The Quantum EM Fields and the Photon Propagator

Total Page:16

File Type:pdf, Size:1020Kb

The Quantum EM Fields and the Photon Propagator The Quantum EM Fields and the Photon Propagator Quantizing the free electromagnetic tension fields E and B is fairly straightforward. The time-independent Maxwell equations Bˆ (x) = 0, Eˆ(x) = Jˆ0(x) 0 (forthefreefields) (1) ∇· ∇· → are imposed as operatorial constraints in the Hilbert space, while the time-dependent Maxwell equations ∂ ∂ Bˆ = Eˆ, Eˆ = + Bˆ (2) ∂t −∇ × ∂t ∇ × follow from the free Hamiltonian ˆ 3 1 ˆ 2 1 ˆ 2 H = d x 2 E (x) + 2 B (x) (3) Z and the equal-time commutation relations Eˆi(x, t), Eˆj(y, t) = 0, i j Bˆ (x, t), Bˆ (y, t) = 0, (4) ∂ Eˆi(x, t), Bˆj(y, t) = iǫijk δ(3)(x y). − ∂xk − In light of eqs. (1), we may expand the vector fields into momentum/polarization modes using only the transverse polarizations, for example the helicity modes λ = 1 only, thus ± k ek,λ = iλ k ek,λ, ek,λ k for λ = 1, × | | ⊥ ± 3 d k ixk † Eˆ(x) = e ek Eˆk , Eˆ = Eˆ , (2π)3 ,λ × ,λ k,λ − −k,λ Z λ=±X1 only 3 d k ixk † Bˆ (x) = e ek Bˆk , Bˆ = Bˆ , (2π)3 ,λ × ,λ k,λ − −k,λ Z λ=±X1 only 3 (3) ′ Eˆ k , Eˆ k′ ′ = 0, Bˆ k , Bˆ k′ ′ = 0, Eˆ k , Bˆ k′ ′ = iλ k δ ′ (2π) δ (k + k ). ,λ ,λ ,λ ,λ ,λ ,λ | | λλ × (5) Consequently, the photonic creation and annihilation operators are defined for the transverse polarizations only, aˆ = λBˆ + iEˆ , aˆ† = λBˆ† iEˆ† , λ = 1 only, (6) k,λ k,λ k,λ k,λ k,λ − k,λ ± 1 † † † 3 (3) ′ aˆ , aˆ ′ ′ = 0, aˆ , aˆ ′ ′ = 0, aˆ , aˆ ′ ′ = δλλ′ 2 k (2π) δ (k k ). (7) k,λ k ,λ k,λ k ,λ k,λ k ,λ × | | − In terms of these operators, the free EM Hamiltonian becomes d3k 1 ˆ † k H 3 ωkaˆk,λaˆk,λ + const for ωk = + , (8) (2π) 2ωk | | Z λ=±X1 only which makes for the usual time-dependence in the Heisenberg picture: aˆ aˆ e−iωkt, aˆ† aˆ† e+iωkt, (9) k,λ → k,λ × k,λ → k,λ × and hence (after a bit of algebra) d3k 1 Eˆ(x, t) = e−iωt+ikx( iωe )ˆa + e+iωt−ikx(+iωe∗ )ˆa† , (2π)3 2ω − k,λ k,λ k,λ k,λ Z λ=±X1 only d3k 1 Bˆ (x, t) = e−iωt+ikx( ik e )ˆa + e+iωt−ikx(+ik e∗ )ˆa† . (2π)3 2ω − × k,λ k,λ × k,λ k,λ Z λ=±X1 only (10) Up to now I have focused on the EM tension fields Fˆµν(x), but to couple the EM to come charged fields — like the electron field Ψ(ˆ x) — I would also need the quantum potential µ fields Aˆµ(x) to spell out the interactions Jµ A . Classically, the potential fields are L ⊃ − × subject to gauge symmetries Aµ (x) Aµ (x) = Aµ (x) ∂µΛ(x), (11) old → new old − but implementing such symmetries in the quantum theory is highly non-trivial since a trans- form changing the time-dependence of quantum fields must change the Hamiltonian oper- ator. There are more problems with quantum gauge theories, and I’ll deal with them next semester. For the moment, let me simply say that the canonical quantization of the potential fields Aµ(x) requires fixing a gauge. That is, we should remove the gauge-redundancy of the potential fields by imposing a local linear constraint at each x, for example A(x) 0 ∇· ≡ µ 3 (the Coulomb gauge), or ∂µA (x) 0 (the Landau gauge), or A (x) 0 (the axial gauge). ≡ ≡ 2 Once we impose such a constraint at the Lagrangian level, we find the canonical conjugates of the remaining independent fields, build the classical Hamiltonian, and then quantize the theory in the usual way to build the quantum fields, the Hilbert space where they act, and the Hamiltonian operator. In general, different gauge-fixing constraints give rise to different quantum theories, each having its own Hilbert space and the Hamiltonian operator. However, all such theories are physically equivalent to each other! For the free electromagnetic fields, I do not have to re-quantize the theory starting from the Lagrangian and some gauge-fixing constraint for the Aµ(x). Since I already have the quantum Eˆ(x) and Bˆ (x) fields as in eqs. (10), I can can obtain the free potential fields Aˆµ(x) by simply solving the equations ∂µAˆν(x) ∂ν Aˆµ(x)= Fˆµν(x) combined with the gauge-fixing − constraint. By linearity and translation invariance, the result has form d3k 1 Aˆµ (x) = e−ikx µ(k,λ) aˆ + e+ikx µ∗(k,λ) aˆ† (12) free (2π)3 2ω E × k,λ E × k,λ Z λ=±X1 only where the plane waves Aµ(x) = e−ikx µ(k,λ) and Aµ(x) = e+ikx µ∗(k,λ) for k0 =+ k (13) ×E ×E | | obey the gauge-fixing conditions as well as the Maxwell equations. For example, in the Coulomb gauge Aˆ = 0, the polarization vectors µ have simple form ∇· E µ(k,λ) = 0, e(k,λ) (14) E while in the axial gauge n A = 0 (where n is a fixed unit 3-vector) we have much messier · formulae n e(k,λ) 0 n e(k,λ) ~(k,λ) = e(k,λ) · k, (k,λ) = · ωk . (15) E − n k E − n k · · 3 Photon Propagator The photon propagator Gµν(x y) = 0 TAˆµ(x)Aˆν(y) 0 (16) F − h | | i depends on the gauge-fixing condition for the quantum potential fields Aˆµ(x). So let me first calculate it for the Coulomb gauge Aˆ 0, and then I’ll deal with the other gauges. ∇· ≡ Instead of calculating the propagator directly from eqs. (16), (12), and (14), let me use the fact that Gµν(x y) is a Green’s function of the Maxwell equations for the Aµ fields, − 2 ν ν µ ν µ 4 µν ∂ A (x) ∂ (∂µA (x)) = J (x) = A (x) = d y ( i)G (x y) Jν(y). (17) − ⇒ − − × Z Let’s start with the scalar potential A0(x, t). In the Coulomb gauge we have µ 0 0 ∂µA = ∂ A + A = ∂ A + 0, (18) 0 ∇· 0 which makes the equation for the A0(x) time-independent. Indeed, 0 2 0 0 µ 2 2 0 0 0 2 0 J = ∂ A (x) ∂ (∂µA (x)) = (∂ )A ∂ (∂ A ) = A (x) (19) − 0 −∇ − 0 −∇ 0 while the terms containing the time derivatives ∂0 cancel out. Consequently, the A (x, t) is the instantaneous Coulomb potential for the electric charge density J0(y, t), J0(y, same t) A0(x, t) = d3y (20) 4π x y Z | − | — which is why this gauge is called the Coulomb gauge. In terms of the Green’s function, this means iδ(x0 y0) G0i 0, G00(x y) = − , (21) ≡ − 4π x y | − | of after Fourier transform to the momentum space, i G00(k) = independent of k0. (22) k2 4 As to the vector potential A, it obeys 2 µ ∂ A + (∂µA ) = J (23) ∇ where in the Coulomb gauge µ 0 1 0 1 ∂µA = ∂ A = ∂ J = + ( J) (24) 0 − 2 0 2 ∇· ∇ ∇ where the last equality follows from the electric current conservation, ∂ J0 = J. Con- 0 −∇ · sequently, 1 (∂2 2)A = J ( J), (25) 0 −∇ − ∇ 2 ∇· ∇ or in momentum basis kikj (k2 k2) Ai(k) = Ji Jj. (26) − 0 − × − k2 × In terms of the Green’s function, this means i kikj Gi0 = 0, Gij(k) = δij . (27) k2 k2 × − k2 0 − Altogether, all the components of the Gµν(k) can be summarized as i Gµν(k) = Cµν(k) (28) k2 k2 × 0 − where kikj k2 k2 k2 Cij = δij , Ci0 = C0i = 0, C00 = 0 − = 0 1. (29) − k2 k2 k2 − A convenient way to summarize this Cµν(k) tensor is (k0, k) Cµν(k) = gµν + kµcν(k) + kνcµ(k) for cµ(k) = − . (30) − 2k2 In the coordinate space d4k ie−ik(x−y) Gµν(x y) = Cµν(k) (31) − (2π)4 k2 × Z where the denominator has poles at k0 = k . As usual, these poles should be regularized ±| | by moving them off the real axis, and different regularizations lead to different Green’s 5 functions. We are interested in the specific Green’s function — the Feynman propagator — so we should move the poles to k0 =+ k i0 and k0 = k + i0, thus | | − −| | d4k ie−ik(x−y) Gµν(x y) = Cµν(k) = gµν + kµcν(k) + kνcµ(k) . (32) F − (2π)4 k2 + i0 × − Z Now consider photon propagators for different gauge conditions for the EM potential fields Aˆµ(x). When we change a gauge condition, the potential fields change as Aˆµ (x) = Aˆµ (x) ∂µΛ(ˆ x) (33) new old − for some Λ(ˆ x) that depends on the old Aˆµ(x) and the new gauge condition. For linear gauge conditions, 4 µ Λ(ˆ x) = d y Lµ(x y)Aˆ (y) (34) − old Z for some kernel Lµ(x y), or in momentum space − ˆ ˆµ Λ(k) = Lµ(k)Aold(k). (35) Consequently, in the new gauge the Feynman propagator becomes Gµν (x y) = 0 TAˆµ (x)Aˆν (y) 0 new − h | new new | i ˆµ ˆν ∂ ˆ ˆν = 0 TAold(x)Aold(y) 0 0 TΛ(x)Aold(y) 0 h | | i − ∂xµ h | | i 2 ∂ ˆµ ˆ ∂ ˆ ˆ 0 TAold(x)Λ(y) 0 + 0 TΛ(x)Λ(y) 0 − ∂yν h | | i ∂xµ∂yν h | | i 2 µν ∂ ν ∂ µ ∂ = Gold(x y) + f (x y) + f (y x) + h(x y) − ∂xµ − ∂yν − ∂xµ∂yν − for some functions f µ(x y) and h(x y).
Recommended publications
  • 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]
  • Quantum Gravity and the Cosmological Constant Problem
    Quantum Gravity and the Cosmological Constant Problem J. W. Moffat Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada and Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada October 15, 2018 Abstract A finite and unitary nonlocal formulation of quantum gravity is applied to the cosmological constant problem. The entire functions in momentum space at the graviton-standard model particle loop vertices generate an exponential suppression of the vacuum density and the cosmological constant to produce agreement with their observational bounds. 1 Introduction A nonlocal quantum field theory and quantum gravity theory has been formulated that leads to a finite, unitary and locally gauge invariant theory [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14]. For quantum gravity the finiteness of quantum loops avoids the problem of the non-renormalizabilty of local quantum gravity [15, 16]. The finiteness of the nonlocal quantum field theory draws from the fact that factors of exp[ (p2)/Λ2] are attached to propagators which suppress any ultraviolet divergences in Euclidean momentum space,K where Λ is an energy scale factor. An important feature of the field theory is that only the quantum loop graphs have nonlocal properties; the classical tree graph theory retains full causal and local behavior. Consider first the 4-dimensional spacetime to be approximately flat Minkowski spacetime. Let us denote by f a generic local field and write the standard local Lagrangian as [f]= [f]+ [f], (1) L LF LI where and denote the free part and the interaction part of the action, respectively, and LF LI arXiv:1407.2086v1 [gr-qc] 7 Jul 2014 1 [f]= f f .
    [Show full text]
  • Gauge-Fixing Degeneracies and Confinement in Non-Abelian Gauge Theories
    PH YSICAL RKVIE% 0 VOLUME 17, NUMBER 4 15 FEBRUARY 1978 Gauge-fixing degeneracies and confinement in non-Abelian gauge theories Carl M. Bender Department of I'hysics, 8'ashington University, St. Louis, Missouri 63130 Tohru Eguchi Stanford Linear Accelerator Center, Stanford, California 94305 Heinz Pagels The Rockefeller University, New York, N. Y. 10021 (Received 28 September 1977) Following several suggestions of Gribov we have examined the problem of gauge-fixing degeneracies in non- Abelian gauge theories. First we modify the usual Faddeev-Popov prescription to take gauge-fixing degeneracies into account. We obtain a formal expression for the generating functional which is invariant under finite gauge transformations and which counts gauge-equivalent orbits only once. Next we examine the instantaneous Coulomb interaction in the canonical formalism with the Coulomb-gauge condition. We find that the spectrum of the Coulomb Green's function in an external monopole-hke field configuration has an accumulation of negative-energy bound states at E = 0. Using semiclassical methods we show that this accumulation phenomenon, which is closely linked with gauge-fixing degeneracies, modifies the usual Coulomb propagator from (k) ' to Iki ' for small )k[. This confinement behavior depends only on the long-range behavior of the field configuration. We thereby demonstrate the conjectured confinement property of non-Abelian gauge theories in the Coulomb gauge. I. INTRODUCTION lution to the theory. The failure of the Borel sum to define the theory and gauge-fixing degeneracy It has recently been observed by Gribov' that in are correlated phenomena. non-Abelian gauge theories, in contrast with Abe- In this article we examine the gauge-fixing de- lian theories, standard gauge-fixing conditions of generacy further.
    [Show full text]
  • 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.
    [Show full text]
  • Gauge Symmetry Breaking in Gauge Theories---In Search of Clarification
    Gauge Symmetry Breaking in Gauge Theories—In Search of Clarification Simon Friederich [email protected] Universit¨at Wuppertal, Fachbereich C – Mathematik und Naturwissenschaften, Gaußstr. 20, D-42119 Wuppertal, Germany Abstract: The paper investigates the spontaneous breaking of gauge sym- metries in gauge theories from a philosophical angle, taking into account the fact that the notion of a spontaneously broken local gauge symmetry, though widely employed in textbook expositions of the Higgs mechanism, is not supported by our leading theoretical frameworks of gauge quantum the- ories. In the context of lattice gauge theory, the statement that local gauge symmetry cannot be spontaneously broken can even be made rigorous in the form of Elitzur’s theorem. Nevertheless, gauge symmetry breaking does occur in gauge quantum field theories in the form of the breaking of remnant subgroups of the original local gauge group under which the theories typi- cally remain invariant after gauge fixing. The paper discusses the relation between these instances of symmetry breaking and phase transitions and draws some more general conclusions for the philosophical interpretation of gauge symmetries and their breaking. 1 Introduction The interpretation of symmetries and symmetry breaking has been recog- nized as a central topic in the philosophy of science in recent years. Gauge symmetries, in particular, have attracted a considerable amount of inter- arXiv:1107.4664v2 [physics.hist-ph] 5 Aug 2012 est due to the central role they play in our most successful theories of the fundamental constituents of nature. The standard model of elementary par- ticle physics, for instance, is formulated in terms of gauge symmetry, and so are its most discussed extensions.
    [Show full text]
  • Path Integrals in Quantum Mechanics
    Path Integrals in Quantum Mechanics Emma Wikberg Project work, 4p Department of Physics Stockholm University 23rd March 2006 Abstract The method of Path Integrals (PI’s) was developed by Richard Feynman in the 1940’s. It offers an alternate way to look at quantum mechanics (QM), which is equivalent to the Schrödinger formulation. As will be seen in this project work, many "elementary" problems are much more difficult to solve using path integrals than ordinary quantum mechanics. The benefits of path integrals tend to appear more clearly while using quantum field theory (QFT) and perturbation theory. However, one big advantage of Feynman’s formulation is a more intuitive way to interpret the basic equations than in ordinary quantum mechanics. Here we give a basic introduction to the path integral formulation, start- ing from the well known quantum mechanics as formulated by Schrödinger. We show that the two formulations are equivalent and discuss the quantum mechanical interpretations of the theory, as well as the classical limit. We also perform some explicit calculations by solving the free particle and the harmonic oscillator problems using path integrals. The energy eigenvalues of the harmonic oscillator is found by exploiting the connection between path integrals, statistical mechanics and imaginary time. Contents 1 Introduction and Outline 2 1.1 Introduction . 2 1.2 Outline . 2 2 Path Integrals from ordinary Quantum Mechanics 4 2.1 The Schrödinger equation and time evolution . 4 2.2 The propagator . 6 3 Equivalence to the Schrödinger Equation 8 3.1 From the Schrödinger equation to PI’s . 8 3.2 From PI’s to the Schrödinger equation .
    [Show full text]
  • Operator Gauge Symmetry in QED
    Symmetry, Integrability and Geometry: Methods and Applications Vol. 2 (2006), Paper 013, 7 pages Operator Gauge Symmetry in QED Siamak KHADEMI † and Sadollah NASIRI †‡ † Department of Physics, Zanjan University, P.O. Box 313, Zanjan, Iran E-mail: [email protected] URL: http://www.znu.ac.ir/Members/khademi.htm ‡ Institute for Advanced Studies in Basic Sciences, IASBS, Zanjan, Iran E-mail: [email protected] Received October 09, 2005, in final form January 17, 2006; Published online January 30, 2006 Original article is available at http://www.emis.de/journals/SIGMA/2006/Paper013/ Abstract. In this paper, operator gauge transformation, first introduced by Kobe, is applied to Maxwell’s equations and continuity equation in QED. The gauge invariance is satisfied after quantization of electromagnetic fields. Inherent nonlinearity in Maxwell’s equations is obtained as a direct result due to the nonlinearity of the operator gauge trans- formations. The operator gauge invariant Maxwell’s equations and corresponding charge conservation are obtained by defining the generalized derivatives of the f irst and second kinds. Conservation laws for the real and virtual charges are obtained too. The additional terms in the field strength tensor are interpreted as electric and magnetic polarization of the vacuum. Key words: gauge transformation; Maxwell’s equations; electromagnetic fields 2000 Mathematics Subject Classification: 81V80; 78A25 1 Introduction Conserved quantities of a system are direct consequences of symmetries inherent in the system. Therefore, symmetries are fundamental properties of any dynamical system [1, 2, 3]. For a sys- tem of electric charges an important quantity that is always expected to be conserved is the total charge.
    [Show full text]
  • 'Diagramology' Types of Feynman Diagram
    `Diagramology' Types of Feynman Diagram Tim Evans (2nd January 2018) 1. Pieces of Diagrams Feynman diagrams1 have four types of element:- • Internal Vertices represented by a dot with some legs coming out. Each type of vertex represents one term in Hint. • Internal Edges (or legs) represented by lines between two internal vertices. Each line represents a non-zero contraction of two fields, a Feynman propagator. • External Vertices. An external vertex represents a coordinate/momentum variable which is not integrated out. Whatever the diagram represents (matrix element M, Green function G or sometimes some other mathematical object) the diagram is a function of the values linked to external vertices. Sometimes external vertices are represented by a dot or a cross. However often no explicit notation is used for an external vertex so an edge ending in space and not at a vertex with one or more other legs will normally represent an external vertex. • External Legs are ones which have at least one end at an external vertex. Note that external vertices may not have an explicit symbol so these are then legs which just end in the middle of no where. Still represents a Feynman propagator. 2. Subdiagrams A subdiagram is a subset of vertices V and all the edges between them. You may also want to include the edges connected at one end to a vertex in our chosen subset, V, but at the other end connected to another vertex or an external leg. Sometimes these edges connecting the subdiagram to the rest of the diagram are not included, in which case we have \amputated the legs" and we are working with a \truncated diagram".
    [Show full text]
  • Avoiding Gauge Ambiguities in Cavity Quantum Electrodynamics Dominic M
    www.nature.com/scientificreports OPEN Avoiding gauge ambiguities in cavity quantum electrodynamics Dominic M. Rouse1*, Brendon W. Lovett1, Erik M. Gauger2 & Niclas Westerberg2,3* Systems of interacting charges and felds are ubiquitous in physics. Recently, it has been shown that Hamiltonians derived using diferent gauges can yield diferent physical results when matter degrees of freedom are truncated to a few low-lying energy eigenstates. This efect is particularly prominent in the ultra-strong coupling regime. Such ambiguities arise because transformations reshufe the partition between light and matter degrees of freedom and so level truncation is a gauge dependent approximation. To avoid this gauge ambiguity, we redefne the electromagnetic felds in terms of potentials for which the resulting canonical momenta and Hamiltonian are explicitly unchanged by the gauge choice of this theory. Instead the light/matter partition is assigned by the intuitive choice of separating an electric feld between displacement and polarisation contributions. This approach is an attractive choice in typical cavity quantum electrodynamics situations. Te gauge invariance of quantum electrodynamics (QED) is fundamental to the theory and can be used to greatly simplify calculations1–8. Of course, gauge invariance implies that physical observables are the same in all gauges despite superfcial diferences in the mathematics. However, it has recently been shown that the invariance is lost in the strong light/matter coupling regime if the matter degrees of freedom are treated as quantum systems with a fxed number of energy levels8–14, including the commonly used two-level truncation (2LT). At the origin of this is the role of gauge transformations (GTs) in deciding the partition between the light and matter degrees of freedom, even if the primary role of gauge freedom is to enforce Gauss’s law.
    [Show full text]
  • 7 Quantized Free Dirac Fields
    7 Quantized Free Dirac Fields 7.1 The Dirac Equation and Quantum Field Theory The Dirac equation is a relativistic wave equation which describes the quantum dynamics of spinors. We will see in this section that a consistent description of this theory cannot be done outside the framework of (local) relativistic Quantum Field Theory. The Dirac Equation (i∂/ m)ψ =0 ψ¯(i∂/ + m) = 0 (1) − can be regarded as the equations of motion of a complex field ψ. Much as in the case of the scalar field, and also in close analogy to the theory of non-relativistic many particle systems discussed in the last chapter, the Dirac field is an operator which acts on a Fock space. We have already discussed that the Dirac equation also follows from a least-action-principle. Indeed the Lagrangian i µ = [ψ¯∂ψ/ (∂µψ¯)γ ψ] mψψ¯ ψ¯(i∂/ m)ψ (2) L 2 − − ≡ − has the Dirac equation for its equation of motion. Also, the momentum Πα(x) canonically conjugate to ψα(x) is ψ δ † Πα(x)= L = iψα (3) δ∂0ψα(x) Thus, they obey the equal-time Poisson Brackets ψ 3 ψα(~x), Π (~y) P B = iδαβδ (~x ~y) (4) { β } − Thus † 3 ψα(~x), ψ (~y) P B = δαβδ (~x ~y) (5) { β } − † In other words the field ψα and its adjoint ψα are a canonical pair. This result follows from the fact that the Dirac Lagrangian is first order in time derivatives. Notice that the field theory of non-relativistic many-particle systems (for both fermions on bosons) also has a Lagrangian which is first order in time derivatives.
    [Show full text]
  • A Supersymmetry Primer
    hep-ph/9709356 version 7, January 2016 A Supersymmetry Primer Stephen P. Martin Department of Physics, Northern Illinois University, DeKalb IL 60115 I provide a pedagogical introduction to supersymmetry. The level of discussion is aimed at readers who are familiar with the Standard Model and quantum field theory, but who have had little or no prior exposure to supersymmetry. Topics covered include: motiva- tions for supersymmetry, the construction of supersymmetric Lagrangians, superspace and superfields, soft supersymmetry-breaking interactions, the Minimal Supersymmetric Standard Model (MSSM), R-parity and its consequences, the origins of supersymmetry breaking, the mass spectrum of the MSSM, decays of supersymmetric particles, experi- mental signals for supersymmetry, and some extensions of the minimal framework. Contents 1 Introduction 3 2 Interlude: Notations and Conventions 13 3 Supersymmetric Lagrangians 17 3.1 The simplest supersymmetric model: a free chiral supermultiplet ............. 18 3.2 Interactions of chiral supermultiplets . ................ 22 3.3 Lagrangians for gauge supermultiplets . .............. 25 3.4 Supersymmetric gauge interactions . ............. 26 3.5 Summary: How to build a supersymmetricmodel . ............ 28 4 Superspace and superfields 30 4.1 Supercoordinates, general superfields, and superspace differentiation and integration . 31 4.2 Supersymmetry transformations the superspace way . ................ 33 4.3 Chiral covariant derivatives . ............ 35 4.4 Chiralsuperfields................................ ........ 37 arXiv:hep-ph/9709356v7 27 Jan 2016 4.5 Vectorsuperfields................................ ........ 38 4.6 How to make a Lagrangian in superspace . .......... 40 4.7 Superspace Lagrangians for chiral supermultiplets . ................... 41 4.8 Superspace Lagrangians for Abelian gauge theory . ................ 43 4.9 Superspace Lagrangians for general gauge theories . ................. 46 4.10 Non-renormalizable supersymmetric Lagrangians . .................. 49 4.11 R symmetries........................................
    [Show full text]
  • 8 Non-Abelian Gauge Theory: Perturbative Quantization
    8 Non-Abelian Gauge Theory: Perturbative Quantization We’re now ready to consider the quantum theory of Yang–Mills. In the first few sec- tions, we’ll treat the path integral formally as an integral over infinite dimensional spaces, without worrying about imposing a regularization. We’ll turn to questions about using renormalization to make sense of these formal integrals in section 8.3. To specify Yang–Mills theory, we had to pick a principal G bundle P M together ! with a connection on P . So our first thought might be to try to define the Yang–Mills r partition function as ? SYM[ ]/~ Z [(M,g), g ] = A e− r (8.1) YM YM D ZA where 1 SYM[ ]= tr(F F ) (8.2) r −2g2 r ^⇤ r YM ZM as before, and is the space of all connections on P . A To understand what this integral might mean, first note that, given any two connections and , the 1-parameter family r r0 ⌧ = ⌧ +(1 ⌧) 0 (8.3) r r − r is also a connection for all ⌧ [0, 1]. For example, you can check that the rhs has the 2 behaviour expected of a connection under any gauge transformation. Thus we can find a 1 path in between any two connections. Since 0 ⌦M (g), we conclude that is an A r r2 1 A infinite dimensional affine space whose tangent space at any point is ⌦M (g), the infinite dimensional space of all g–valued covectors on M. In fact, it’s easy to write down a flat (L2-)metric on using the metric on M: A 2 1 µ⌫ a a d ds = tr(δA δA)= g δAµ δA⌫ pg d x.
    [Show full text]