|ψ(t)iS. (5) The Heisenberg picture is favored over the Schr¨odinger picture in the case of relativistic quantum mechanics: we don’t have to say which reference 1Copyright, 2000, D. E. Soper [email protected] 1 frame we use to define t in |ψ(t)iS. For operators, we can deal with local operators like, for instance, the electric field F µν(~x, t).
    [Show full text]
  • Path Integrals in Quantum Mechanics
    Path Integrals in Quantum Mechanics Dennis V. Perepelitsa MIT Department of Physics 70 Amherst Ave. Cambridge, MA 02142 Abstract We present the path integral formulation of quantum mechanics and demon- strate its equivalence to the Schr¨odinger picture. We apply the method to the free particle and quantum harmonic oscillator, investigate the Euclidean path integral, and discuss other applications. 1 Introduction A fundamental question in quantum mechanics is how does the state of a particle evolve with time? That is, the determination the time-evolution ψ(t) of some initial | i state ψ(t ) . Quantum mechanics is fully predictive [3] in the sense that initial | 0 i conditions and knowledge of the potential occupied by the particle is enough to fully specify the state of the particle for all future times.1 In the early twentieth century, Erwin Schr¨odinger derived an equation specifies how the instantaneous change in the wavefunction d ψ(t) depends on the system dt | i inhabited by the state in the form of the Hamiltonian. In this formulation, the eigenstates of the Hamiltonian play an important role, since their time-evolution is easy to calculate (i.e. they are stationary). A well-established method of solution, after the entire eigenspectrum of Hˆ is known, is to decompose the initial state into this eigenbasis, apply time evolution to each and then reassemble the eigenstates. That is, 1In the analysis below, we consider only the position of a particle, and not any other quantum property such as spin. 2 D.V. Perepelitsa n=∞ ψ(t) = exp [ iE t/~] n ψ(t ) n (1) | i − n h | 0 i| i n=0 X This (Hamiltonian) formulation works in many cases.
    [Show full text]
  • 22 Scattering in Supersymmetric Matter Chern-Simons Theories at Large N
    2 2 scattering in supersymmetric matter ! Chern-Simons theories at large N Karthik Inbasekar 10th Asian Winter School on Strings, Particles and Cosmology 09 Jan 2016 Scattering in CS matter theories In QFT, Crossing symmetry: analytic continuation of amplitudes. Particle-antiparticle scattering: obtained from particle-particle scattering by analytic continuation. Naive crossing symmetry leads to non-unitary S matrices in U(N) Chern-Simons matter theories.[ Jain, Mandlik, Minwalla, Takimi, Wadia, Yokoyama] Consistency with unitarity required Delta function term at forward scattering. Modified crossing symmetry rules. Conjecture: Singlet channel S matrices have the form sin(πλ) = 8πpscos(πλ)δ(θ)+ i S;naive(s; θ) S πλ T S;naive: naive analytic continuation of particle-particle scattering. T Scattering in U(N) CS matter theories at large N Particle: fund rep of U(N), Antiparticle: antifund rep of U(N). Fundamental Fundamental Symm(Ud ) Asymm(Ue ) ⊗ ! ⊕ Fundamental Antifundamental Adjoint(T ) Singlet(S) ⊗ ! ⊕ C2(R1)+C2(R2)−C2(Rm) Eigenvalues of Anyonic phase operator νm = 2κ 1 νAsym νSym νAdj O ;ν Sing O(λ) ∼ ∼ ∼ N ∼ symm, asymm and adjoint channels- non anyonic at large N. Scattering in the singlet channel is effectively anyonic at large N- naive crossing rules fail unitarity. Conjecture beyond large N: general form of 2 2 S matrices in any U(N) Chern-Simons matter theory ! sin(πνm) (s; θ) = 8πpscos(πνm)δ(θ)+ i (s; θ) S πνm T Universality and tests Delta function and modified crossing rules conjectured by Jain et al appear to be universal. Tests of the conjecture: Unitarity of the S matrix.
    [Show full text]
  • Arxiv:Math/0212058V2 [Math.DG] 18 Nov 2003 Nosbrusof Subgroups Into Rdc C.[Oc 00 Et .].Eape O Uhsae a Spaces Such for Examples 3.2])
    THE SPINOR BUNDLE OF RIEMANNIAN PRODUCTS FRANK KLINKER Abstract. In this note we compare the spinor bundle of a Riemannian mani- fold (M = M1 ×···×MN ,g) with the spinor bundles of the Riemannian factors (Mi,gi). We show that - without any holonomy conditions - the spinor bundle of (M,g) for a special class of metrics is isomorphic to a bundle obtained by tensoring the spinor bundles of (Mi,gi) in an appropriate way. For N = 2 and an one dimensional factor this construction was developed in [Baum 1989a]. Although the fact for general factors is frequently used in (at least physics) literature, a proof was missing. I would like to thank Shahram Biglari, Mario Listing, Marc Nardmann and Hans-Bert Rademacher for helpful comments. Special thanks go to Helga Baum, who pointed out some difficulties arising in the pseudo-Riemannian case. We consider a Riemannian manifold (M = M MN ,g), which is a product 1 ×···× of Riemannian spin manifolds (Mi,gi) and denote the projections on the respective factors by pi. Furthermore the dimension of Mi is Di such that the dimension of N M is given by D = i=1 Di. The tangent bundle of M is decomposed as P ∗ ∗ (1) T M = p Tx1 M p Tx MN . (x0,...,xN ) 1 1 ⊕···⊕ N N N We omit the projections and write TM = i=1 TMi. The metric g of M need not be the product metric of the metrics g on M , but L i i is assumed to be of the form c d (2) gab(x) = Ai (x)gi (xi)Ai (x), T Mi a cd b for D1 + + Di−1 +1 a,b D1 + + Di, 1 i N ··· ≤ ≤ ··· ≤ ≤ In particular, for those metrics the splitting (1) is orthogonal, i.e.
    [Show full text]
  • Old Supersymmetry As New Mathematics
    old supersymmetry as new mathematics PILJIN YI Korea Institute for Advanced Study with help from Sungjay Lee Atiyah-Singer Index Theorem ~ 1963 1975 ~ Bogomolnyi-Prasad-Sommerfeld (BPS) Calabi-Yau ~ 1978 1977 ~ Supersymmetry Calibrated Geometry ~ 1982 1982 ~ Index Thm by Path Integral (Alvarez-Gaume) (Harvey & Lawson) 1985 ~ Calabi-Yau Compactification 1988 ~ Mirror Symmetry 1992~ 2d Wall-Crossing / tt* (Cecotti & Vafa etc) Homological Mirror Symmetry ~ 1994 1994 ~ 4d Wall-Crossing (Seiberg & Witten) (Kontsevich) 1998 ~ Wall-Crossing is Bound State Dissociation (Lee & P.Y.) Stability & Derived Category ~ 2000 2000 ~ Path Integral Proof of Mirror Symmetry (Hori & Vafa) Wall-Crossing Conjecture ~ 2008 2008 ~ Konstevich-Soibelman Explained (conjecture by Kontsevich & Soibelman) (Gaiotto & Moore & Neitzke) 2011 ~ KS Wall-Crossing proved via Quatum Mechanics (Manschot , Pioline & Sen / Kim , Park, Wang & P.Y. / Sen) 2012 ~ S2 Partition Function as tt* (Jocker, Kumar, Lapan, Morrison & Romo /Gomis & Lee) quantum and geometry glued by superstring theory when can we perform path integrals exactly ? counting geometry with supersymmetric path integrals quantum and geometry glued by superstring theory Einstein this theory famously resisted quantization, however on the other hand, five superstring theories, with a consistent quantum gravity inside, live in 10 dimensional spacetime these superstring theories say, spacetime is composed of 4+6 dimensions with very small & tightly-curved (say, Calabi-Yau) 6D manifold sitting at each and every point of
    [Show full text]
  • Motion of the Reduced Density Operator
    Motion of the Reduced Density Operator Nicholas Wheeler, Reed College Physics Department Spring 2009 Introduction. Quantum mechanical decoherence, dissipation and measurements all involve the interaction of the system of interest with an environmental system (reservoir, measurement device) that is typically assumed to possess a great many degrees of freedom (while the system of interest is typically assumed to possess relatively few degrees of freedom). The state of the composite system is described by a density operator ρ which in the absence of system-bath interaction we would denote ρs ρe, though in the cases of primary interest that notation becomes unavailable,⊗ since in those cases the states of the system and its environment are entangled. The observable properties of the system are latent then in the reduced density operator ρs = tre ρ (1) which is produced by “tracing out” the environmental component of ρ. Concerning the specific meaning of (1). Let n) be an orthonormal basis | in the state space H of the (open) system, and N) be an orthonormal basis s !| " in the state space H of the (also open) environment. Then n) N) comprise e ! " an orthonormal basis in the state space H = H H of the |(closed)⊗| composite s e ! " system. We are in position now to write ⊗ tr ρ I (N ρ I N) e ≡ s ⊗ | s ⊗ | # ! " ! " ↓ = ρ tr ρ in separable cases s · e The dynamics of the composite system is generated by Hamiltonian of the form H = H s + H e + H i 2 Motion of the reduced density operator where H = h I s s ⊗ e = m h n m N n N $ | s| % | % ⊗ | % · $ | ⊗ $ | m,n N # # $% & % &' H = I h e s ⊗ e = n M n N M h N | % ⊗ | % · $ | ⊗ $ | $ | e| % n M,N # # $% & % &' H = m M m M H n N n N i | % ⊗ | % $ | ⊗ $ | i | % ⊗ | % $ | ⊗ $ | m,n M,N # # % &$% & % &'% & —all components of which we will assume to be time-independent.
    [Show full text]
  • Quantum Field Theory*
    Quantum Field Theory y Frank Wilczek Institute for Advanced Study, School of Natural Science, Olden Lane, Princeton, NJ 08540 I discuss the general principles underlying quantum eld theory, and attempt to identify its most profound consequences. The deep est of these consequences result from the in nite number of degrees of freedom invoked to implement lo cality.Imention a few of its most striking successes, b oth achieved and prosp ective. Possible limitation s of quantum eld theory are viewed in the light of its history. I. SURVEY Quantum eld theory is the framework in which the regnant theories of the electroweak and strong interactions, which together form the Standard Mo del, are formulated. Quantum electro dynamics (QED), b esides providing a com- plete foundation for atomic physics and chemistry, has supp orted calculations of physical quantities with unparalleled precision. The exp erimentally measured value of the magnetic dip ole moment of the muon, 11 (g 2) = 233 184 600 (1680) 10 ; (1) exp: for example, should b e compared with the theoretical prediction 11 (g 2) = 233 183 478 (308) 10 : (2) theor: In quantum chromo dynamics (QCD) we cannot, for the forseeable future, aspire to to comparable accuracy.Yet QCD provides di erent, and at least equally impressive, evidence for the validity of the basic principles of quantum eld theory. Indeed, b ecause in QCD the interactions are stronger, QCD manifests a wider variety of phenomena characteristic of quantum eld theory. These include esp ecially running of the e ective coupling with distance or energy scale and the phenomenon of con nement.
    [Show full text]
  • Introductory Lectures on Quantum Field Theory
    Introductory Lectures on Quantum Field Theory a b L. Álvarez-Gaumé ∗ and M.A. Vázquez-Mozo † a CERN, Geneva, Switzerland b Universidad de Salamanca, Salamanca, Spain Abstract In these lectures we present a few topics in quantum field theory in detail. Some of them are conceptual and some more practical. They have been se- lected because they appear frequently in current applications to particle physics and string theory. 1 Introduction These notes are based on lectures delivered by L.A.-G. at the 3rd CERN–Latin-American School of High- Energy Physics, Malargüe, Argentina, 27 February–12 March 2005, at the 5th CERN–Latin-American School of High-Energy Physics, Medellín, Colombia, 15–28 March 2009, and at the 6th CERN–Latin- American School of High-Energy Physics, Natal, Brazil, 23 March–5 April 2011. The audience on all three occasions was composed to a large extent of students in experimental high-energy physics with an important minority of theorists. In nearly ten hours it is quite difficult to give a reasonable introduction to a subject as vast as quantum field theory. For this reason the lectures were intended to provide a review of those parts of the subject to be used later by other lecturers. Although a cursory acquaintance with the subject of quantum field theory is helpful, the only requirement to follow the lectures is a working knowledge of quantum mechanics and special relativity. The guiding principle in choosing the topics presented (apart from serving as introductions to later courses) was to present some basic aspects of the theory that present conceptual subtleties.
    [Show full text]
  • Estimations of the Trace of Powers of Positive Self-Adjoint Operators by Extrapolation of the Moments∗
    Electronic Transactions on Numerical Analysis. ETNA Volume 39, pp. 144-155, 2012. Kent State University Copyright 2012, Kent State University. http://etna.math.kent.edu ISSN 1068-9613. ESTIMATIONS OF THE TRACE OF POWERS OF POSITIVE SELF-ADJOINT OPERATORS BY EXTRAPOLATION OF THE MOMENTS∗ CLAUDE BREZINSKI†, PARASKEVI FIKA‡, AND MARILENA MITROULI‡ Abstract. Let A be a positive self-adjoint linear operator on a real separable Hilbert space H. Our aim is to build estimates of the trace of Aq, for q ∈ R. These estimates are obtained by extrapolation of the moments of A. Applications of the matrix case are discussed, and numerical results are given. Key words. Trace, positive self-adjoint linear operator, symmetric matrix, matrix powers, matrix moments, extrapolation. AMS subject classifications. 65F15, 65F30, 65B05, 65C05, 65J10, 15A18, 15A45. 1. Introduction. Let A be a positive self-adjoint linear operator from H to H, where H is a real separable Hilbert space with inner product denoted by (·, ·). Our aim is to build estimates of the trace of Aq, for q ∈ R. These estimates are obtained by extrapolation of the integer moments (z, Anz) of A, for n ∈ N. A similar procedure was first introduced in [3] for estimating the Euclidean norm of the error when solving a system of linear equations, which corresponds to q = −2. The case q = −1, which leads to estimates of the trace of the inverse of a matrix, was studied in [4]; on this problem, see [10]. Let us mention that, when only positive powers of A are used, the Hilbert space H could be infinite dimensional, while, for negative powers of A, it is always assumed to be a finite dimensional one, and, obviously, A is also assumed to be invertible.
    [Show full text]
  • An S-Matrix for Massless Particles Arxiv:1911.06821V2 [Hep-Th]
    An S-Matrix for Massless Particles Holmfridur Hannesdottir and Matthew D. Schwartz Department of Physics, Harvard University, Cambridge, MA 02138, USA Abstract The traditional S-matrix does not exist for theories with massless particles, such as quantum electrodynamics. The difficulty in isolating asymptotic states manifests itself as infrared divergences at each order in perturbation theory. Building on insights from the literature on coherent states and factorization, we construct an S-matrix that is free of singularities order-by-order in perturbation theory. Factorization guarantees that the asymptotic evolution in gauge theories is universal, i.e. independent of the hard process. Although the hard S-matrix element is computed between well-defined few particle Fock states, dressed/coherent states can be seen to form as intermediate states in the calculation of hard S-matrix elements. We present a framework for the perturbative calculation of hard S-matrix elements combining Lorentz-covariant Feyn- man rules for the dressed-state scattering with time-ordered perturbation theory for the asymptotic evolution. With hard cutoffs on the asymptotic Hamiltonian, the cancella- tion of divergences can be seen explicitly. In dimensional regularization, where the hard cutoffs are replaced by a renormalization scale, the contribution from the asymptotic evolution produces scaleless integrals that vanish. A number of illustrative examples are given in QED, QCD, and N = 4 super Yang-Mills theory. arXiv:1911.06821v2 [hep-th] 24 Aug 2020 Contents 1 Introduction1 2 The hard S-matrix6 2.1 SH and dressed states . .9 2.2 Computing observables using SH ........................... 11 2.3 Soft Wilson lines . 14 3 Computing the hard S-matrix 17 3.1 Asymptotic region Feynman rules .
    [Show full text]
  • What's in a Name? the Matrix As an Introduction to Mathematics
    St. John Fisher College Fisher Digital Publications Mathematical and Computing Sciences Faculty/Staff Publications Mathematical and Computing Sciences 9-2008 What's in a Name? The Matrix as an Introduction to Mathematics Kris H. Green St. John Fisher College, [email protected] Follow this and additional works at: https://fisherpub.sjfc.edu/math_facpub Part of the Mathematics Commons How has open access to Fisher Digital Publications benefited ou?y Publication Information Green, Kris H. (2008). "What's in a Name? The Matrix as an Introduction to Mathematics." Math Horizons 16.1, 18-21. Please note that the Publication Information provides general citation information and may not be appropriate for your discipline. To receive help in creating a citation based on your discipline, please visit http://libguides.sjfc.edu/citations. This document is posted at https://fisherpub.sjfc.edu/math_facpub/12 and is brought to you for free and open access by Fisher Digital Publications at St. John Fisher College. For more information, please contact [email protected]. What's in a Name? The Matrix as an Introduction to Mathematics Abstract In lieu of an abstract, here is the article's first paragraph: In my classes on the nature of scientific thought, I have often used the movie The Matrix to illustrate the nature of evidence and how it shapes the reality we perceive (or think we perceive). As a mathematician, I usually field questions elatedr to the movie whenever the subject of linear algebra arises, since this field is the study of matrices and their properties. So it is natural to ask, why does the movie title reference a mathematical object? Disciplines Mathematics Comments Article copyright 2008 by Math Horizons.
    [Show full text]
  • Multivector Differentiation and Linear Algebra 0.5Cm 17Th Santaló
    Multivector differentiation and Linear Algebra 17th Santalo´ Summer School 2016, Santander Joan Lasenby Signal Processing Group, Engineering Department, Cambridge, UK and Trinity College Cambridge [email protected], www-sigproc.eng.cam.ac.uk/ s jl 23 August 2016 1 / 78 Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. 2 / 78 Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. 3 / 78 Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. 4 / 78 Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... 5 / 78 Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary 6 / 78 We now want to generalise this idea to enable us to find the derivative of F(X), in the A ‘direction’ – where X is a general mixed grade multivector (so F(X) is a general multivector valued function of X). Let us use ∗ to denote taking the scalar part, ie P ∗ Q ≡ hPQi. Then, provided A has same grades as X, it makes sense to define: F(X + tA) − F(X) A ∗ ¶XF(X) = lim t!0 t The Multivector Derivative Recall our definition of the directional derivative in the a direction F(x + ea) − F(x) a·r F(x) = lim e!0 e 7 / 78 Let us use ∗ to denote taking the scalar part, ie P ∗ Q ≡ hPQi.
    [Show full text]