S-matrix

Last updated

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).

Contents

More formally, in the context of QFT, the S-matrix is defined as the unitary matrix connecting sets of asymptotically free particle states (the in-states and the out-states) in the Hilbert space of physical states. A multi-particle state is said to be free (or non-interacting) if it transforms under Lorentz transformations as a tensor product, or direct product in physics parlance, of one-particle states as prescribed by equation (1) below. Asymptotically free then means that the state has this appearance in either the distant past or the distant future.

While the S-matrix may be defined for any background (spacetime) that is asymptotically solvable and has no event horizons, it has a simple form in the case of the Minkowski space. In this special case, the Hilbert space is a space of irreducible unitary representations of the inhomogeneous Lorentz group (the Poincaré group); the S-matrix is the evolution operator between (the distant past), and (the distant future). It is defined only in the limit of zero energy density (or infinite particle separation distance).

It can be shown that if a quantum field theory in Minkowski space has a mass gap, the state in the asymptotic past and in the asymptotic future are both described by Fock spaces.

History

The initial elements of S-matrix theory are found in Paul Dirac's 1927 paper "Über die Quantenmechanik der Stoßvorgänge". [1] [2] The S-matrix was first properly introduced by John Archibald Wheeler in the 1937 paper "On the Mathematical Description of Light Nuclei by the Method of Resonating Group Structure". [3] In this paper Wheeler introduced a scattering matrix – a unitary matrix of coefficients connecting "the asymptotic behaviour of an arbitrary particular solution [of the integral equations] with that of solutions of a standard form", [4] but did not develop it fully.

In the 1940s, Werner Heisenberg independently developed and substantiated the idea of the S-matrix. Because of the problematic divergences present in quantum field theory at that time, Heisenberg was motivated to isolate the essential features of the theory that would not be affected by future changes as the theory developed. In doing so, he was led to introduce a unitary "characteristic" S-matrix. [4]

Today, however, exact S-matrix results are important for conformal field theory, integrable systems, and several further areas of quantum field theory and string theory. S-matrices are not substitutes for a field-theoretic treatment, but rather, complement the end results of such.

Motivation

In high-energy particle physics one is interested in computing the probability for different outcomes in scattering experiments. These experiments can be broken down into three stages:

  1. Making a collection of incoming particles collide (usually two kinds of particles with high energies).
  2. Allowing the incoming particles to interact. These interactions may change the types of particles present (e.g. if an electron and a positron annihilate they may produce two photons).
  3. Measuring the resulting outgoing particles.

The process by which the incoming particles are transformed (through their interaction) into the outgoing particles is called scattering. For particle physics, a physical theory of these processes must be able to compute the probability for different outgoing particles when different incoming particles collide with different energies.

The S-matrix in quantum field theory achieves exactly this. It is assumed that the small-energy-density approximation is valid in these cases.

Use

The S-matrix is closely related to the transition probability amplitude in quantum mechanics and to cross sections of various interactions; the elements (individual numerical entries) in the S-matrix are known as scattering amplitudes. Poles of the S-matrix in the complex-energy plane are identified with bound states, virtual states or resonances. Branch cuts of the S-matrix in the complex-energy plane are associated to the opening of a scattering channel.

In the Hamiltonian approach to quantum field theory, the S-matrix may be calculated as a time-ordered exponential of the integrated Hamiltonian in the interaction picture; it may also be expressed using Feynman's path integrals. In both cases, the perturbative calculation of the S-matrix leads to Feynman diagrams.

In scattering theory, the S-matrix is an operator mapping free particle in-states to free particle out-states (scattering channels) in the Heisenberg picture. This is very useful because often we cannot describe the interaction (at least, not the most interesting ones) exactly.

In one-dimensional quantum mechanics

A simple prototype in which the S-matrix is 2-dimensional is considered first, for the purposes of illustration. In it, particles with sharp energy E scatter from a localized potential V according to the rules of 1-dimensional quantum mechanics. Already this simple model displays some features of more general cases, but is easier to handle.

Each energy E yields a matrix S = S(E) that depends on V. Thus, the total S-matrix could, figuratively speaking, be visualized, in a suitable basis, as a "continuous matrix" with every element zero except for 2 × 2-blocks along the diagonal for a given V.

Definition

Consider a localized one dimensional potential barrier V(x), subjected to a beam of quantum particles with energy E. These particles are incident on the potential barrier from left to right.

The solutions of Schrödinger's equation outside the potential barrier are plane waves given by

for the region to the left of the potential barrier, and

for the region to the right to the potential barrier, where

is the wave vector. The time dependence is not needed in our overview and is hence omitted. The term with coefficient A represents the incoming wave, whereas term with coefficient C represents the outgoing wave. B stands for the reflecting wave. Since we set the incoming wave moving in the positive direction (coming from the left), D is zero and can be omitted.

The "scattering amplitude", i.e., the transition overlap of the outgoing waves with the incoming waves is a linear relation defining the S-matrix,

The above relation can be written as

where

The elements of S completely characterize the scattering properties of the potential barrier V(x).

Unitary property

The unitary property of the S-matrix is directly related to the conservation of the probability current in quantum mechanics.

The probability current density J of the wave function ψ(x) is defined as

The probability current density of to the left of the barrier is

while the probability current density of to the right of the barrier is

For conservation of the probability current, JL = JR. This implies the S-matrix is a unitary matrix.

Proof

Time-reversal symmetry

If the potential V(x) is real, then the system possesses time-reversal symmetry. Under this condition, if ψ(x) is a solution of Schrödinger's equation, then ψ*(x) is also a solution.

The time-reversed solution is given by

for the region to the left to the potential barrier, and

for the region to the right to the potential barrier, where the terms with coefficient B*, C* represent incoming wave, and terms with coefficient A*, D* represent outgoing wave.

They are again related by the S-matrix,

that is,

Now, the relations

together yield a condition

This condition, in conjunction with the unitarity relation, implies that the S-matrix is symmetric, as a result of time reversal symmetry,

By combining the symmetry and the unitarity, the S-matrix can be expressed in the form:

with and . So the S-matrix is determined by three real parameters.

Transfer matrix

The transfer matrix relates the plane waves and on the right side of scattering potential to the plane waves and on the left side: [5]

and its components can be derived from the components of the S-matrix via: [6] and , whereby time-reversal symmetry is assumed.

In the case of time-reversal symmetry, the transfer matrix can be expressed by three real parameters:

with and (in case r = 1 there would be no connection between the left and the right side)

Finite square well

The one-dimensional, non-relativistic problem with time-reversal symmetry of a particle with mass m that approaches a (static) finite square well, has the potential function V with

The scattering can be solved by decomposing the wave packet of the free particle into plane waves with wave numbers for a plane wave coming (faraway) from the left side or likewise (faraway) from the right side.

The S-matrix for the plane wave with wave number k has the solution: [6]

and  ; hence and therefore and in this case.

Whereby is the (increased) wave number of the plane wave inside the square well, as the energy eigenvalue associated with the plane wave has to stay constant:

The transmission is

In the case of then and therefore and i.e. a plane wave with wave number k passes the well without reflection if for a

Finite square barrier

The square barrier is similar to the square well with the difference that for .

There are three different cases depending on the energy eigenvalue of the plane waves (with wave numbers k resp. k) far away from the barrier:

  • : In this case and the formulas for have the same form as is in the square well case, and the transmission is
  • : In this case and the wave function has the property inside the barrier and

    and

    The transmission is: . This intermediate case is not singular, it's the limit ( resp. ) from both sides.
  • :In this case is an imaginary number. So the wave function inside the barrier has the components and with .

    The solution for the S-matrix is: [7]

    and likewise: and also in this case .

    The transmission is .

Transmission coefficient and reflection coefficient

The transmission coefficient from the left of the potential barrier is, when D = 0,

The reflection coefficient from the left of the potential barrier is, when D = 0,

Similarly, the transmission coefficient from the right of the potential barrier is, when A = 0,

The reflection coefficient from the right of the potential barrier is, when A = 0,

The relations between the transmission and reflection coefficients are

and

This identity is a consequence of the unitarity property of the S-matrix.

With time-reversal symmetry, the S-matrix is symmetric and hence and .

Optical theorem in one dimension

In the case of free particles V(x) = 0, the S-matrix is [8]

Whenever V(x) is different from zero, however, there is a departure of the S-matrix from the above form, to

This departure is parameterized by two complex functions of energy, r and t. From unitarity there also follows a relationship between these two functions,

The analogue of this identity in three dimensions is known as the optical theorem.

Definition in quantum field theory

Interaction picture

A straightforward way to define the S-matrix begins with considering the interaction picture. [9] Let the Hamiltonian H be split into the free part H0 and the interaction V, H = H0 + V. In this picture, the operators behave as free field operators and the state vectors have dynamics according to the interaction V. Let

denote a state that has evolved from a free initial state

The S-matrix element is then defined as the projection of this state on the final state

Thus

where S is the S-operator. The great advantage of this definition is that the time-evolution operatorU evolving a state in the interaction picture is formally known, [10]

where T denotes the time-ordered product. Expressed in this operator,

from which

Expanding using the knowledge about U gives a Dyson series,

or, if V comes as a Hamiltonian density,

Being a special type of time-evolution operator, S is unitary. For any initial state and any final state one finds

This approach is somewhat naïve in that potential problems are swept under the carpet. [11] This is intentional. The approach works in practice and some of the technical issues are addressed in the other sections.

In and out states

Here a slightly more rigorous approach is taken in order to address potential problems that were disregarded in the interaction picture approach of above. The final outcome is, of course, the same as when taking the quicker route. For this, the notions of in and out states are needed. These will be developed in two ways, from vacua, and from free particle states. Needless to say, the two approaches are equivalent, but they illuminate matters from different angles.

From vacua

If a(k) is a creation operator, its hermitian adjoint is an annihilation operator and destroys the vacuum,

In Dirac notation, define

as a vacuum quantum state, i.e. a state without real particles. The asterisk signifies that not all vacua are necessarily equal, and certainly not equal to the Hilbert space zero state 0. All vacuum states are assumed Poincaré invariant, invariance under translations, rotations and boosts, [11] formally,

where Pμ is the generator of translation in space and time, and Mμν is the generator of Lorentz transformations. Thus the description of the vacuum is independent of the frame of reference. Associated to the in and out states to be defined are the in and out field operators (aka fields) Φi and Φo. Attention is here focused to the simplest case, that of a scalar theory in order to exemplify with the least possible cluttering of the notation. The in and out fields satisfy

the free Klein–Gordon equation. These fields are postulated to have the same equal time commutation relations (ETCR) as the free fields,

where πi,j is the field canonically conjugate to Φi,j. Associated to the in and out fields are two sets of creation and annihilation operators, ai(k) and af (k), acting in the same Hilbert space, [12] on two distinct complete sets (Fock spaces; initial space i, final space f). These operators satisfy the usual commutation rules,

The action of the creation operators on their respective vacua and states with a finite number of particles in the in and out states is given by

where issues of normalization have been ignored. See the next section for a detailed account on how a general n-particle state is normalized. The initial and final spaces are defined by

The asymptotic states are assumed to have well defined Poincaré transformation properties, i.e. they are assumed to transform as a direct product of one-particle states. [13] This is a characteristic of a non-interacting field. From this follows that the asymptotic states are all eigenstates of the momentum operator Pμ, [11]

In particular, they are eigenstates of the full Hamiltonian,

The vacuum is usually postulated to be stable and unique, [11] [nb 1]

The interaction is assumed adiabatically turned on and off.

Heisenberg picture

The Heisenberg picture is employed henceforth. In this picture, the states are time-independent. A Heisenberg state vector thus represents the complete spacetime history of a system of particles. [13] The labeling of the in and out states refers to the asymptotic appearance. A state Ψα, in is characterized by that as t → −∞ the particle content is that represented collectively by α. Likewise, a state Ψβ, out will have the particle content represented by β for t → +∞. Using the assumption that the in and out states, as well as the interacting states, inhabit the same Hilbert space and assuming completeness of the normalized in and out states (postulate of asymptotic completeness [11] ), the initial states can be expanded in a basis of final states (or vice versa). The explicit expression is given later after more notation and terminology has been introduced. The expansion coefficients are precisely the S-matrix elements to be defined below.

While the state vectors are constant in time in the Heisenberg picture, the physical states they represent are not. If a system is found to be in a state Ψ at time t = 0, then it will be found in the state U(τ)Ψ = eiHτΨ at time t = τ. This is not (necessarily) the same Heisenberg state vector, but it is an equivalent state vector, meaning that it will, upon measurement, be found to be one of the final states from the expansion with nonzero coefficient. Letting τ vary one sees that the observed Ψ (not measured) is indeed the Schrödinger picture state vector. By repeating the measurement sufficiently many times and averaging, one may say that the same state vector is indeed found at time t = τ as at time t = 0. This reflects the expansion above of an in state into out states.

From free particle states

For this viewpoint, one should consider how the archetypical scattering experiment is performed. The initial particles are prepared in well defined states where they are so far apart that they don't interact. They are somehow made to interact, and the final particles are registered when they are so far apart that they have ceased to interact. The idea is to look for states in the Heisenberg picture that in the distant past had the appearance of free particle states. This will be the in states. Likewise, an out state will be a state that in the distant future has the appearance of a free particle state. [13]

The notation from the general reference for this section, Weinberg (2002) will be used. A general non-interacting multi-particle state is given by

where

  • p is momentum,
  • σ is spin z-component or, in the massless case, helicity,
  • n is particle species.

These states are normalized as

Permutations work as such; if sSk is a permutation of k objects (for a k-particle state) such that

then a nonzero term results. The sign is plus unless s involves an odd number of fermion transpositions, in which case it is minus. The notation is usually abbreviated letting one Greek letter stand for the whole collection describing the state. In abbreviated form the normalization becomes

When integrating over free-particle states one writes in this notation

where the sum includes only terms such that no two terms are equal modulo a permutation of the particle type indices. The sets of states sought for are supposed to be complete. This is expressed as

which could be paraphrased as

where for each fixed α, the right hand side is a projection operator onto the state α. Under an inhomogeneous Lorentz transformation (Λ, a), the field transforms according to the rule

where W(Λ, p) is the Wigner rotation and D(j) is the (2j + 1)-dimensional representation of SO(3). By putting Λ = 1, a = (τ, 0, 0, 0), for which U is exp(iHτ), in (1) , it immediately follows that

so the in and out states sought after are eigenstates of the full Hamiltonian that are necessarily non-interacting due to the absence of mixed particle energy terms. The discussion in the section above suggests that the in states Ψ+ and the out states Ψ should be such that

for large positive and negative τ has the appearance of the corresponding package, represented by g, of free-particle states, g assumed smooth and suitably localized in momentum. Wave packages are necessary, else the time evolution will yield only a phase factor indicating free particles, which cannot be the case. The right hand side follows from that the in and out states are eigenstates of the Hamiltonian per above. To formalize this requirement, assume that the full Hamiltonian H can be divided into two terms, a free-particle Hamiltonian H0 and an interaction V, H = H0 + V such that the eigenstates Φγ of H0 have the same appearance as the in- and out-states with respect to normalization and Lorentz transformation properties,

The in and out states are defined as eigenstates of the full Hamiltonian,

satisfying

for τ → −∞ or τ → +∞ respectively. Define

then

This last expression will work only using wave packages.From these definitions follow that the in and out states are normalized in the same way as the free-particle states,

and the three sets are unitarily equivalent. Now rewrite the eigenvalue equation,

where the ± terms has been added to make the operator on the LHS invertible. Since the in and out states reduce to the free-particle states for V → 0, put

on the RHS to obtain

Then use the completeness of the free-particle states,

to finally obtain

Here H0 has been replaced by its eigenvalue on the free-particle states. This is the Lippmann–Schwinger equation.

In states expressed as out states

The initial states can be expanded in a basis of final states (or vice versa). Using the completeness relation,

where |Cm|2 is the probability that the interaction transforms

into

By the ordinary rules of quantum mechanics,

and one may write

The expansion coefficients are precisely the S-matrix elements to be defined below.

The S-matrix

The S-matrix is now defined by [13]

Here α and β are shorthands that represent the particle content but suppresses the individual labels. Associated to the S-matrix there is the S-operatorS defined by [13]

where the Φγ are free particle states. [13] [nb 2] This definition conforms with the direct approach used in the interaction picture. Also, due to unitary equivalence,

As a physical requirement, S must be a unitary operator. This is a statement of conservation of probability in quantum field theory. But

By completeness then,

so S is the unitary transformation from in-states to out states. Lorentz invariance is another crucial requirement on the S-matrix. [13] [nb 3] The S-operator represents the quantum canonical transformation of the initial in states to the final out states. Moreover, S leaves the vacuum state invariant and transforms in-space fields to out-space fields, [nb 4]

In terms of creation and annihilation operators, this becomes

hence

A similar expression holds when S operates to the left on an out state. This means that the S-matrix can be expressed as

If S describes an interaction correctly, these properties must be also true:

Evolution operator U

Define a time-dependent creation and annihilation operator as follows,

so, for the fields,

where

We allow for a phase difference, given by

because for S,

Substituting the explicit expression for U, one has

where is the interaction part of the Hamiltonian and is the time ordering.

By inspection, it can be seen that this formula is not explicitly covariant.

Dyson series

The most widely used expression for the S-matrix is the Dyson series. This expresses the S-matrix operator as the series:

where:

The not-S-matrix

Since the transformation of particles from black hole to Hawking radiation could not be described with an S-matrix, Stephen Hawking proposed a "not-S-matrix", for which he used the dollar sign ($), and which therefore was also called "dollar matrix". [14]

See also

Remarks

  1. This is not true if an open system is studied. Under an influence of an external field the in and out vacua can differ since the external field can produce particles.
  2. Here it is assumed that the full Hamiltonian H can be divided into two terms, a free-particle Hamiltonian H0 and an interaction V, H = H0 + V such that the eigenstates Φγ of H0 have the same appearance as the in- and out-states with respect to normalization and Lorentz transformation properties. See Weinberg (2002), page 110.
  3. If Λ is a (inhomogeneous) proper orthochronous Lorentz transformation, then Wigner's theorem guarantees the existence of a unitary operator U(Λ) acting either on HiorHf. A theory is said to be Lorentz invariant if the same U(Λ) acts on HiandHf. Using the unitarity of U(Λ), Sβα = ⟨i, β|f, α⟩ = ⟨i, β|U(Λ)U(Λ)|f, α. The right-hand side can be expanded using knowledge about how the non-interacting states transform to obtain an expression, and that expression is to be taken as a definition of what it means for the S-matrix to be Lorentz invariant. See Weinberg (2002), equation 3.3.1 gives an explicit form.
  4. Here the postulate of asymptotic completeness is employed. The in and out states span the same Hilbert space, which is assumed to agree with the Hilbert space of the interacting theory. This is not a trivial postulate. If particles can be permanently combined into bound states, the structure of the Hilbert space changes. See Greiner & Reinhardt 1996 , section 9.2.

Notes

  1. Dirac, Paul (1927-08-01). "Über die Quantenmechanik der Stoßvorgänge". Zeitschrift für Physik (in German). 44 (8): 585–595. doi:10.1007/BF01451660. ISSN   0044-3328.
  2. Sanyuk, Valerii I.; Sukhanov, Alexander D. (2003-09-01). "Dirac in 20th century physics: a centenary assessment". Physics-Uspekhi. 46 (9): 937–956. ISSN   1063-7869.
  3. John Archibald Wheeler, "On the Mathematical Description of Light Nuclei by the Method of Resonating Group Structure", Phys. Rev. 52, 1107–1122 (1937).
  4. 1 2 Jagdish Mehra, Helmut Rechenberg, The Historical Development of Quantum Theory (Pages 990 and 1031) Springer, 2001 ISBN   0-387-95086-9, ISBN   978-0-387-95086-0
  5. "Transfer Matrix Formulation of Scattering Theory in Arbitrary Dimensions" (PDF). gemma.ujf.cas.cz. Retrieved 29 October 2022.
  6. 1 2 "EE201/MSE207 Lecture 6" (PDF). intra.ece.ucr.edu. Retrieved 29 October 2022.
  7. "The Potential Barrier". quantummechanics.ucsd.edu. Retrieved 1 November 2022.
  8. Merzbacher 1961 Ch 6. A more common convention, utilized below, is to have the S-matrix go to the identity in the free particle case.
  9. Greiner & Reinhardt 1996 Section 8.2.
  10. Greiner & Reinhardt 1996 Equation 8.44.
  11. 1 2 3 4 5 Greiner & Reinhardt 1996 Chapter 9.
  12. Weinberg 2002 Chapter 3. See especially remark at the beginning of section 3.2.
  13. 1 2 3 4 5 6 7 Weinberg 2002 Chapter 3.
  14. Leonard Susskind, Black Hole War, chapter 11.

Related Research Articles

Bra–ket notation, also called Dirac notation, is a notation for linear algebra and linear operators on complex vector spaces together with their dual space both in the finite-dimensional and infinite-dimensional case. It is specifically designed to ease the types of calculations that frequently come up in quantum mechanics. Its use in quantum mechanics is quite widespread.

<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 that 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.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system that is entangled with another, without describing their combined state; this case is typical for a system interacting with some environment.
<span class="mw-page-title-main">Wave function</span> Mathematical description of the quantum state of a system

In quantum physics, a wave function is a mathematical description of the quantum state of an isolated quantum system. The most common symbols for a wave function are the Greek letters ψ and Ψ. Wave functions are complex-valued. For example, a wave function might assign a complex number to each point in a region of space. The Born rule provides the means to turn these complex probability amplitudes into actual probabilities. In one common form, it says that the squared modulus of a wave function that depends upon position is the probability density of measuring a particle as being at a given place. The integral of a wavefunction's squared modulus over all the system's degrees of freedom must be equal to 1, a condition called normalization. Since the wave function is complex-valued, only its relative phase and relative magnitude can be measured; its value does not, in isolation, tell anything about the magnitudes or directions of measurable observables. One has to apply quantum operators, whose eigenvalues correspond to sets of possible results of measurements, to the wave function ψ and calculate the statistical distributions for measurable quantities.

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

In physics, an operator is a function over a space of physical states onto another space of physical states. The simplest example of the utility of operators is the study of symmetry. Because of this, they are useful tools in classical mechanics. Operators are even more important in quantum mechanics, where they form an intrinsic part of the formulation of the theory.

<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.

<span class="mw-page-title-main">Probability amplitude</span> Complex number whose squared absolute value is a probability

In quantum mechanics, a probability amplitude is a complex number used for describing the behaviour of systems. The square of the modulus of this quantity represents a probability density.

<span class="mw-page-title-main">Fermi's interaction</span> Mechanism of beta decay proposed in 1933

In particle physics, Fermi's interaction is an explanation of the beta decay, proposed by Enrico Fermi in 1933. The theory posits four fermions directly interacting with one another. This interaction explains beta decay of a neutron by direct coupling of a neutron with an electron, a neutrino and a proton.

<span class="mw-page-title-main">Bloch sphere</span> Geometrical representation of the pure state space of a two-level quantum mechanical system

In quantum mechanics and computing, the Bloch sphere is a geometrical representation of the pure state space of a two-level quantum mechanical system (qubit), named after the physicist Felix Bloch.

<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–Zimmermann (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">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

In physics, the Majorana equation is a relativistic wave equation. It is named after the Italian physicist Ettore Majorana, who proposed it in 1937 as a means of describing fermions that are their own antiparticle. Particles corresponding to this equation are termed Majorana particles, although that term now has a more expansive meaning, referring to any fermionic particle that is its own anti-particle.

The Hückel method or Hückel molecular orbital theory, proposed by Erich Hückel in 1930, is a simple method for calculating molecular orbitals as linear combinations of atomic orbitals. The theory predicts the molecular orbitals for π-electrons in π-delocalized molecules, such as ethylene, benzene, butadiene, and pyridine. It provides the theoretical basis for Hückel's rule that cyclic, planar molecules or ions with π-electrons are aromatic. It was later extended to conjugated molecules such as pyridine, pyrrole and furan that contain atoms other than carbon and hydrogen (heteroatoms). A more dramatic extension of the method to include σ-electrons, known as the extended Hückel method (EHM), was developed by Roald Hoffmann. The extended Hückel method gives some degree of quantitative accuracy for organic molecules in general and was used to provide computational justification for the Woodward–Hoffmann rules. To distinguish the original approach from Hoffmann's extension, the Hückel method is also known as the simple Hückel method (SHM).

The time-evolving block decimation (TEBD) algorithm is a numerical scheme used to simulate one-dimensional quantum many-body systems, characterized by at most nearest-neighbour interactions. It is dubbed Time-evolving Block Decimation because it dynamically identifies the relevant low-dimensional Hilbert subspaces of an exponentially larger original Hilbert space. The algorithm, based on the Matrix Product States formalism, is highly efficient when the amount of entanglement in the system is limited, a requirement fulfilled by a large class of quantum many-body systems in one dimension.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

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.

In mathematics, the spin representations are particular projective representations of the orthogonal or special orthogonal groups in arbitrary dimension and signature. More precisely, they are two equivalent representations of the spin groups, which are double covers of the special orthogonal groups. They are usually studied over the real or complex numbers, but they can be defined over other fields.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

References