Cluster decomposition

Last updated

In physics, the cluster decomposition property states that experiments carried out far from each other cannot influence each other. Usually applied to quantum field theory, it requires that vacuum expectation values of operators localized in bounded regions factorize whenever these regions becomes sufficiently distant from each other. First formulated by Eyvind Wichmann and James H. Crichton in 1963 in the context of the S-matrix, [1] it was conjectured by Steven Weinberg that in the low energy limit the cluster decomposition property, together with Lorentz invariance and quantum mechanics, inevitably lead to quantum field theory. String theory satisfies all three of the conditions and so provides a counter-example against this being true at all energy scales. [2]

Contents

Formulation

The S-matrix describes the amplitude for a process with an initial state evolving into a final state . If the initial and final states consist of two clusters, with and close to each other but far from the pair and , then the cluster decomposition property requires the S-matrix to factorize

as the distance between the two clusters increases. The physical interpretation of this is that any two spatially well separated experiments and cannot influence each other. [3] This condition is fundamental to the ability to doing physics without having to know the state of the entire universe. By expanding the S-matrix into a sum of a product of connected S-matrix elements , which at the perturbative level are equivalent to connected Feynman diagrams, the cluster decomposition property can be restated as demanding that connected S-matrix elements must vanish whenever some of its clusters of particles are far apart from each other.

This position space formulation can also be reformulated in terms of the momentum space S-matrix . [4] Since its Fourier transformation gives the position space connected S-matrix, this only depends on position through the exponential terms. Therefore, performing a uniform translation in a direction on a subset of particles will effectively change the momentum space S-matrix as

By translational invariance, a translation of all particles cannot change the S-matrix, therefore must be proportional to a momentum conserving delta function to ensure that the translation exponential factor vanishes. If there is an additional delta function of only a subset of momenta corresponding to some cluster of particles, then this cluster can be moved arbitrarily far through a translation without changing the S-matrix, which would violate cluster decomposition. This means that in momentum space the property requires that the S-matrix only has a single delta function.

Cluster decomposition can also be formulated in terms of correlation functions, where for any two operators and localized to some region, the vacuum expectation values factorize as the two operators become distantly separated

This formulation allows for the property to be applied to theories that lack an S-matrix such as conformal field theories. It is in terms of these Wightman functions that the property is usually formulated in axiomatic quantum field theory. [5] In some formulations, such as Euclidean constructive field theory, it is explicitly introduced as an axiom. [6]

Properties

If a theory is constructed from creation and annihilation operators, then the cluster decomposition property automatically holds. This can be seen by expanding out the S-matrix as a sum of Feynman diagrams which allows for the identification of connected S-matrix elements with connected Feynman diagrams. Vertices arise whenever creation and annihilation operators commute past each other leaving behind a single momentum delta function. In any connected diagram with V vertices, I internal lines and L loops, I-L of the delta functions go into fixing internal momenta, leaving V-(I-L) delta functions unfixed. A form of Euler's formula states that any graph with C disjoint connected components satisfies C = V-I+L. Since the connected S-matrix elements correspond to C=1 diagrams, these only have a single delta function and thus the cluster decomposition property, as formulated above in momentum space in terms of delta functions, holds.

Microcausality, the locality condition requiring commutation relations of local operators to vanish for spacelike separations, is a sufficient condition for the S-matrix to satisfy cluster decomposition. In this sense cluster decomposition serves a similar purpose for the S-matrix as microcausality does for fields, preventing causal influence from propagating between regions that are distantly separated. [7] However, cluster decomposition is weaker than having no superluminal causation since it can be formulated for classical theories as well. [8]

One key requirement for cluster decomposition is that it requires a unique vacuum state, with it failing if the vacuum state is a mixed state. [9] The rate at which the correlation functions factorize depends on the spectrum of the theory, where if it has mass gap of mass then there is an exponential falloff while if there are massless particles present then it can be as slow as . [10]

Related Research Articles

<span class="mw-page-title-main">Lorentz transformation</span> Family of linear transformations

In physics, the Lorentz transformations are a six-parameter family of linear transformations from a coordinate frame in spacetime to another frame that moves at a constant velocity relative to the former. The respective inverse transformation is then parameterized by the negative of this velocity. The transformations are named after the Dutch physicist Hendrik Lorentz.

<span class="mw-page-title-main">Pauli matrices</span> Matrices important in quantum mechanics and the study of spin

In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices which are Hermitian, involutory and unitary. Usually indicated by the Greek letter sigma, they are occasionally denoted by tau when used in connection with isospin symmetries.

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-12 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

<span class="mw-page-title-main">Stress–energy tensor</span> Tensor describing energy momentum density in spacetime

The stress–energy tensor, sometimes called the stress–energy–momentum tensor or the energy–momentum tensor, is a tensor physical quantity that describes the density and flux of energy and momentum in spacetime, generalizing the stress tensor of Newtonian physics. It is an attribute of matter, radiation, and non-gravitational force fields. This density and flux of energy and momentum are the sources of the gravitational field in the Einstein field equations of general relativity, just as mass density is the source of such a field in Newtonian gravity.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

<span class="mw-page-title-main">Second quantization</span> Formulation of the quantum many-body problem

Second quantization, also referred to as occupation number representation, is a formalism used to describe and analyze quantum many-body systems. In quantum field theory, it is known as canonical quantization, in which the fields are thought of as field operators, in a manner similar to how the physical quantities are thought of as operators in first quantization. The key ideas of this method were introduced in 1927 by Paul Dirac, and were later developed, most notably, by Pascual Jordan and Vladimir Fock. In this approach, the quantum many-body states are represented in the Fock state basis, which are constructed by filling up each single-particle state with a certain number of identical particles. The second quantization formalism introduces the creation and annihilation operators to construct and handle the Fock states, providing useful tools to the study of the quantum many-body theory.

In physics, the S-matrix or scattering matrix relates the initial state and the final state of a physical system undergoing a scattering process. It is used in quantum mechanics, scattering theory and quantum field theory (QFT).

<span class="mw-page-title-main">LSZ reduction formula</span> Connection between correlation functions and the S-matrix

In quantum field theory, the Lehmann–Symanzik–Zimmerman (LSZ) reduction formula is a method to calculate S-matrix elements from the time-ordered correlation functions of a quantum field theory. It is a step of the path that starts from the Lagrangian of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists Harry Lehmann, Kurt Symanzik and Wolfhart Zimmermann.

<span class="mw-page-title-main">Electromagnetic tensor</span> Mathematical object that describes the electromagnetic field in spacetime

In electromagnetism, the electromagnetic tensor or electromagnetic field tensor is a mathematical object that describes the electromagnetic field in spacetime. The field tensor was first used after the four-dimensional tensor formulation of special relativity was introduced by Hermann Minkowski. The tensor allows related physical laws to be written very concisely, and allows for the quantization of the electromagnetic field by Lagrangian formulation described below.

In general relativity, the Gibbons–Hawking–York boundary term is a term that needs to be added to the Einstein–Hilbert action when the underlying spacetime manifold has a boundary.

The normal-inverse Gaussian distribution is a continuous probability distribution that is defined as the normal variance-mean mixture where the mixing density is the inverse Gaussian distribution. The NIG distribution was noted by Blaesild in 1977 as a subclass of the generalised hyperbolic distribution discovered by Ole Barndorff-Nielsen. In the next year Barndorff-Nielsen published the NIG in another paper. It was introduced in the mathematical finance literature in 1997.

The Wigner D-matrix is a unitary matrix in an irreducible representation of the groups SU(2) and SO(3). It was introduced in 1927 by Eugene Wigner, and plays a fundamental role in the quantum mechanical theory of angular momentum. The complex conjugate of the D-matrix is an eigenfunction of the Hamiltonian of spherical and symmetric rigid rotors. The letter D stands for Darstellung, which means "representation" in German.

In many-body theory, the term Green's function is sometimes used interchangeably with correlation function, but refers specifically to correlators of field operators or creation and annihilation operators.

Non-linear least squares is the form of least squares analysis used to fit a set of m observations with a model that is non-linear in n unknown parameters (m ≥ n). It is used in some forms of nonlinear regression. The basis of the method is to approximate the model by a linear one and to refine the parameters by successive iterations. There are many similarities to linear least squares, but also some significant differences. In economic theory, the non-linear least squares method is applied in (i) the probit regression, (ii) threshold regression, (iii) smooth regression, (iv) logistic link regression, (v) Box–Cox transformed regressors ().

In physics, relativistic quantum mechanics (RQM) is any Poincaré covariant formulation of quantum mechanics (QM). This theory is applicable to massive particles propagating at all velocities up to those comparable to the speed of light c, and can accommodate massless particles. The theory has application in high energy physics, particle physics and accelerator physics, as well as atomic physics, chemistry and condensed matter physics. Non-relativistic quantum mechanics refers to the mathematical formulation of quantum mechanics applied in the context of Galilean relativity, more specifically quantizing the equations of classical mechanics by replacing dynamical variables by operators. Relativistic quantum mechanics (RQM) is quantum mechanics applied with special relativity. Although the earlier formulations, like the Schrödinger picture and Heisenberg picture were originally formulated in a non-relativistic background, a few of them also work with special relativity.

The purpose of this page is to provide supplementary materials for the ordinary least squares article, reducing the load of the main article with mathematics and improving its accessibility, while at the same time retaining the completeness of exposition.

In statistics, the matrix t-distribution is the generalization of the multivariate t-distribution from vectors to matrices. The matrix t-distribution shares the same relationship with the multivariate t-distribution that the matrix normal distribution shares with the multivariate normal distribution. For example, the matrix t-distribution is the compound distribution that results from sampling from a matrix normal distribution having sampled the covariance matrix of the matrix normal from an inverse Wishart distribution.

<span class="mw-page-title-main">Bargmann–Wigner equations</span> Wave equation for arbitrary spin particles

In relativistic quantum mechanics and quantum field theory, the Bargmann–Wigner equations describe free particles with non-zero mass and arbitrary spin j, an integer for bosons or half-integer for fermions. The solutions to the equations are wavefunctions, mathematically in the form of multi-component spinor fields.

<span class="mw-page-title-main">Gluon field strength tensor</span> Second rank tensor in quantum chromodynamics

In theoretical particle physics, the gluon field strength tensor is a second order tensor field characterizing the gluon interaction between quarks.

In the ADM formulation of general relativity one splits spacetime into spatial slices and time, the basic variables are taken to be the induced metric, , on the spatial slice, and its conjugate momentum variable related to the extrinsic curvature, ,. These are the metric canonical coordinates.

References

  1. Wichmann, E.H.; Crichton, J.H. (1963). "Cluster Decomposition Properties of the S Matrix". Phys. Rev. American Physical Society. 132 (6): 2788–2799. Bibcode:1963PhRv..132.2788W. doi:10.1103/PhysRev.132.2788.
  2. Weinberg, S. (1996). What is quantum field theory, and what did we think it is?. Conference on Historical Examination and Philosophical Reflections on the Foundations of Quantum Field Theory. pp. 241–251. arXiv: hep-th/9702027 .
  3. Schwartz, M. D. (2014). "7". Quantum Field Theory and the Standard Model. Cambridge University Press. pp. 96–97. ISBN   9781107034730.
  4. Weinberg, S. (1995). "4". The Quantum Theory of Fields: Foundations. Vol. 1. Cambridge University Press. pp. 177–188. ISBN   9780521670531.
  5. Bogolubov, N.N.; Logunov, A.A.; Todorov, I.T. (1975). Introduction to Axiomatic Quantum Field Theory. Translated by Fulling, S.A.; Popova, L.G. (1 ed.). Benjamin. pp. 272–282. ISBN   9780805309829.
  6. Iagolnitzer, D. (1993). "3". Scattering in Quantum Field Theories The Axiomatic and Constructive Approaches. Princeton University Press. pp. 155–156. ISBN   9780691633282.
  7. Brown, L.S. (1992). "6". Quantum Field Theory. Cambridge: Cambridge University Press. pp. 311–313. doi:10.1017/CBO9780511622649. ISBN   978-0521469463.
  8. Bain, J. (1998). "Weinberg on Qft: Demonstrative Induction and Underdetermination". Synthese. 117 (1): 7–8. doi:10.1023/A:1005025424031. JSTOR   20118095. S2CID   9049200.
  9. Weinberg, S. (1995). "19". The Quantum Theory of Fields: Modern Applications. Vol. 2. Cambridge University Press. p. 167. ISBN   9780521670548.
  10. Streater, R.F.; Wightman, A.S. (2000) [1964]. "3". PCT, Spin and Statistics, and All That. Princeton: Princeton University Press. p. 113. ISBN   978-0691070629.