Light-front computational methods

The light front quantization[1]
[2]
[3]
of quantum field theories
provides a useful alternative to ordinary equal-time
quantization. In
particular, it can lead to a relativistic description of bound systems
in terms of quantum-mechanical wave functions. The quantization is
based on the choice of light-front coordinates,[4]
where
plays the role of time and the corresponding spatial
coordinate is
. Here,
is the ordinary time,
is one Cartesian coordinate,
and
is the speed of light. The other
two Cartesian coordinates,
and
, are untouched and often called
transverse or perpendicular, denoted by symbols of the type
. The choice of the
frame of reference where the time
and
-axis are defined can be left unspecified in an exactly
soluble relativistic theory, but in practical calculations some choices may be more suitable than others.
The solution of the LFQCD Hamiltonian eigenvalue equation will utilize the available mathematical methods of quantum mechanics and contribute to the development of advanced computing techniques for large quantum systems, including nuclei. For example, in the discretized light-cone quantization method (DLCQ),[5] [6] [7] [8] [9] [10] periodic conditions are introduced such that momenta are discretized and the size of the Fock space is limited without destroying Lorentz invariance. Solving a quantum field theory is then reduced to diagonalizing a large sparse Hermitian matrix. The DLCQ method has been successfully used to obtain the complete spectrum and light-front wave functions in numerous model quantum field theories such as QCD with one or two space dimensions for any number of flavors and quark masses. An extension of this method to supersymmetric theories, SDLCQ,[11] [12] takes advantage of the fact that the light-front Hamiltonian can be factorized as a product of raising and lowering ladder operators. SDLCQ has provided new insights into a number of supersymmetric theories including direct numerical evidence[13] for a supergravity/super-Yang—Mills duality conjectured by Maldacena.
It is convenient to work in a Fock basis
where the light-front
momenta
and
are diagonal.
The state
is given by an expansion
with
is
interpreted as the wave function of the contribution from states
with
particles. The eigenvalue problem
is a set of coupled integral equations for
these wave functions. Although the notation as presented supports
only one particle type, the generalization to more than one is trivial.
Discrete light-cone quantization
A systematic approach to discretization of the eigenvalue problem is the DLCQ method originally suggested by Pauli and Brodsky.[5][6] In essence it is the replacement of integrals by trapezoidal approximations, with equally-spaced intervals in the longitudinal and transverse momenta
corresponding to periodic boundary conditions on the
intervals and
.
The length scales
and
determine the resolution of the
calculation. Because the plus component of momentum is always
positive, the limit
can be exchanged for a limit
in terms of the integer {\em resolution}
.
The combination of momentum components that defines
is then
independent of
. The longitudinal momentum fractions
become
ratios of integers
. Because the
are all positive, DLCQ
automatically limits the number of particles to be no more than
.
When a limit on transverse momentum is supplied via a chosen cutoff,
a finite matrix problem is obtained; however, the matrix may be too
large for present numerical techniques. An explicit truncation in
particle number, the light-cone equivalent
of the Tamm—Dancoff approximation, can then be made.
Large basis sizes require special techniques for matrix diagonalization;
the one typically used is the Lanczos algorithm.
For the case of one space dimension, one
can readily solve for the hadron spectrum of QCD for any quark masses
and colors.
Most DLCQ calculations are done without zero modes. However, in principle, any DLCQ basis with periodic boundary conditions may include them as constrained modes, dependent on the other modes with nonzero momentum. The constraint comes from the spatial average of the Euler-Lagrange equation for the field. This constraint equation can be difficult to solve, even for the simplest theories. However, an approximate solution can be found, consistent with the underlying approximations of the DLCQ method itself.[14] This solution generates the effective zero-mode interactions for the light-front Hamiltonian.
Calculations in the massive sector that are done without zero modes will usually yield the correct answer. The neglect of zero modes merely worsens the convergence. One exception is that of cubic scalar theories, where the spectrum extends to minus infinity. A DLCQ calculation without zero modes will require careful extrapolation to detect this infinity, whereas a calculation that includes zero modes yields the correct result immediately. The zero modes are avoided if one uses antiperiodic boundary conditions.
Supersymmetric discrete light-cone quantization
The supersymmetric form of DLCQ
(SDLCQ)[11][12]
is specifically designed
to maintain supersymmetry in the discrete approximation.
Ordinary DLCQ violates supersymmetry by terms that do not survive
the continuum limit. The SDLCQ construction discretizes the
supercharge and {\em defines} the Hamiltonian
by the superalgebra relation
.
The range of transverse
momentum is limited by a simple cutoff in the momentum value.
Effects of zero modes are expected to cancel.
In addition to calculations of spectra, this technique can
be used to calculate expectation values. One such quantity,
a correlator
of
the stress energy tensor, has been
computed as a test of a Maldacena conjecture. A very efficient
Lanczos-based method was developed for this calculation. The most
recent results provide direct evidence for the
conjecture.[13]
Transverse lattice
The transverse lattice
method[15][16]
brings together two powerful ideas in quantum field theory: light-front
Hamiltonian quantization and lattice gauge theory.
Lattice gauge theory is a very popular means of regulating
for calculation the gauge theories that describe all visible
matter in the universe; in particular, it manifestly demonstrates
the linear confinement of QCD that holds quarks and gluons inside
the protons and neutrons of the atomic nucleus. In general, to
obtain solutions of a quantum field theory, with its continuously
infinite degrees of freedom, one must put kinematical cutoffs
or other restrictions on the space of quantum states. To remove the
errors this introduces, one may then extrapolate these cutoffs,
provided a continuum limit exists, and/or
renormalize observables to account for degrees of freedom above
the cutoff. For the purposes of Hamiltonian quantization, one
must have a continuous time direction. In the case of light-front
Hamiltonian quantization, in addition to continuous light-front
time , it is necessary to keep the
direction continuous
if one wants to preserve the manifest Lorentz boost invariance in
one direction and to include small light-front energies
.
Therefore, at most one can impose a lattice cutoff on the
remaining transverse spatial directions. Such a transverse
lattice gauge theory was first suggested by Bardeen and
Pearson in 1976.[15]
Most practical calculations performed with transverse lattice
gauge theory have utilized one further ingredient:
the color-dielectric expansion. A dielectric formulation is one
in which the gauge group elements, whose generators are the gluon
fields in the case of QCD, are replaced by collective
(smeared, blocked, etc.) variables which represent an
average over their fluctuations on short distance scales.
These dielectric variables are massive, carry
color, and form an effective gauge field theory with classical
action minimized at zero field, meaning that color flux is expelled
from the vacuum at the classical level. This maintains the triviality
of the light-front vacuum structure, but arises only for a low momentum
cutoff on the effective theory (corresponding to transverse lattice
spacings of order 1/2 fm in QCD). As a result, the effective cutoff
Hamiltonian is initially poorly constrained.
The color-dielectric expansion, together with requirements of
Lorentz symmetry restoration, has nevertheless been
successfully used to organize the interactions in the Hamiltonian
in a way suitable for practical solution. The most accurate spectrum
of large- glueballs has been obtained in this way, and as well
as pion light-front wave functions in agreement with a range of experimental data.
Basis Light-Front Quantization
The basis light-front quantization (BLFQ)
approach[17]
uses expansions in products of single-particle basis functions to represent
the Fock-state wave functions. Typically, the longitudinal ()
dependence is represented in the DLCQ basis of plane waves, and
the transverse dependence is represented by two-dimensional
harmonic oscillator functions. The latter are ideal for
applications to confining cavities and are consistent with
light-front holographic QCD.[18]
[19]
[20]
[21]
[22]
The use of products of single
particle basis functions is also convenient for incorporation
of boson and fermion statistics, because the products are
readily (anti)symmetrized. By employing two-dimensional basis
functions with rotational symmetry about the longitudinal
direction (where the harmonic oscillator functions serve as
an example), one preserves the total angular momentum projection
quantum number which facilitates determination
of the total angular momentum of the mass eigenstates.
For applications without an
external cavity, where transverse momentum is conserved,
a Lagrange multiplier method is used to separate the
relative transverse motion from the total system's motion.
The first application of BLFQ to QED solved for the electron in a two-dimensional transverse confining cavity and showed how the anomalous magnetic moment behaved as a function of the strength of the cavity.[23] The second application of BLFQ to QED solved for the electron's anomalous magnetic moment in free space[24] [25] and demonstrated agreement with the Schwinger moment in the appropriate limit.
The extension of BLFQ to the time-dependent regime, namely, time-dependent BLFQ (tBLFQ) is straightforward and is currently under active development. The goal of tBLFQ is to solve light-front field theory in real-time (with or without time-dependent background fields). The typical application areas include intense lasers (see Light-front quantization#Intense lasers}) and relativistic heavy-ion collisions.
Light-front coupled-cluster method
The light-front coupled cluster (LFCC) method[26] is a particular form of truncation for the infinite coupled system of integral equations for light-front wave functions. The system of equations that comes from the field-theoretic Schrödinger equation also requires regularization, to make the integral operators finite. The traditional Fock-space truncation of the system, where the allowed number of particles is limited, typically disrupts the regularization by removing infinite parts that would otherwise cancel against parts that are retained. Although there are ways to circumvent this, they are not completely satisfactory.
The LFCC method avoids these difficulties by truncating
the set of equations in a very different way. Instead
of truncating the number of particles, it truncates
the way in which wave functions are related to each
other; the wave functions of higher Fock states are
determined by the lower-state wave functions and the
exponentiation of an operator . Specifically,
the eigenstate is written in the form
,
where
is a normalization factor and
is a state with the minimal number
of constituents. The operator
increases particle
number and conserves all relevant quantum numbers,
including light-front momentum. This is in principle
exact but also still infinite, because
can have an
infinite number of terms. Zero modes can be included
by inclusion of their creation as terms in
; this
generates a nontrivial vacuum as a generalized
coherent state of zero modes.
The truncation made is a truncation of .
The original eigenvalue problem becomes a finite-sized eigenvalue
problem for the valence state
, combined
with auxiliary equations for the terms retained in
:
Here is a projection onto the valence sector, and
is the LFCC effective
Hamiltonian. The projection
is truncated to
provide just enough auxiliary equations to determine
the functions in the truncated
operator. The effective
Hamiltonian is computed from its Baker--Hausdorff expansion
,
which can be terminated at the point where more particles
are being created than are kept by the truncated
projection
. The use of the exponential of
rather
than some other function is convenient, not only because
of the Baker—Hausdorff expansion but more generally because
it is invertible; in principle, other functions could be used
and would also provide an exact representation until a
truncation is made.
The truncation of can be handled systematically. Terms
can be classified by the number of annihilated constituents
and the net increase in particle number. For example, in QCD
the lowest-order contributions annihilate one particle and increase
the total by one. These are one-gluon emission from a quark,
quark pair creation from one gluon, and gluon pair creation
from one gluon. Each involves a function of relative momentum
for the transition from one to two particles. Higher order terms
annihilate more particles and/or increase the total by
more than one. These provide additional contributions to
higher-order wave functions and even to low-order wave
functions for more complicated valence states. For example,
the wave function for the
Fock state
of a meson can have a contribution from a term in
that annihilates a
pair and creates a
pair plus a gluon, when this acts on the meson
valence state
.
The mathematics of the LFCC method has its origin in the
many-body coupled cluster method used in
nuclear physics and quantum chemistry.[27]
The physics is, however, quite different. The many-body
method works with a state of a large number
of particles and uses the exponentiation of to
build in correlations of excitations to higher
single-particle states; the particle number does not change.
The LFCC method starts from a small number of constituents
in a valence state and uses
to build states with
more particles; the method of solution of the valence-state
eigenvalue problem is left unspecified.
The computation of physical observables from matrix elements of operators requires some care. Direct computation would require an infinite sum over Fock space. One can instead borrow from the many-body coupled cluster method[27] a construction that computes expectation values from right and left eigenstates. This construction can be extended to include off-diagonal matrix elements and gauge projections. Physical quantities can then be computed from the right and left LFCC eigenstates.
Renormalization group
Renormalization concepts, especially the renormalization group methods in quantum theories and statistical mechanics, have a long history and a very broad scope. The concepts of renormalization that appear useful in theories quantized in the front form of dynamics are essentially of two types, as in other areas of theoretical physics. The two types of concepts are associated with two types of theoretical tasks involved in applications of a theory. One task is to calculate observables (values of operationally defined quantities) in a theory that is unambiguously defined. The other task is to define a theory unambiguously. This is explained below.
Since the front form of dynamics aims at explaining hadrons as bound states of quarks and gluons, and the binding mechanism is not describable using perturbation theory, the definition of a theory needed in this case cannot be limited to perturbative expansions. For example, it is not sufficient to construct a theory using regularization of loop integrals order-by-order and correspondingly redefining the masses, coupling constants, and field normalization constants also order-by-order. In other words, one needs to design the Minkowski space-time formulation of a relativistic theory that is not based on any a priori perturbative scheme. The front form of Hamiltonian dynamics is perceived by many researchers as the most suitable framework for this purpose among the known options.[1] [2][3]
The desired definition of a relativistic theory involves calculations of as many observables as one must use in order to fix all the parameters that appear in the theory. The relationship between the parameters and observables may depend on the number of degrees of freedom that are included in the theory.
For example, consider virtual particles in a candidate formulation of the theory. Formally, special relativity requires that the range of momenta of the particles is infinite because one can change the momentum of a particle by an arbitrary amount through a change of frame of reference. If the formulation is not to distinguish any inertial frame of reference, the particles must be allowed to carry any value of momentum. Since the quantum field modes corresponding to particles with different momenta form different degrees of freedom, the requirement of including infinitely many values of momentum means that one requires the theory to involve infinitely many degrees of freedom. But for mathematical reasons, being forced to use computers for sufficiently precise calculations, one has to work with a finite number of degrees of freedom. One must limit the momentum range by some cutoff.
Setting up a theory with a finite cutoff for mathematical reasons, one hopes that the cutoff can be made sufficiently large to avoid its appearance in observables of physical interest, but in local quantum field theories that are of interest in hadronic physics the situation is not that simple. Namely, particles of different momenta are coupled through the dynamics in a nontrivial way, and the calculations aiming at predicting observables yield results that depend on the cutoffs. Moreover, they do so in a diverging fashion.
There may be more cutoff parameters than just for momentum. For example, one may assume that the volume of space is limited, which would interfere with translation invariance of a theory, or assume that the number of virtual particles is limited, which would interfere with the assumption that every virtual particle may split into more virtual particles. All such restrictions lead to a set of cutoffs that becomes a part of a definition of a theory.
Consequently, every result of a calculation for any
observable characterized by its
physical scale
has the form of a function of the
set of parameters of the theory,
, the set of cutoffs,
say
, and the scale
. Thus, the results
take the form
However, experiments provide values of observables
that characterize natural processes irrespective of
the cutoffs in a theory used to explain them. If the
cutoffs do not describe properties of nature and are
introduced merely for making a theory computable, one
needs to understand how the dependence on
may drop out from
.
The cutoffs may also reflect some natural features of
a physical system at hand, such as in the model case
of an ultraviolet cutoff on the wave vectors of sound
waves in a crystal due to the spacing of atoms in
the crystal lattice. The natural cutoffs may be of
enormous size in comparison to the scale
. Then,
one faces the question of how it happens in the theory
that its results for observables at scale
are not
also of the enormous size of the cutoff and, if they
are not, then how they depend on the scale
.
The two types of concepts of renormalization mentioned above are associated with the following two questions:
- How should the parameters
depend on the cutoffs
so that all observables
of physical interest do not depend on
, including the case where one removes the cutoffs by sending them formally to infinity?
- What is the required set of parameters
?
The renormalization group concept associated
with the first
question[28]
[29]
predates the concept associated with the second
question.[30]
[31]
[32]
[33]
Certainly, if one were in possession of a
good answer to the second question, the first
question could also be answered. In the absence
of a good answer to the second question, one
may wonder why any specific choice of parameters
and their cutoff dependence could secure cutoff
independence of all observables
with finite scales
.
The renormalization group concept associated
with the first question above relies on the
circumstance that some finite set
yields the desired result,
In this way of thinking, one can expect that
in a theory with parameters a calculation
of
observables at some scale
is
sufficient to fix all parameters as functions
of
. So, one may hope that there exists
a collection of
effective parameters at
scale
, corresponding to
observables
at scale
, that are sufficient to parametrize
the theory in such a way that predictions expressed
in terms of these parameters are free from dependence
on
. Since the scale
is arbitrary,
a whole family of such
-parameter sets labeled
by
should exist, and every member of that
family corresponds to the same physics. Moving
from one such family to another by changing one
value of
to another is described as action
of ``the renormalization group. The word group
is justified because the group axioms are
satisfied: two such changes form another such
change, one can invert a change, etc.
The question remains, however, why fixing the
cutoff dependence of parameters
on
, using
conditions that
selected
observables do not depend on
, is good
enough to make all observables in the physical
range of
not depend on
. In some
theories such a miracle may happen but in others
it may not. The ones where it happens are
called renormalizable, because one can normalize
the parameters properly to obtain cutoff independent
results.
Typically, the set is established
using perturbative calculations that are combined
with models for description of nonperturbative
effects. For example, perturbative QCD diagrams
for quarks and gluons are combined with the parton
models for description of binding of quarks and
gluons into hadrons. The set of parameters
includes cutoff dependent masses, charges and field
normalization constants. The predictive power of
a theory set up this way relies on the circumstance
that the required set of parameters is relatively
small. The regularization is designed order-by-order
so that as many formal symmetries as possible of a
local theory are preserved and employed in calculations,
as in the dimensional regularization of Feynman
diagrams. The claim that the set of parameters
leads to finite, cutoff independent
limits for all observables is qualified by the
need to use some form of perturbation theory and
inclusion of model assumptions concerning bound
states.
The renormalization group concept associated with
the second question above is conceived to explain
how it may be so that the concept of
renormalization group associated with the first
question can make sense, instead of being at best
a successful recipe to deal with divergences in
perturbative
calculations.[34]
Namely, to answer the second question, one designs
a calculation (see below) that identifies the
required set of parameters to define the theory,
the starting point being some specific initial
assumption, such as some local Lagrangian density
which is a function of field variables and needs
to be modified by including all the required
parameters. Once the required set of parameters
is known, one can establish a set of observables
that are sufficient to define the cutoff dependence
of the required set. The observables can have any
finite scale , and one can use any scale
to define the parameters
, up to their
finite parts that must be fitted to experiment,
including features such as the observed symmetries.
Thus, not only the possibility that a renormalization group of the first type may exist can be understood, but also the alternative situations are found where the set of required cutoff dependent parameters does not have to be finite. Predictive power of latter theories results from known relationships among the required parameters and options to establish all the relevant ones.[35]
The renormalization group concept of the second
kind is associated with the nature of the mathematical
computation used to discover the set of parameters
. In its essence, the calculation starts with
some specific form of a theory with cutoff
and derives a corresponding theory
with a smaller cutoff, in the sense of more
restrictive, say
. After re-parameterization
using the cutoff as a unit, one obtains a new
theory of similar type but with new terms. This
means that the starting theory with cutoff
should also contain such new terms for its form to
be consistent with the presence of a cutoff. Eventually,
one can find a set of terms that reproduces itself up
to changes in the coefficients of the required terms.
These coefficients evolve with the number of steps one
makes, in each and every step reducing the cutoff by
factor of two and rescaling variables. One could use
other factors than two, but two is convenient.
In summary, one obtains a trajectory of a point
in a space of dimension equal to the number of
required parameters and motion along the trajectory
is described by transformations that form new
kind of a group. Different initial points might
lead to different trajectories, but if the
steps are self-similar and reduce to a multiple
action of one and the same transformation, say
, one may describe what happens in terms of
the features of
, called the renormalization
group transformation. The transformation
may transform points in the parameter space
making some of the parameters decrease, some grow, and some
stay unchanged. It may have fixed points,
limit cycles, or even lead to chaotic motion.
Suppose that has a fixed point. If one starts
the procedure at this point, an infinitely long
sequence of reductions of the cutoff by factors of
two changes nothing in the structure of the theory,
except the scale of its cutoff. This means that
the initial cutoff can be arbitrarily large. Such a
theory may possess the symmetries of special
relativity, since there is no price to pay for
extending the cutoff as required when one wishes
to make the Lorentz transformation that yields
momenta which exceed the cutoff.
Both concepts of the renormalization group can
be considered in quantum theories constructed
using the front form of dynamics. The first
concept allows one to play with a small set
of parameters and seek consistency, which
is a useful strategy in perturbation theory if
one knows from other approaches what to expect.
In particular, one may study new perturbative
features that appear in the front form of dynamics,
since it differs from the instant form. The main
difference is that the front variables
(or
) are considerably different from the
transverse variables
(or
),
so that there is no simple rotational symmetry
among them.
One can also study sufficiently simplified models for which computers can be used to carry out calculations and see if a procedure suggested by perturbation theory may work beyond it. The second concept allows one to address the issue of defining a relativistic theory ab initio without limiting the definition to perturbative expansions. This option is particularly relevant to the issue of describing bound states in QCD. However, to address this issue one needs to overcome certain difficulties that the renormalization group procedures based on the idea of reduction of cutoffs are not capable of easily resolving. To avoid the difficulties, one can employ the similarity renormalization group procedure. Both the difficulties and similarity are explained in the next section.
Similarity transformations
A glimpse of the difficulties of the procedure of reducing a
cutoff to cutoff
in the front form
of Hamiltonian dynamics of strong interactions can be
gained by considering the eigenvalue problem for the
Hamiltonian
,
where ,
has a known spectrum and
describes the interactions. Let us assume that
the eigenstate
can be written as a superposition
of eigenstates of
and let us introduce two
projection operators,
and
, such that
projects
on eigenstates of
with eigenvalues smaller than
and
projects on eigenstates of
with
eigenvalues between
and
. The result
of projecting the eigenvalue problem for
using
and
is a set of two coupled equations
The first equation can be used to evaluate
in terms of
,
This expression allows one to write an
equation for in the form
where
The equation for appears to resemble
an eigenvalue problem for
. It is
valid in a theory with cutoff
, but
its effective ``Hamiltonian
depends
on the unknown eigenvalue
. However, if
is much greater than
of interest,
one can neglect
in comparison to
provided that
is small in comparison
to
.
In QCD, which is asymptotically free, one indeed has
as the dominant term in the energy denominator
in
for small eigenvalues
. In practice,
this happens for cutoffs
so much larger
than the smallest eigenvalues
of physical interest
that the corresponding eigenvalue problems are too
complex for solving them with required precision.
Namely, there are still too many degrees of freedom.
One needs to reduce cutoffs
considerably further. This issue appears in all approaches to
the bound state problem in QCD, not only in the front
form of the dynamics.
Even if interactions are sufficiently small, one faces
an additional difficulty with eliminating -states.
Namely, for small interactions one can eliminate the
eigenvalue
from a proper effective Hamiltonian in
-subspace in favor of eigenvalues of
. Consequently,
the denominators analogous to the one that appears
above in
only contain differences of
eigenvalues of
, one above
and one
below.[30][31]
Unfortunately, such
differences can become arbitrarily small near the
cutoff
, and they generate strong interactions
in the effective theory due to the coupling between the
states just below and just above the cutoff
.
This is particularly bothersome when the eigenstates
of
near the cutoff are highly degenerate and
splitting of the bound state problem into parts
below and above the cutoff cannot be accomplished
through any simple expansion in powers of the coupling
constant.
In any case, when one reduces the cutoff to
, and then
to
and so on, the strength of interaction in QCD
Hamiltonians increases and, especially if the
interaction is attractive,
can cancel
and
cannot be ignored no matter how small
it is in comparison to the reduced cutoff. In particular, this
difficulty concerns bound states, where interactions
must prevent free relative motion of constituents
from dominating the scene and a spatially compact
systems have to be formed. So far, it appears
not possible to precisely eliminate the eigenvalue
from the effective dynamics obtained by projecting
on sufficiently low energy eigenstates of
to facilitate reliable calculations.
Fortunately, one can use instead a change of
basis.[36]
Namely, it is possible to define
a procedure in which the basis states are rotated in
such a way that the matrix elements of vanish
between basis states that according to
differ
in energy by more than a running cutoff, say
.
The running cutoff is called the energy bandwidth.
The name comes from the band-diagonal form of the
Hamiltonian matrix in the new basis ordered in energy
using
. Different values of the running cutoff
correspond to using differently rotated
basis states. The rotation is designed not to depend
at all on the eigenvalues
one wants to compute.
As a result, one obtains in the rotated basis an
effective Hamiltonian matrix eigenvalue problem
in which the dependence on cutoff may
manifest itself only in the explicit dependence
of matrix elements of the new
.[36]
The two features of similarity that
(1) the
-dependence becomes explicit
before one tackles the problem of solving
the eigenvalue problem for
and
(2) the effective Hamiltonian with small energy
bandwidth may not depend on the eigenvalues
one tries to find,
allow one to discover in advance the required
counterterms to the diverging cutoff dependence.
A complete set of counterterms defines the set of
parameters required for defining the theory which
has a finite energy bandwidth
and no
cutoff dependence in the band. In the course of
discovering the counterterms and corresponding
parameters, one keeps changing the initial
Hamiltonian. Eventually, the complete Hamiltonian
may have cutoff independent eigenvalues, including
bound states.
In the case of the front-form Hamiltonian for QCD, a perturbative version of the similarity renormalization group procedure is outlined by Wilson et al.[37] Further discussion of computational methods stemming from the similarity renormalization group concept is provided in the next section.
Renormalization group procedure for effective particles
The similarity renormalization group procedure, discussed in #Similarity transformations, can be applied to the problem of describing bound states of quarks and gluons using QCD according to the general computational scheme outlined by Wilson et al.[37] and illustrated in a numerically soluble model by Glazek and Wilson.[38] Since these works were completed, the method has been applied to various physical systems using a weak-coupling expansion. More recently, similarity has evolved into a computational tool called the renormalization group procedure for effective particles, or RGPEP. In principle, the RGPEP is now defined without a need to refer to some perturbative expansion. The most recent explanation of the RGPEP is given by Glazek in terms of an elementary and exactly solvable model for relativistic fermions that interact through a mass mixing term of arbitrary strength in their Hamiltonian.[39][40]
The effective particles can be seen as resulting
from a dynamical transformation akin to the
Melosh transformation from current to constituent
quarks.[41]
Namely, the RGPEP transformation
changes the bare quanta in a canonical theory
to the effective quanta in an equivalent effective theory
with a Hamiltonian that has the energy
bandwidth ;
see #Similarity transformations
and references therein for an explanation of the band.
The transformations that
change
form a group.
The effective particles are introduced through a transformation
where is a quantum field operator built
from creation and annihilation operators for
effective particles of size
and
is the original quantum field
operator built from creation and annihilation
operators for point-like bare quanta of a
canonical theory. In great brevity, a canonical
Hamiltonian density is built from fields
and the effective Hamiltonian at scale
is
built from fields
, but without actually
changing the Hamiltonian. Thus,
which means that the same dynamics is expressed
in terms of different operators for different
values of . The coefficients
in the
expansion of a Hamiltonian in powers of the
field operators
depend on
and the
field operators depend on
, but the Hamiltonian
is not changing with
. The RGPEP provides an
equation for the coefficients
as functions
of
.
In principle, if one had solved the RGPEP
equation for the front form Hamiltonian of QCD
exactly, the eigenvalue problem could be written
using effective quarks and gluons corresponding
to any . In particular, for
very small,
the eigenvalue problem would involve very large
numbers of virtual constituents capable of
interacting with large momentum transfers up to
about the bandwidth
. In
contrast, the same eigenvalue problem written
in terms of quanta corresponding to a large
,
comparable with the size of hadrons, is hoped to
take the form of a simple equation that resembles
the constituent quark models. To demonstrate
mathematically that this is precisely what happens
in the RGPEP in QCD is a serious challenge.
Bethe-Salpeter equation
The Bethe-Salpeter amplitude, which satisfies the
Bethe-Salpeter equation[42]
[43]
[44]
(see the reviews by
Nakanishi[45]
[46]
),
when projected on the light-front plane, results in the light-front wave function. The
meaning of the ``light-front projection" is the following.
In the coordinate space, the Bethe-Salpeter amplitude is a function of
two four-dimensional coordinates ,
namely:
, where
is the total four-momentum
of the system. In momentum space, it is given by the Fourier transform:
(the momentum space Bethe-Salpeter amplitude defined in this way
includes in itself the delta-function responsible for the momenta conservation
). The light-front projection means that the arguments
are
on the light-front plane, i.e., they are constrained by the condition (in the
covariant formulation):
. This is achieved by inserting in the Fourier
transform the corresponding delta functions
:
In this way, we can find the light-front wave function .
Applying this formula to the Bethe-Salpeter amplitude with a given total
angular momentum, one reproduces the angular momentum structure of the
light-front wave function described in Light front quantization#Angular momentum.
In particular, projecting the Bethe-Salpeter amplitude corresponding to
a system of two spinless particles with the angular momentum
,
one reproduces the light-front wave function
given in Light front quantization#Angular momentum.
The Bethe-Salpeter amplitude includes the propagators of the external particles,
and, therefore, it is singular.
It can be represented in the form of the Nakanishi
integral[47]
through a non-singular function :
-
(1)
where is the relative four-momentum. The Nakanishi weight
function
is found from an equation and has the properties:
,
.
Projecting the Bethe-Salpeter amplitude (1) on the light-front plane, we get the
following useful representation for the light-front wave function (see the
review by Carbonell and
Karmanov[48]):
It turns out that the masses of a two-body system, found from the
Bethe-Salpeter equation for and from the light-front equation
for
with the
kernel corresponding to the same physical content, say, one-boson
exchange (which, however, in the both approaches have very different
analytical forms)
are very close to each other.
The same is true for the electromagnetic form
factors<ref name="ck_trento_09}.
On the contrary, the masses of a three-body system found in the two approaches are rather different. They become very close to each other after incorporating the three-body forces of relativistic origin.[49] This undoubtedly proves the existence of three-body forces, though the contribution of relativistic origin does not exhaust, of course, all the contributions. The same relativistic dynamics should generate four-body forces, etc. Since in nuclei the small binding energies (relative to the nucleon mass) result from cancellations between the kinetic and potentials energies (which are comparable with nucleon mass, and, hence relativistic), the relativistic effects in nuclei are noticeable. Therefore, many-body forces should be taken into account for fine tuning to experimental data.
Vacuum structure and zero modes
One of the advantages of light-front quantization is that the
empty state, the so-called perturbative vacuum, is the physical
vacuum.[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
The massive states of a theory can then be built
on this lowest state without having any contributions from
vacuum structure, and the wave functions for these massive
states do not contain vacuum contributions. This occurs because
each is positive, and the interactions of the theory
cannot produce particles from the zero-momentum vacuum without
violating momentum conservation. There is no need to normal-order
the light-front vacuum.
However, certain aspects of some theories
are associated with vacuum structure. For example, the
Higgs mechanism of the Standard Model relies on spontaneous
symmetry breaking in the vacuum of the
theory.[61]
[62]
[63]
[64]
[65]
[66]
The usual Higgs vacuum expectation value in the instant form is
replaced by zero mode analogous to a constant Stark field
when one quantizes the Standard model using
the front form.[67]
Chiral symmetry breaking of quantum chromodynamics is
often associated in the instant form with quark and gluon condensates
in the QCD vacuum.
However, these effects become properties of the hadron wave functions
themselves using the front
form.[59]
[60]
[68]
[69]
This also eliminates the many orders of magnitude conflict
between the measured cosmological constant and quantum field
theory.[68]
Some aspects of vacuum structure in light-front quantization can be analyzed by studying properties of massive states. In particular, by studying the appearance of degeneracies among the lowest massive states, one can determine the critical coupling strength associated with spontaneous symmetry breaking. One can also use a limiting process, where the analysis begins in equal-time quantization but arrives in light-front coordinates as the limit of some chosen parameter.[70] [71] A much more direct approach is to include modes of zero longitudinal momentum (zero modes) in a calculation of a nontrivial light-front vacuum built from these modes; the Hamiltonian then contains effective interactions that determine the vacuum structure and provide for zero-mode exchange interactions between constituents of massive states.
See also
- Light front quantization
- Light-front quantization applications
- Quantum field theories
- Quantum chromodynamics
- Quantum electrodynamics
- Light-front holography
References
- 1 2 B. L. G. Bakker, A. Bassetto, S. J. Brodsky, W. Broniowski, S. Dalley, T. Frederico, S. D. Glazek, J. R. Hiller; et al. (2014). "Light-Front Quantum Chromodynamics: A framework for the analysis of hadron physics". Nuclear Physics Proceedings Supplement. 251-252: 165–174. arXiv:1309.6333. Bibcode:2014NuPhS.251..165B. doi:10.1016/j.nuclphysbps.2014.05.004.
- 1 2 M. Burkardt (2002). "Light front quantization". Advances in Nuclear Physics 23: 1–74. doi:10.1007/0-306-47067-5_1.
- 1 2 S.J. Brodsky, H.-C. Pauli, and S.S. Pinsky (1998). "Quantum chromodynamics and other field theories on the light cone". Physics Reports 301: 299–486. Bibcode:1998PhR...301..299B. doi:10.1016/S0370-1573(97)00089-6.
- ↑ P. A. M. Dirac (1949). "Forms of Relativistic Dynamics". Reviews of Modern Physics 21: 392–399. Bibcode:1949RvMP...21..392D. doi:10.1103/RevModPhys.21.392.
- 1 2 H.-C. Pauli and S. J. Brodsky (1985). "Solving field theory in one space one time dimension". Physical Review D 32: 1993–2000. Bibcode:1985PhRvD..32.1993P. doi:10.1103/PhysRevD.32.1993.
- 1 2 H.-C. Pauli and S. J. Brodsky (1985). "Discretized light cone quantization: Solution to a field theory in one space one time dimensions". Physical Review D 32: 2001–2013. Bibcode:1985PhRvD..32.2001P. doi:10.1103/PhysRevD.32.2001.
- ↑ T. Maskawa and K. Yamawaki (1976). "The Problem of
Mode in the Null Plane Field Theory and Dirac's Method of Quantization". Progress of Theoretical Physics 56: 270–283. doi:10.1143/PTP.56.270.
- ↑ T. Eller, H.-C. Pauli, and S. J. Brodsky (1987). "Discretized Light Cone Quantization: The Massless and the Massive Schwinger Model". Physical Review D 35: 1493–1507. doi:10.1103/PhysRevD.35.1493.
- ↑ K. Hornbostel, S. J. Brodsky, and H.-C. Pauli title=Light Cone Quantized QCD in (1+1)-Dimensions (1990). "Light-cone-quantized QCD in 1+1 dimensions". Physical Review D 41: 3814–3821. doi:10.1103/PhysRevD.41.3814.
- ↑ A. C. Tang, S. J. Brodsky, and H.-C. Pauli title=Discretized light cone quantization: Formalism for quantum electrodynamics (1991). "Discretized light-cone quantization: Formalism for quantum electrodynamics". Physical Review D 44: 1842–1865. doi:10.1103/PhysRevD.44.1842.
- 1 2 Y. Matsumura, N. Sakai, and T. Sakai (1995). "Mass spectra of supersymmetric Yang-Mills theories in (1+1)-dimensions". Physical Review D 52: 2446–2461. doi:10.1103/PhysRevD.52.2446.
- 1 2 O. Lunin and S. Pinsky (1999). "SDLCQ: Supersymmetric discrete light cone quantization". AIP Conference Proceedings 494: 140. doi:10.1063/1.1301663.
- 1 2 J. R. Hiller, S. S. Pinsky, N. Salwen and U. Trittmann (2005). "Direct evidence for the Maldacena conjecture for N=(8,8) super Yang-Mills theory in 1+1 dimensions". Physics Letters B 624: 105–114. doi:10.1016/j.physletb.2005.08.003.
- ↑ S.S. Chabysheva and J.R. Hiller (2009). "Zero momentum modes in discrete light-cone quantization". Physical Review D 79: 096012. doi:10.1103/PhysRevD.79.096012.
- 1 2 W. A. Bardeen and R. B. Pearson (1976). "Local Gauge Invariance and the Bound State Nature of Hadrons". Physical Review D 14: 547–551. doi:10.1103/PhysRevD.14.547.
- ↑ M. Burkardt and S. Dalley (2002). "The relativistic bound state problem in QCD: transverse lattice methods". Progress in Particle and Nuclear Physics 48: 317–362. doi:10.1016/S0146-6410(02)00140-0.
- ↑ J. P. Vary, H. Honkanen, J. Li, P. Maris, S. J. Brodsky, A. Harindranath, G. F. de Teramond, and P. Sternberg (2010). "Hamiltonian light-front field theory in a basis function approach". Physical Review C 81: 035205. Bibcode:2010PhRvC..81c5205V. doi:10.1103/PhysRevC.81.035205.
- ↑ G. F. de Teramond and S. J. Brodsky (2005). "Hadronic spectrum of a holographic dual of QCD". Physical Review Letters 94: 201601. Bibcode:2005PhRvL..94t1601D. doi:10.1103/PhysRevLett.94.201601.
- ↑ G. F. de Teramond and S. J. Brodsky (2009). "Light-Front Holography: A First Approximation to QCD". Physical Review Letters 102: 081601. doi:10.1103/PhysRevLett.102.081601. PMID 19257731.
- ↑ S. J. Brodsky, F. -G. Cao and G. F. de Teramond (2012). "AdS/QCD and Applications of Light-Front Holography". Communications in Theoretical Physics 57: 641–664. Bibcode:2012CoTPh..57..641S. doi:10.1088/0253-6102/57/4/21.
- ↑ T. Gutsche, V. E. Lyubovitskij, I. Schmidt and A. Vega (2013). "Nucleon resonances in AdS/QCD". Physical Review D 87: 016017. Bibcode:2013PhRvD..87a6017G. doi:10.1103/PhysRevD.87.016017.
- ↑ T. Gutsche, V. E. Lyubovitskij, I. Schmidt and A. Vega (2013). "Chiral Symmetry Breaking and Meson Wave Functions in Soft-Wall AdS/QCD". Physical Review D 87: 056001. Bibcode:2013PhRvD..87e6001G. doi:10.1103/PhysRevD.87.056001.
- ↑ H. Honkanen, P. Maris, J. P. Vary and S. J. Brodsky (2011). "Electron in a transverse harmonic cavity". Physical Review Letters 106: 061603. Bibcode:2011PhRvL.106f1603H. doi:10.1103/PhysRevLett.106.061603.
- ↑ X. Zhao, H. Honkanen, P. Maris, J.P. Vary, and S.J. Brodsky (2012). "Electron Anomalous Magnetic Moment in Basis Light-Front Quantization Approach". Few Body Systems 52: 339–344. Bibcode:2012FBS....52..339Z. doi:10.1007/s00601-011-0273-2.
- ↑ X. Zhao, H. Honkanen, P. Maris, J. P. Vary and S. J. Brodsky. "Electron g-2 in Light-Front Quantization". arXiv:1402.4195.
- ↑ S. S. Chabysheva and J. R. Hiller (2012). "A Light-Front Coupled-Cluster Method for the Nonperturbative Solution of Quantum Field Theories". Physics Letters B 711: 417–422. Bibcode:2012PhLB..711..417C. doi:10.1016/j.physletb.2012.04.032.
- 1 2 R.J. Bartlett and M. Musial (2007). "Coupled-cluster theory in quantum chemistry". Reviews of Modern Physics 79: 291–352. Bibcode:2007RvMP...79..291B. doi:10.1103/RevModPhys.79.291.
- ↑ E. C. G. Stueckelberg and A. Petermann (1953). "Normalization of constants in the quanta theory". Helvetica Physica Acta 26: 499.
- ↑ M. Gell-Mann and F. E. Low (1954). "Quantum electrodynamics at small distances". Physical Review 95: 1300–1312. doi:10.1103/PhysRev.95.1300.
- 1 2 K. G. Wilson (1965). "Model Hamiltonians for Local Quantum Field Theory". Physical Review 140: B445. doi:10.1103/PhysRev.140.B445.
- 1 2 K. G. Wilson (1970). "A Model Of Coupling Constant Renormalization". Physical Review D 2: 1438–1472. doi:10.1103/PhysRevD.2.1438.
- ↑ G. 't Hooft (1973). "Dimensional regularization and the renormalization group". Nuclear Physics B 61: 455–468. doi:10.1016/0550-3213(73)90376-3.
- ↑ K. G. Wilson and J. B. Kogut (1974). "The Renormalization group and the epsilon expansion". Physics Reports 12: 75–199. doi:10.1016/0370-1573(74)90023-4.
- ↑ P. A. M. Dirac (1965). "Quantum Electrodynamics without Dead Wood". Physical Review 139: B684. doi:10.1103/PhysRev.139.B684.
- ↑ R. J. Perry and K. G. Wilson (1993). "Perturbative renormalizability with an infinite number of relevant and marginal operators". Nuclear Physics B 403: 587–601. doi:10.1016/0550-3213(93)90363-T.
- 1 2 S. D. Glazek and K. G. Wilson (1993). "Renormalization of Hamiltonians". Physical Review D 48: 5863–5872. Bibcode:1993PhRvD..48.5863G. doi:10.1103/PhysRevD.48.5863.
- 1 2 K. G. Wilson, T. S. Walhout, A. Harindranath, W.-M. Zhang, R. J. Perry, S. D. Glazek (1994). "Nonperturbative QCD: A Weak coupling treatment on the light front". Physical Review D 49: 6720–6766. Bibcode:1994PhRvD..49.6720W. doi:10.1103/PhysRevD.49.6720.
- ↑ S. D. Glazek, K. G. Wilson title=Asymptotic freedom and bound states in Hamiltonian dynamics (1998). "Asymptotic freedom and bound states in Hamiltonian dynamics". Physical Review D 57: 3558–3566. Bibcode:1998PhRvD..57.3558G. doi:10.1103/PhysRevD.57.3558.
- ↑ S. D. Glazek (2012). "Perturbative Formulae for Relativistic Interactions of Effective Particles". Acta Physics Polinca B 43: 1843. doi:10.5506/APhysPolB.43.1843.
- ↑ S. D. Glazek (2013). "Fermion mass mixing and vacuum triviality in the renormalization group procedure for effective particles". Physical Review D 87: 125032. Bibcode:2013PhRvD..87l5032G. doi:10.1103/PhysRevD.87.125032.
- ↑ H. J. Melosh (1974). "Quarks: Currents and constituents". Physical Review D 9: 1095–1112. Bibcode:1974PhRvD...9.1095M. doi:10.1103/PhysRevD.9.1095.
- ↑ E.E. Salpeter and H.A. Bethe (1951). "A Relativistic Equation for Bound-State Problems". Physical Review 84: 1232–1242. doi:10.1103/PhysRev.84.1232.
- ↑ J. H. O. Sales, T. Frederico, B. V. Carlson and P. U. Sauer (2000). "Light-front Bethe-Salpeter equation". Physical Review C 61: 044003. doi:10.1103/PhysRevC.61.044003.
- ↑ J. H. Sales, T. Frederico, B. V. Carlson, and P. U. Sauer (2001). "Renormalization of the ladder light-front Bethe-Salpeter equation in the Yukawa model journal=Physical Review C". Physical Review C 63: 064003. doi:10.1103/PhysRevC.63.064003.
- ↑ N. Nakanishi (1969). "A General survey of the theory of the Bethe-Salpeter equation". Progress of Theoretical Physics Supplement 43: 1–81. doi:10.1143/PTPS.43.1.
- ↑ N. Nakanishi (1988). "Review of the Wick-cutkosky Model". Progress of Theoretical Physics Supplement 95: 1–24. doi:10.1143/PTPS.95.1.
- ↑ Nakanishi, N. (1971). Graph Theory and Feynman Integrals. New York: Gordon and Breach.
- ↑ J. Carbonell and V.A. Karmanov (2011). "Solutions of the Bethe-Salpeter equation in Minkowski space and applications to electromagnetic form factors". Few Body Systems 49: 205–222. doi:10.1007/s00601-010-0133-5.
- ↑ V. A. Karmanov and P. Maris (2009). "Manifestation of three-body forces in three-body Bethe-Salpeter and light-front equations". Few Body Systems 46: 95–113. doi:10.1007/s00601-009-0054-3.
- ↑ Y. Nambu and G. Jona-Lasinio (1961). "Dynamical model of elementary particles based on an analogy with auperconductivity". Physical Review 122: 345–358. Bibcode:1961PhRv..122..345N. doi:10.1103/PhysRev.122.345.
- ↑ M. Gell-Mann, R. J. Oakes, and B. Renner (1968). "Behavior of current divergences under SU(3) x SU(3)". Physical Review 175: 2195–2199. Bibcode:1968PhRv..175.2195G. doi:10.1103/PhysRev.175.2195.
- ↑ G. 't Hooft and M. Veltman (1972). "Regularization and renormalization of gauge fields". Nuclear Physics B 44: 189–213. Bibcode:1972NuPhB..44..189T. doi:10.1016/0550-3213(72)90279-9.
- ↑ M. A. Shifman, A.I. Vainshtein, and V. I. Zakharov (1979). "QCD and Resonance Physics: Applications". Nuclear Physics B 147: 448–518. Bibcode:1979NuPhB.147..448S. doi:10.1016/0550-3213(79)90023-3.
- ↑ R. P. Feynman (1981). "The Qualitative Behavior of Yang-Mills Theory in (2+1)-Dimensions". Nuclear Physics B 188: 479–512. Bibcode:1981NuPhB.188..479F. doi:10.1016/0550-3213(81)90005-5.
- ↑ E. Witten (1981). "Dynamical Breaking of Supersymmetry". Nuclear Physics B 188: 513–554. Bibcode:1981NuPhB.188..513W. doi:10.1016/0550-3213(81)90006-7.
- ↑ J. Gasser and H. Leutwyler (1984). "Chiral Perturbation Theory to One Loop". Annals of Physics 158: 142–210. Bibcode:1984AnPhy.158..142G. doi:10.1016/0003-4916(84)90242-2.
- ↑ S. D. Glazek (1988). "Light Front QCD in the Vacuum Background". Physical Review D 38: 3277–3286. Bibcode:1988PhRvD..38.3277G. doi:10.1103/PhysRevD.38.3277.
- ↑ P. Maris, C. D. Roberts, and P. C. Tandy (1998). "Pion mass and decay constant". Physics Letters B 420: 267–273. Bibcode:1998PhLB..420..267M. doi:10.1016/S0370-2693(97)01535-9.
- 1 2 S. J. Brodsky, C. D. Roberts, R. Shrock and P. C. Tandy (2012). "Confinement contains condensates". Physical Review C 85: 065202. Bibcode:2012PhRvC..85f5202B. doi:10.1103/PhysRevC.85.065202.
- 1 2 A. Casher and L. Susskind (1974). "Chiral magnetism (or magnetohadrochironics)". Physical Review D 9: 436–460. Bibcode:1974PhRvD...9..436C. doi:10.1103/PhysRevD.9.436.
- ↑ C. M. Bender, S. S. Pinsky, and B. van de Sande (1993). "Spontaneous symmetry breaking of
in (1+1)-dimensions in light front field theory". Physical Review D 48: 816–821. Bibcode:1993PhRvD..48..816B. doi:10.1103/PhysRevD.48.816.
- ↑ S. S. Pinsky and B. van de Sande (1994). "Spontaneous symmetry breaking of (1+1)-dimensional
theory in light front field theory. 2". Physical Review D 49: 2001–2013. Bibcode:1994PhRvD..49.2001P. doi:10.1103/PhysRevD.49.2001.
- ↑ S. S. Pinsky, B. van de Sande, and J.R. Hiller (1995). "Spontaneous symmetry breaking of (1+1)-dimensional
theory in light front field theory. 3". Physical Review D 51: 726–733. Bibcode:1995PhRvD..51..726P. doi:10.1103/PhysRevD.51.726.
- ↑ J. S. Rozowsky and C. B. Thorn (2000). "Spontaneous symmetry breaking at infinite momentum without P+ zero modes". Physical Review Letters 85: 1614–1617. Bibcode:2000PhRvL..85.1614R. doi:10.1103/PhysRevLett.85.1614.
- ↑ D. Chakrabarti, A. Harindranath, L. Martinovic, G. B. Pivovarov and J. P. Vary (2005). "Ab initio results for the broken phase of scalar light front field theory". Physics Letters B 617: 92–98. Bibcode:2005PhLB..617...92C. doi:10.1016/j.physletb.2005.05.012.
- ↑ V. T. Kim, G. B. Pivovarov and J. P. Vary (2004). "Phase transition in light front
". Physical Review D 69: 085008. Bibcode:2004PhRvD..69h5008K. doi:10.1103/PhysRevD.69.085008.
- ↑ P. P. Srivastava and S. J. Brodsky (2002). "A Unitary and renormalizable theory of the standard model in ghost free light cone gauge". Physical Review D 66: 045019. doi:10.1103/PhysRevD.66.045019.
- 1 2 S. J. Brodsky and R. Shrock (2011). "Condensates in Quantum Chromodynamics and the Cosmological Constant". Proceedings of the National Academy of Sciences 108: 45–50. doi:10.1073/pnas.1010113107.
- ↑ S. J. Brodsky, C. D. Roberts, R. Shrock and P. C. Tandy (2010). "Essence of the vacuum quark condensate". Physical Review C 82: 022201. doi:10.1103/PhysRevC.82.022201.
- ↑ K. Hornbostel (1992). "Nontrivial vacua from equal time to the light cone". Physical Review D 45: 3781–3801. doi:10.1103/PhysRevD.45.3781.
- ↑ C.-R. Ji and A. Suzuki (2013). "Interpolating scattering amplitudes between the instant form and the front form of relativistic dynamics". Physical Review D 87: 065015. doi:10.1103/PhysRevD.87.065015.
External links
- ILCAC, Inc., the International Light-Cone Advisory Committee.
- Publications on light-front dynamics, maintained by A. Harindranath.