Soft graviton theorem

Last updated
Scattering of n incoming and m outcoming particles with an outgoing graviton added to one outcoming leg. SoftGravitonTheorem.svg
Scattering of n incoming and m outcoming particles with an outgoing graviton added to one outcoming leg.

In physics, the soft graviton theorem, first formulated by Steven Weinberg in 1965, [1] allows calculation of the S-matrix, used in calculating the outcome of collisions between particles, when low-energy (soft) gravitons come into play.

Contents

Specifically, if in a collision between n incoming particles from which m outgoing particles arise, the outcome of the collision depends on a certain S matrix, by adding one or more gravitons to the n + m particles, the resulting S matrix (let it be S') differs from the initial S only by a factor that does not depend in any way, except for the momentum, on the type of particles to which the gravitons couple. [2]

The theorem also holds by putting photons in place of gravitons, thus obtaining a corresponding soft photon theorem.

The theorem is used in the context of attempts to formulate a theory of quantum gravity in the form of a perturbative quantum theory, that is, as an approximation of a possible, as yet unknown, exact theory of quantum gravity. [3]

In 2014 Andrew Strominger and Freddy Cachazo expanded the soft graviton theorem, gauge invariant under translation, to the subleading term of the series, obtaining the gauge invariance under rotation (implying global angular momentum conservation) and connected this to the gravitational spin memory effect. [4]

Formulation

Given particles whose interaction is described by a certain initial S matrix, by adding a soft graviton (i.e., whose energy is negligible compared to the energy of the other particles) that couples to one of the incoming or outgoing particles, the resulting S' matrix is, leaving off some kinematic factors,

, [1] [5]

where p is the momentum of the particle interacting with the graviton, ϵμν is the graviton polarization, pG is the momentum of the graviton, ε is an infinitesimal real quantity which helps to shape the integration contour, and the factor η is equal to 1 for outgoing particles and -1 for incoming particles.

The formula comes from a power series and the last term with the big O indicates that terms of higher order are not considered. Although the series differs depending on the spin of the particle coupling to the graviton, the lowest-order term shown above is the same for all spins. [1]

In the case of multiple soft gravitons involved, the factor in front of S is the sum of the factors due to each individual graviton.

If a soft photon (whose energy is negligible compared to the energy of the other particles) is added instead of the graviton, the resulting matrix S' is

, [1] [5]

with the same parameters as before but with pγ momentum of the photon, ϵ is its polarization, and q the charge of the particle coupled to the photon.

As for the graviton, in case of more photons, a sum over all the terms occurs.

Subleading order expansion

The expansion of the formula to the subleading term of the series for the graviton was calculated by Andrew Strominger and Freddy Cachazo: [4]

,

where represents the angular momentum of the particle interacting with the graviton.

This formula is gauge invariant under rotation and is connected to the gravitational spin memory effect. [4]

See Also

Related Research Articles

<span class="mw-page-title-main">Noether's theorem</span> Statement relating differentiable symmetries to conserved quantities

Noether's theorem or Noether's first theorem states that every differentiable symmetry of the action of a physical system with conservative forces has a corresponding conservation law. The theorem was proven by mathematician Emmy Noether in 1915 and published in 1918. The action of a physical system is the integral over time of a Lagrangian function, from which the system's behavior can be determined by the principle of least action. This theorem only applies to continuous and smooth symmetries over physical space.

<span class="mw-page-title-main">Poincaré group</span> Group of flat spacetime symmetries

The Poincaré group, named after Henri Poincaré (1906), was first defined by Hermann Minkowski (1908) as the group of Minkowski spacetime isometries. It is a ten-dimensional non-abelian Lie group that is of importance as a model in our understanding of the most basic fundamentals of physics.

<span class="mw-page-title-main">Propagator</span> Function in quantum field theory showing probability amplitudes of moving particles

In quantum mechanics and quantum field theory, the propagator is a function that specifies the probability amplitude for a particle to travel from one place to another in a given period of time, or to travel with a certain energy and momentum. In Feynman diagrams, which serve to calculate the rate of collisions in quantum field theory, virtual particles contribute their propagator to the rate of the scattering event described by the respective diagram. These may also be viewed as the inverse of the wave operator appropriate to the particle, and are, therefore, often called (causal) Green's functions.

In the study of Dirac fields in quantum field theory, Richard Feynman invented the convenient Feynman slash notation. If A is a covariant vector,

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

In differential geometry, a tensor density or relative tensor is a generalization of the tensor field concept. A tensor density transforms as a tensor field when passing from one coordinate system to another, except that it is additionally multiplied or weighted by a power W of the Jacobian determinant of the coordinate transition function or its absolute value. A tensor density with a single index is called a vector density. A distinction is made among (authentic) tensor densities, pseudotensor densities, even tensor densities and odd tensor densities. Sometimes tensor densities with a negative weight W are called tensor capacity. A tensor density can also be regarded as a section of the tensor product of a tensor bundle with a density bundle.

In the theory of general relativity, linearized gravity is the application of perturbation theory to the metric tensor that describes the geometry of spacetime. As a consequence, linearized gravity is an effective method for modeling the effects of gravity when the gravitational field is weak. The usage of linearized gravity is integral to the study of gravitational waves and weak-field gravitational lensing.

In mathematical physics, the gamma matrices, also called the Dirac matrices, are a set of conventional matrices with specific anticommutation relations that ensure they generate a matrix representation of the Clifford algebra It is also possible to define higher-dimensional gamma matrices. When interpreted as the matrices of the action of a set of orthogonal basis vectors for contravariant vectors in Minkowski space, the column vectors on which the matrices act become a space of spinors, on which the Clifford algebra of spacetime acts. This in turn makes it possible to represent infinitesimal spatial rotations and Lorentz boosts. Spinors facilitate spacetime computations in general, and in particular are fundamental to the Dirac equation for relativistic spin particles. Gamma matrices were introduced by Dirac in 1928.

In theoretical physics, massive gravity is a theory of gravity that modifies general relativity by endowing the graviton with a nonzero mass. In the classical theory, this means that gravitational waves obey a massive wave equation and hence travel at speeds below the speed of light.

In theoretical physics, a source field is a background field coupled to the original field as

<span class="mw-page-title-main">Electromagnetic stress–energy tensor</span>

In relativistic physics, the electromagnetic stress–energy tensor is the contribution to the stress–energy tensor due to the electromagnetic field. The stress–energy tensor describes the flow of energy and momentum in spacetime. The electromagnetic stress–energy tensor contains the negative of the classical Maxwell stress tensor that governs the electromagnetic interactions.

<span class="mw-page-title-main">Covariant formulation of classical electromagnetism</span> Ways of writing certain laws of physics

The covariant formulation of classical electromagnetism refers to ways of writing the laws of classical electromagnetism in a form that is manifestly invariant under Lorentz transformations, in the formalism of special relativity using rectilinear inertial coordinate systems. These expressions both make it simple to prove that the laws of classical electromagnetism take the same form in any inertial coordinate system, and also provide a way to translate the fields and forces from one frame to another. However, this is not as general as Maxwell's equations in curved spacetime or non-rectilinear coordinate systems.

In quantum field theory, and especially in quantum electrodynamics, the interacting theory leads to infinite quantities that have to be absorbed in a renormalization procedure, in order to be able to predict measurable quantities. The renormalization scheme can depend on the type of particles that are being considered. For particles that can travel asymptotically large distances, or for low energy processes, the on-shell scheme, also known as the physical scheme, is appropriate. If these conditions are not fulfilled, one can turn to other schemes, like the minimal subtraction scheme.

The table of chords, created by the Greek astronomer, geometer, and geographer Ptolemy in Egypt during the 2nd century AD, is a trigonometric table in Book I, chapter 11 of Ptolemy's Almagest, a treatise on mathematical astronomy. It is essentially equivalent to a table of values of the sine function. It was the earliest trigonometric table extensive enough for many practical purposes, including those of astronomy. Since the 8th and 9th centuries, the sine and other trigonometric functions have been used in Islamic mathematics and astronomy, reforming the production of sine tables. Khwarizmi and Habash al-Hasib later produced a set of trigonometric tables.

In quantum field theory, a Ward–Takahashi identity is an identity between correlation functions that follows from the global or gauge symmetries of the theory, and which remains valid after renormalization.

The spin angular momentum of light (SAM) is the component of angular momentum of light that is associated with the quantum spin and the rotation between the polarization degrees of freedom of the photon.

<span class="mw-page-title-main">Two-body Dirac equations</span> Quantum field theory equations

In quantum field theory, and in the significant subfields of quantum electrodynamics (QED) and quantum chromodynamics (QCD), the two-body Dirac equations (TBDE) of constraint dynamics provide a three-dimensional yet manifestly covariant reformulation of the Bethe–Salpeter equation for two spin-1/2 particles. Such a reformulation is necessary since without it, as shown by Nakanishi, the Bethe–Salpeter equation possesses negative-norm solutions arising from the presence of an essentially relativistic degree of freedom, the relative time. These "ghost" states have spoiled the naive interpretation of the Bethe–Salpeter equation as a quantum mechanical wave equation. The two-body Dirac equations of constraint dynamics rectify this flaw. The forms of these equations can not only be derived from quantum field theory they can also be derived purely in the context of Dirac's constraint dynamics and relativistic mechanics and quantum mechanics. Their structures, unlike the more familiar two-body Dirac equation of Breit, which is a single equation, are that of two simultaneous quantum relativistic wave equations. A single two-body Dirac equation similar to the Breit equation can be derived from the TBDE. Unlike the Breit equation, it is manifestly covariant and free from the types of singularities that prevent a strictly nonperturbative treatment of the Breit equation. In applications of the TBDE to QED, the two particles interact by way of four-vector potentials derived from the field theoretic electromagnetic interactions between the two particles. In applications to QCD, the two particles interact by way of four-vector potentials and Lorentz invariant scalar interactions, derived in part from the field theoretic chromomagnetic interactions between the quarks and in part by phenomenological considerations. As with the Breit equation a sixteen-component spinor Ψ is used.

<span class="mw-page-title-main">Symmetry in quantum mechanics</span> Properties underlying modern physics

Symmetries in quantum mechanics describe features of spacetime and particles which are unchanged under some transformation, in the context of quantum mechanics, relativistic quantum mechanics and quantum field theory, and with applications in the mathematical formulation of the standard model and condensed matter physics. In general, symmetry in physics, invariance, and conservation laws, are fundamentally important constraints for formulating physical theories and models. In practice, they are powerful methods for solving problems and predicting what can happen. While conservation laws do not always give the answer to the problem directly, they form the correct constraints and the first steps to solving a multitude of problems.

In mathematical physics, the Gordon decomposition of the Dirac current is a splitting of the charge or particle-number current into a part that arises from the motion of the center of mass of the particles and a part that arises from gradients of the spin density. It makes explicit use of the Dirac equation and so it applies only to "on-shell" solutions of the Dirac equation.

<span class="mw-page-title-main">Dual graviton</span> Hypothetical particle found in supergravity

In theoretical physics, the dual graviton is a hypothetical elementary particle that is a dual of the graviton under electric-magnetic duality, as an S-duality, predicted by some formulations of supergravity in eleven dimensions.

References

  1. 1 2 3 4 Weinberg, Steven (1965-10-25). "Infrared Photons and Gravitons". Physical Review. 140 (2B): B516–B524. Bibcode:1965PhRv..140..516W. doi:10.1103/PhysRev.140.B516.
  2. He, Temple; Lysov, Vyacheslav; Mitra, Prahar; Strominger, Andrew (2015-05-07). "BMS supertranslations and Weinberg's soft graviton theorem". Journal of High Energy Physics. 2015 (5): 151. arXiv: 1401.7026 . Bibcode:2015JHEP...05..151H. doi:10.1007/JHEP05(2015)151. ISSN   1029-8479. S2CID   256013139.
  3. Verma, Mritunjay. Soft Graviton Theorem in Generic Quantum Theory of Gravity (PDF). Harish-Chandra Research Institute.
  4. 1 2 3 Cachazo, Freddy; Strominger, Andrew (2014). "Evidence for a New Soft Graviton Theorem". arXiv: 1404.4091 [hep-th].
  5. 1 2 Strominger, Andrew (2018-03-06). Lectures on the Infrared Structure of Gravity and Gauge Theory. Princeton University Press. pp. 35–36. ISBN   978-0-691-17950-6.