LSZ reduction formula
   HOME

TheInfoList



OR:

In quantum field theory, the LSZ reduction formula is a method to calculate ''S''-matrix elements (the
scattering amplitude In quantum physics, the scattering amplitude is the probability amplitude of the outgoing spherical wave relative to the incoming plane wave in a stationary-state scattering process.time-ordered In theoretical physics, path-ordering is the procedure (or a meta-operator \mathcal P) that orders a product of operators according to the value of a chosen parameter: :\mathcal P \left\ \equiv O_(\sigma_) O_(\sigma_) \cdots O_(\sigma_). H ...
correlation functions of a quantum field theory. It is a step of the path that starts from the
Lagrangian Lagrangian may refer to: Mathematics * Lagrangian function, used to solve constrained minimization problems in optimization theory; see Lagrange multiplier ** Lagrangian relaxation, the method of approximating a difficult constrained problem with ...
of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists
Harry Lehmann Harry Lehmann (21 March 1924 in Güstrow22 November 1998 in Hamburg) was a German physicist. Biography Lehmann studied physics at Rostock and the Humboldt-Universität zu Berlin. In 1952 he worked at the Max-Planck-Institut in Göttingen, and ...
, Kurt Symanzik and Wolfhart Zimmermann. Although the LSZ reduction formula cannot handle
bound state Bound or bounds may refer to: Mathematics * Bound variable * Upper and lower bounds, observed limits of mathematical functions Physics * Bound state, a particle that has a tendency to remain localized in one or more regions of space Geography * ...
s,
massless particle In particle physics, a massless particle is an elementary particle whose invariant mass is zero. There are two known gauge boson massless particles: the photon (carrier of electromagnetism) and the gluon (carrier of the strong force). However, g ...
s and
topological soliton A topological soliton occurs when two adjoining structures or spaces are in some way "out of phase" with each other in ways that make a seamless transition between them impossible. One of the simplest and most commonplace examples of a topological ...
s, it can be generalized to cover bound states, by use of composite fields which are often nonlocal. Furthermore, the method, or variants thereof, have turned out to be also fruitful in other fields of theoretical physics. For example, in
statistical physics Statistical physics is a branch of physics that evolved from a foundation of statistical mechanics, which uses methods of probability theory and statistics, and particularly the mathematical tools for dealing with large populations and approxim ...
they can be used to get a particularly general formulation of the
fluctuation-dissipation theorem The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the th ...
.


In and out fields

''S''-matrix elements are amplitudes of transitions between ''in'' states and ''out'' states. An ''in'' state , \\ \mathrm\rangle describes the state of a system of particles which, in a far away past, before interacting, were moving freely with definite momenta and, conversely, an ''out'' state , \\ \mathrm\rangle describes the state of a system of particles which, long after interaction, will be moving freely with definite momenta ''In'' and ''out'' states are states in
Heisenberg picture In physics, the Heisenberg picture (also called the Heisenberg representation) is a formulation (largely due to Werner Heisenberg in 1925) of quantum mechanics in which the operators (observables and others) incorporate a dependency on time, but ...
so they should not be thought to describe particles at a definite time, but rather to describe the system of particles in its entire evolution, so that the S-matrix element: :S_=\langle \\ \mathrm, \\ \mathrm\rangle is the
probability amplitude In quantum mechanics, a probability amplitude is a complex number used for describing the behaviour of systems. The modulus squared of this quantity represents a probability density. Probability amplitudes provide a relationship between the qu ...
for a set of particles which were prepared with definite momenta to interact and be measured later as a new set of particles with momenta The easy way to build ''in'' and ''out'' states is to seek appropriate field operators that provide the right
creation and annihilation operators Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator (usually d ...
. These fields are called respectively ''in'' and ''out'' fields. Just to fix ideas, suppose we deal with a Klein–Gordon field that interacts in some way which doesn't concern us: :\mathcal L= \frac 1 2 \partial_\mu \varphi\partial^\mu \varphi - \frac 1 2 m_0^2 \varphi^2 +\mathcal L_ \mathcal L_ may contain a self interaction or interaction with other fields, like a
Yukawa interaction In particle physics, Yukawa's interaction or Yukawa coupling, named after Hideki Yukawa, is an interaction between particles according to the Yukawa potential. Specifically, it is a scalar field (or pseudoscalar field) and a Dirac field of the ...
g\ \varphi\bar\psi\psi. From this
Lagrangian Lagrangian may refer to: Mathematics * Lagrangian function, used to solve constrained minimization problems in optimization theory; see Lagrange multiplier ** Lagrangian relaxation, the method of approximating a difficult constrained problem with ...
, using
Euler–Lagrange equation In the calculus of variations and classical mechanics, the Euler–Lagrange equations are a system of second-order ordinary differential equations whose solutions are stationary points of the given action functional. The equations were discovered ...
s, the equation of motion follows: :\left(\partial^2+m_0^2\right)\varphi(x)=j_0(x) where, if \mathcal L_ does not contain derivative couplings: :j_0=\frac We may expect the ''in'' field to resemble the asymptotic behaviour of the free field as , making the assumption that in the far away past interaction described by the current is negligible, as particles are far from each other. This hypothesis is named the adiabatic hypothesis. However self interaction never fades away and, besides many other effects, it causes a difference between the Lagrangian mass and the physical mass of the
boson In particle physics, a boson ( ) is a subatomic particle whose spin quantum number has an integer value (0,1,2 ...). Bosons form one of the two fundamental classes of subatomic particle, the other being fermions, which have odd half-integer spi ...
. This fact must be taken into account by rewriting the equation of motion as follows: :\left(\partial^2+m^2\right)\varphi(x)=j_0(x)+\left(m^2-m_0^2\right)\varphi(x)=j(x) This equation can be solved formally using the retarded
Green's function In mathematics, a Green's function is the impulse response of an inhomogeneous linear differential operator defined on a domain with specified initial conditions or boundary conditions. This means that if \operatorname is the linear differenti ...
of the Klein–Gordon operator \partial^2+m^2: :\Delta_(x)=i\theta\left(x^0\right)\int \frac \left(e^-e^\right)_\qquad \omega_k=\sqrt allowing us to split interaction from asymptotic behaviour. The solution is: :\varphi(x)=\sqrt Z \varphi_(x) +\int \mathrm^4y \Delta_(x-y)j(y) The factor is a normalization factor that will come handy later, the field is a solution of the
homogeneous equation In mathematics, a system of linear equations (or linear system) is a collection of one or more linear equations involving the same variable (math), variables. For example, :\begin 3x+2y-z=1\\ 2x-2y+4z=-2\\ -x+\fracy-z=0 \end is a system of three ...
associated with the equation of motion: :\left(\partial^2+m^2\right) \varphi_(x)=0, and hence is a
free field In physics a free field is a field without interactions, which is described by the terms of motion and mass. Description In classical physics, a free field is a field whose equations of motion are given by linear partial differential equat ...
which describes an incoming unperturbed wave, while the last term of the solution gives the perturbation of the wave due to interaction. The field is indeed the ''in'' field we were seeking, as it describes the asymptotic behaviour of the interacting field as , though this statement will be made more precise later. It is a free scalar field so it can be expanded in plane waves: :\varphi_(x)=\int \mathrm^3k \left\ where: :f_k(x)=\left.\frac\_ The inverse function for the coefficients in terms of the field can be easily obtained and put in the elegant form: :a_(\mathbf)=i\int \mathrm^3x f^*_k(x)\overleftrightarrow\varphi_(x) where: :\overleftrightarrow f = \mathrm\partial_0 f -f\partial_0 \mathrm. The
Fourier coefficient A Fourier series () is a summation of harmonically related sinusoidal functions, also known as components or harmonics. The result of the summation is a periodic function whose functional form is determined by the choices of cycle length (or ''p ...
s satisfy the algebra of
creation and annihilation operators Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator (usually d ...
: : _(\mathbf),a_(\mathbf)0;\quad _(\mathbf),a^\dagger_(\mathbf)\delta^3(\mathbf-\mathbf); and they can be used to build ''in'' states in the usual way: :\left, k_1,\ldots,k_n\ \mathrm\right\rangle=\sqrta_^\dagger(\mathbf_1)\ldots \sqrta_^\dagger(\mathbf_n), 0\rangle The relation between the interacting field and the ''in'' field is not very simple to use, and the presence of the retarded Green's function tempts us to write something like: :\varphi(x)\sim\sqrt Z\varphi_(x)\qquad \mathrm\quad x^0\to-\infty implicitly making the assumption that all interactions become negligible when particles are far away from each other. Yet the current contains also self interactions like those producing the mass shift from to . These interactions do not fade away as particles drift apart, so much care must be used in establishing asymptotic relations between the interacting field and the ''in'' field. The correct prescription, as developed by Lehmann, Symanzik and Zimmermann, requires two normalizable states , \alpha\rangle and , \beta\rangle, and a normalizable solution of the Klein–Gordon equation (\partial^2+m^2)f(x)=0. With these pieces one can state a correct and useful but very weak asymptotic relation: :\lim_ \int \mathrm^3x \langle\alpha, f(x)\overleftrightarrow\varphi(x), \beta\rangle= \sqrt Z \int \mathrm^3x \langle\alpha, f(x)\overleftrightarrow\varphi_(x), \beta\rangle The second member is indeed independent of time as can be shown by differentiating and remembering that both and satisfy the Klein–Gordon equation. With appropriate changes the same steps can be followed to construct an ''out'' field that builds ''out'' states. In particular the definition of the ''out'' field is: :\varphi(x)=\sqrt Z \varphi_(x) +\int \mathrm^4y \Delta_(x-y)j(y) where is the advanced Green's function of the Klein–Gordon operator. The weak asymptotic relation between ''out'' field and interacting field is: : \lim_ \int \mathrm^3x \langle\alpha, f(x)\overleftrightarrow\varphi(x), \beta\rangle= \sqrt Z \int \mathrm^3x \langle\alpha, f(x)\overleftrightarrow\varphi_(x), \beta\rangle


The reduction formula for scalars

The asymptotic relations are all that is needed to obtain the LSZ reduction formula. For future convenience we start with the matrix element: :\mathcal M=\langle \beta\ \mathrm, \mathrm\varphi(y_1)\ldots\varphi(y_n), \alpha\ \mathrm\rangle which is slightly more general than an ''S''-matrix element. Indeed, \mathcal M is the expectation value of the time-ordered product of a number of fields \varphi(y_1)\cdots\varphi(y_n) between an ''out'' state and an ''in'' state. The ''out'' state can contain anything from the vacuum to an undefined number of particles, whose momenta are summarized by the index . The ''in'' state contains at least a particle of momentum , and possibly many others, whose momenta are summarized by the index . If there are no fields in the time-ordered product, then \mathcal M is obviously an ''S''-matrix element. The particle with momentum can be 'extracted' from the ''in'' state by use of a creation operator: : \mathcal M=\sqrt\ \left \langle \beta\ \mathrm \bigg, \mathrm T\left varphi(y_1)\ldots\varphi(y_n)\righta_^\dagger(\mathbf p) \bigg, \alpha'\ \mathrm \right \rangle where the prime on \alpha denotes that one particle has been taken out. With the assumption that no particle with momentum is present in the ''out'' state, that is, we are ignoring forward scattering, we can write: :\mathcal M=\sqrt\ \left \langle \beta\ \mathrm \bigg, \left\ \bigg, \alpha'\ \mathrm \right \rangle because a_^\dagger acting on the left gives zero. Expressing the construction operators in terms of ''in'' and ''out'' fields, we have: :\mathcal M=-i\sqrt\ \int \mathrm^3x f_p(x)\overleftrightarrow \left\langle \beta\ \mathrm \bigg, \left\ \bigg, \alpha'\ \mathrm\right \rangle Now we can use the asymptotic condition to write: :\mathcal M= -i\sqrt \left\ Then we notice that the field can be brought inside the time-ordered product, since it appears on the right when and on the left when : :\mathcal M=-i\sqrt \left(\lim_-\lim_\right) \int \mathrm^3x f_p(x) \overleftrightarrow \langle \beta\ \mathrm, \mathrm T\left varphi(x)\varphi(y_1)\ldots\varphi(y_n)\right, \alpha'\ \mathrm \rangle In the following, dependence in the time-ordered product is what matters, so we set: :\langle \beta\ \mathrm, \mathrm T\left varphi(x)\varphi(y_1)\ldots\varphi(y_n)\right, \alpha'\ \mathrm\rangle= \eta(x) It's easy to show by explicitly carrying out the time integration that: : \mathcal M=i\sqrt \int \mathrm(x^0)\partial_0 \int \mathrm^3x f_p(x)\overleftrightarrow\eta(x) so that, by explicit time derivation, we have: :\mathcal M=i\sqrt \int \mathrm^4 x\left\ By its definition we see that is a solution of the Klein–Gordon equation, which can be written as: :\partial_0^2f_p(x)=\left(\Delta-m^2\right) f_p(x) Substituting into the expression for \mathcal M and integrating by parts, we arrive at: :\mathcal M=i\sqrt \int \mathrm^4 x f_p(x)\left(\partial_0^2-\Delta+m^2\right)\eta(x) That is: : \mathcal M=\frac \int \mathrm^4 x e^ \left(\Box+m^2\right)\langle \beta\ \mathrm, \mathrm T\left varphi(x)\varphi(y_1)\ldots\varphi(y_n)\right, \alpha'\ \mathrm\rangle Starting from this result, and following the same path another particle can be extracted from the ''in'' state, leading to the insertion of another field in the time-ordered product. A very similar routine can extract particles from the ''out'' state, and the two can be iterated to get vacuum both on right and on left of the time-ordered product, leading to the general formula: :\langle p_1,\ldots,p_n\ \mathrm, q_1,\ldots,q_m\ \mathrm\rangle=\int \prod_^ \left\ \prod_^ \left\ \langle \Omega, \mathrm \varphi(x_1)\ldots\varphi(x_m)\varphi(y_1)\ldots\varphi(y_n), \Omega\rangle Which is the LSZ reduction formula for Klein–Gordon scalars. It gains a much better looking aspect if it is written using the Fourier transform of the correlation function: : \Gamma \left (p_1,\ldots,p_n \right )=\int \prod_^ \left\ \langle \Omega, \mathrm\ \varphi(x_1)\ldots\varphi(x_n), \Omega\rangle Using the inverse transform to substitute in the LSZ reduction formula, with some effort, the following result can be obtained: :\langle p_1,\ldots,p_n\ \mathrm, q_1,\ldots,q_m\ \mathrm\rangle= \prod_^ \left\ \prod_^ \left\ \Gamma \left (p_1,\ldots,p_n;-q_1,\ldots,-q_m \right ) Leaving aside normalization factors, this formula asserts that ''S''-matrix elements are the residues of the poles that arise in the Fourier transform of the correlation functions as four-momenta are put on-shell.


Reduction formula for fermions

Recall that solutions to the quantized free-field
Dirac equation 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- massive particles, called "Dirac par ...
may be written as :\Psi(x)=\sum_\int\!\mathrm\tilde\big(b^s_\textbfu^s_\textbf\mathrm^+d^_\textbfv^s_\textbf\mathrm^\big), where the metric signature is mostly plus, b^s_\textbf is an annihilation operator for b-type particles of momentum \textbf and spin s=\pm, d^_\textbf is a creation operator for d-type particles of spin s, and the spinors u^s_\textbf and v^s_\textbf satisfy (p\!\!\!/+m)u^s_\textbf=0 and (p\!\!\!/-m)v^s_\textbf = 0. The Lorentz-invariant measure is written as \mathrm\tilde:=\mathrm^3 p/(2\pi)^3 2\omega_\textbf, with \omega_\textbf = \sqrt. Consider now a scattering event consisting of an ''in'' state , \alpha\ \mathrm\rangle of non-interacting particles approaching an interaction region at the origin, where scattering occurs, followed by an ''out'' state , \beta\ \mathrm\rangle of outgoing non-interacting particles. The probability amplitude for this process is given by :\mathcal = \langle \beta\ \mathrm, \alpha\ \mathrm\rangle, where no extra time-ordered product of field operators has been inserted, for simplicity. The situation considered will be the scattering of n b-type particles to n' b-type particles. Suppose that the ''in'' state consists of n particles with momenta \ and spins \, while the ''out'' state contains particles of momenta \ and spins \. The ''in'' and ''out'' states are then given by :, \alpha\ \mathrm\rangle = , \textbf_1^,...,\textbf_n^\rangle\quad\text\quad, \beta\ \mathrm\rangle = , \textbf_1^,...,\textbf_^\rangle. Extracting an ''in'' particle from , \alpha\ \mathrm\rangle yields a free-field creation operator b^_ acting on the state with one less particle. Assuming that no outgoing particle has that same momentum, we then can write :\mathcal = \langle\beta\ \mathrm, b^_-b^_, \alpha'\ \mathrm\rangle, where the prime on \alpha denotes that one particle has been taken out. Now recall that in the free theory, the b-type particle operators can be written in terms of the field using the inverse relation :b^_\textbf = \int\!\mathrm^3 x\;\mathrm^\bar(x)\gamma^0 u^s_\textbf, where \bar(x)=\Psi^\dagger(x)\gamma^0. Denoting the asymptotic free fields by \Psi_\text and \Psi_\text, we find :\mathcal = \int\!\mathrm^3x_1\;\mathrm^\langle\beta\ \mathrm, \bar_\text(x_1)\gamma^0 u^_-\bar_\text(x_1)\gamma^0 u^_, \alpha'\ \mathrm\rangle. The weak asymptotic condition needed for a Dirac field, analogous to that for scalar fields, reads :\lim_\int\!\mathrm^3 x\langle \beta, \mathrm^\bar(x)\gamma^0 u^_, \alpha\rangle=\sqrt\int\!\mathrm^3 x\langle \beta, \mathrm^\bar_\text(x)\gamma^0 u^_, \alpha\rangle, and likewise for the ''out'' field. The scattering amplitude is then :\mathcal = \frac\Big(\lim_-\lim_\Big)\int\!\mathrm^3 x_1\;\mathrm^\langle\beta\ \mathrm, \bar(x_1)\gamma^0 u^_, \alpha'\ \mathrm\rangle, where now the interacting field appears in the inner product. Rewriting the limits in terms of the integral of a time derivative, we have :\mathcal = -\frac\int\!\mathrm^4 x_1\partial_0\big(\mathrm^\langle\beta\ \mathrm, \bar(x_1)\gamma^0 u^_, \alpha'\ \mathrm\rangle\big) :: =-\frac\int\!\mathrm^4 x_1(\partial_0\mathrm^\eta(x_1)+\mathrm^\partial_0\eta(x_1)\big)\gamma^0 u^_, where the row vector of matrix elements of the barred Dirac field is written as \eta(x_1):=\langle\beta\ \mathrm, \bar(x_1), \alpha'\ \mathrm\rangle. Now, recall that \mathrm^u^s_\textbf is a solution to the Dirac equation: :(-i\partial\!\!\!/+m)\mathrm^u^s_\textbf=0. Solving for \gamma^0\partial_0 \mathrm^u^s_\textbf, substituting it into the first term in the integral, and performing an integration by parts, yields :\mathcal = \frac\int\!\mathrm^4x_1\mathrm^\big(i\partial_\mu\eta(x_1)\gamma^\mu + \eta(x_1)m\big)u^_. Switching to Dirac index notation (with sums over repeated indices) allows for a neater expression, in which the quantity in square brackets is to be regarded as a differential operator: :\mathcal = \frac\int\!\mathrm^4x_1\mathrm^ i_ + m)u^_\langle\beta\ \mathrm, \bar_(x_1), \alpha'\ \mathrm\rangle. Consider next the matrix element appearing in the integral. Extracting an ''out'' state creation operator and subtracting the corresponding ''in'' state operator, with the assumption that no incoming particle has the same momentum, we have :\langle\beta\ \mathrm, \bar_(x_1), \alpha'\ \mathrm\rangle = \langle\beta'\ \mathrm, b^_\bar_(x_1) - \bar_(x_1)b^_, \alpha'\ \mathrm\rangle. Remembering that (\bar\gamma^0u^s_\textbf)^\dagger = \bar^s_\textbf\gamma^0\Psi, where \bar^s_\textbf:=u^_\textbf\beta, we can replace the annihilation operators with ''in'' fields using the adjoint of the inverse relation. Applying the asymptotic relation, we find :\langle\beta\ \mathrm, \bar_(x_1), \alpha'\ \mathrm\rangle =\frac\Big(\lim_-\lim_\Big)\int\!\mathrm^3 y_1\mathrm^ bar^_\gamma^0\langle\beta'\ \mathrm, \mathrm Psi_(y_1)\bar_(x_1)\alpha'\ \mathrm\rangle. Note that a time-ordering symbol has appeared, since the first term requires \Psi_(y_1) on the left, while the second term requires it on the right. Following the same steps as before, this expression reduces to :\langle\beta\ \mathrm, \bar_(x_1), \alpha'\ \mathrm\rangle =\frac\int\!\mathrm^4y_1\mathrm^ bar^_(-i\partial\!\!\!/_+m)\langle\beta'\ \mathrm, \mathrm Psi_(y_1)\bar_(x_1)\alpha'\ \mathrm\rangle. The rest of the ''in'' and ''out'' states can then be extracted and reduced in the same way, ultimately resulting in :\langle \beta\ \mathrm, \alpha\ \mathrm\rangle=\int\!\prod_^n \mathrm^4 x_j \frac i_+m)u^_\prod_^\mathrm^4 y_l\frac bar^_(-i_+m) \langle 0, \mathrm Psi_(y_1)...\Psi_(y_)\bar_(x_1)...\bar_(x_n)0\rangle. The same procedure can be done for the scattering of d-type particles, for which u^s_\textbf's are replaced by v^s_\textbf's, and \Psi's and \bar's are swapped.


Field strength normalization

The reason of the normalization factor in the definition of ''in'' and ''out'' fields can be understood by taking that relation between the vacuum and a single particle state , p\rangle with four-moment on-shell: :\langle 0, \varphi(x), p\rangle= \sqrt Z \langle 0, \varphi_(x), p\rangle + \int \mathrm^4y \Delta_(x-y) \langle 0, j(y), p\rangle Remembering that both and are scalar fields with their Lorentz transform according to: :\varphi(x)=e^\varphi(0)e^ where is the four-momentum operator, we can write: : e^\langle 0, \varphi(0), p\rangle= \sqrt Z e^ \langle 0, \varphi_(0), p\rangle + \int \mathrm^4y \Delta_(x-y)\langle 0, j(y), p\rangle Applying the Klein–Gordon operator on both sides, remembering that the four-moment is on-shell and that is the Green's function of the operator, we obtain: : 0=0 + \int \mathrm^4y \delta^4(x-y) \langle 0, j(y), p\rangle; \quad\Leftrightarrow\quad \langle 0, j(x), p\rangle=0 So we arrive to the relation: :\langle 0, \varphi(x), p\rangle= \sqrt Z \langle 0, \varphi_(x), p\rangle which accounts for the need of the factor . The ''in'' field is a free field, so it can only connect one-particle states with the vacuum. That is, its expectation value between the vacuum and a many-particle state is null. On the other hand, the interacting field can also connect many-particle states to the vacuum, thanks to interaction, so the expectation values on the two sides of the last equation are different, and need a normalization factor in between. The right hand side can be computed explicitly, by expanding the ''in'' field in creation and annihilation operators: :\langle 0, \varphi_(x), p\rangle= \int \frac e^ \langle 0, a_(\mathbf q), p\rangle= \int \frac e^ \langle 0, a_(\mathbf q)a^\dagger_(\mathbf p), 0\rangle Using the commutation relation between and a^\dagger_ we obtain: : \langle 0, \varphi_(x), p\rangle= \frac leading to the relation: :\langle 0, \varphi(0), p\rangle= \sqrt \frac by which the value of may be computed, provided that one knows how to compute \langle 0, \varphi(0), p\rangle.


References

* The original paper is: * A pedagogical derivation of the LSZ reduction formula can be found in: Peskin and Schroeder, Section 7.2, also in Srednicki, Section I.5, or in Weinberg, pp. 436–438. {{DEFAULTSORT:Lsz Reduction Formula Quantum field theory