You are on page 1of 18

Theoretical and Mathematical Physics, 132(2): 1119–1136 (2002)

RELATIVISTIC QUARK MODELS IN THE QUASIPOTENTIAL


APPROACH
V. A. Matveev,∗ V. I. Savrin,† A. N. Sissakian,‡ and A. N. Tavkhelidze∗

We describe the interaction of relativistic particles in the framework of the quasipotential approach. The
description is based on the so-called covariant single-time formulation of quantum field theory in which
the theory is considered on a spacelike three-dimensional hypersurface in the Minkowski space. Special
attention is paid to methods for constructing various quasipotentials as well as to use of the quasipotential
approach to describe the characteristics of relativistic particle interactions in quark models such as the
amplitudes of hadron elastic scattering, the mass spectra and widths of meson decays, and the cross
sections of deep inelastic scattering of leptons on hadrons.

Keywords: quasipotential equation, relativistic quark models, hadron elastic scattering, form factors and
meson spectra, structure functions of hadrons

1. Introduction

Numerous experiments show that the majority of known particles admit an internal structure, i.e., are
composite objects. This presumably pertains to hadrons, which are composed from colored quarks and
gluons. But “exotic” particles such as leptoquarks and excited leptons, which may also have a composite
nature, are now being intensely sought experimentally.
Describing the spectra, decay probabilities, and scattering cross sections of composite objects requires
constructing a consistent theory of bound states. Such a theory must be based on the main principles of
local quantum field theory and must use its apparatus [1]. But direct calculation of the characteristics
of composite objects is hardly possible in the framework of a local quantum field theory because the only
known calculation method in this theory is based on the perturbation theory, while the nature of interacting
particle bound-state production must certainly be nonperturbative.
The most advanced method for calculations outside the perturbation theory when constructing a bound-
state theory is the use of dynamic equations. Indeed, if we are only able to construct kernels (potentials) of
dynamic equations in the lowest orders of a perturbation theory, we can develop methods for finding their
exact or approximate solution that do not use the perturbation theory and allow taking contributions of
nonperturbative interaction effects into account when calculating the observable characteristics of bound
states. In the nonrelativistic case, such a theory can be constructed using the dynamic Schrödinger equation
in the language of the classical potential. But for large mass defects and high energies intrinsic to accelerated
composite particles, the corresponding theory must be essentially relativistic. This is why a way to solve
this problem was proposed half a century ago based on using dynamic equations in a local quantum field
theory such as the Bethe–Salpeter equation [2], the quasipotential equation [3], and others [4].

∗ Institute for Nuclear Research, RAS, Moscow, Russia.


† Skobeltsyn Institute of Nuclear Physics, Moscow State University, Moscow, Russia.
‡ Joint Institute for Nuclear Research, Dubna, Moscow Oblast, Russia.

Translated from Teoreticheskaya i Matematicheskaya Fizika, Vol. 132, No. 2, pp. 267–287, August, 2002. Orig-
inal article submitted February 26, 2002.

0040-5779/02/1322-1119$27.00 
c 2002 Plenum Publishing Corporation 1119
In this paper, we use the quasipotential approach to describe relativistic particle interaction, which is
based on the so-called covariant single-time formulation of quantum field theory [5] in which this theory is
constructed on a spacelike three-dimensional hypersurface in the Minkowski space. The main argument for
using such a formulation is that to calculate observable characteristics of an interaction process in a particle
system, we do not need to describe such a process in the total four-dimensional Minkowski space. It suffices
to observe the system evolution on the “section,” e.g., on a spacelike hyperplane. In this case, using the
Lorentz transformations (hyperplane rotations), we can ensure that the times of all particles become equal.
Because, in contrast to the four-dimensional Bethe–Salpeter formalism, we can exclude relative times
of different particles in the system from consideration in the quasipotential approach, we do not need to
formulate boundary conditions for these variables, which is practically impossible to do starting from any
physical considerations. Moreover, the presence of a unique time when describing a particle system allows
using the probability interpretation to describe the quasipotential wave function of this system.
Therefore, the quasipotential approach for describing relativistic composite systems, although based on
the principles and methods of local quantum field theory, allows describing systems of interacting particles
in a common language of a wave function satisfying a relativistic three-dimensional equation analogous
to the Schrödinger equation with physical boundary conditions. The latter are actually determined by
introducing the procedures for projecting and smoothing on the chosen spacelike hypersurface.
The main problem in the quasipotential approach is choosing the interaction quasipotential. In this
paper, we therefore pay special attention to methods for constructing various quasipotentials as well as to
uses of the quasipotential approach to describe the characteristics of relativistic particle interactions in quark
models, such as the amplitudes of hadron elastic scattering, the mass spectra and widths of meson decays,
and the cross sections of deep inelastic scattering of leptons on hadrons. The explicit analogy between
the relativistic system description using a three-dimensional quasipotential equation and the nonrelativistic
interaction picture thus turns out to be very useful because it permits using purely empirical considerations
of classical physics when constructing the interaction quasipotential.

2. Quasipotential equation and properties of the quasipotential

We now consider the Bethe–Salpeter wave function for the fermion–antifermion system [2]

 
ΦP (x1 , x2 ) = 0|T ψ1 (x1 )ψ̄2 (x2 )S |P , (1)

where |P  is a state vector from the Hilbert space of asymptotic states with a definite particle number and
the total four-momentum P . We note that this state can be (a) a one-particle state when we consider a wave
function of a bound state, (b) a two-particle state when we consider a wave function corresponding to a
scattering process, and (c) a multiparticle state when wave function (1) corresponds to a more complicated
inelastic scattering.
The quasipotential approach, or the single-time formulation of the quantum field theory, is based on the
procedure of smoothing and projecting wave function (1) using one-particle wave functions of the fermion
and the antifermion u(∓) (x; k) on the hyperplane λx = τ in the Minkowski space with a timelike unit
normal four-vector λ (λ2 = 1). Applying this procedure to wave function (1), we obtain the momentum
space wave function depending on the single invariant time τ ,

 
(−) (+) (+)
ΦP (k1 , k2 |τ ) = d x1 δ(λx1 − τ )
4
d4 x2 δ(λx2 − τ )ū1 (k1 ; x1 )λ̂ΦP (x1 , x2 )λ̂u2 (x2 ; k2 ) =

 
 (−) (k1 ),
= (2π)3 2k20 δ (3) k1 + k2 − P − (εk1 + εk2 − εp )λ exp{i(εk1 + εk2 − εp )τ }Φ (2)
P

1120
where εp = (λp) is the invariant energy and we introduce the stationary wave function


 (k1 ) =
2εk2 Φ
(−) (+) (+)
d4 x δ(λx)eik1 x v̄1 (k1 )λ̂ΦP (x, 0)λ̂v2 (k2 ), (3)
P

which includes the one-particle wave function amplitudes v (∓) (k).


Because of the connection between three-dimensional momenta, which are imposed by the δ-function
in (2), and because of the properties of the vector λ, these four-momenta satisfy the relation

k1 + k2 − (εk1 + εk2 )λ = P − εp λ, (4)

i.e., not the total four-momentum of the system but only its component orthogonal to λ is conserved. An-
other feature of wave function (2) is that because fermion and antifermion wave functions satisfy the free
Dirac equation, the particles always lie on the mass hyperboloids k12 = m21 and k22 = m22 . The described
procedure for smoothing and projecting therefore ensures the presence of a single time for the whole system
and excludes processes of virtual particle creation. These properties permit using the probability interpre-
tation for the wave function of a system of interacting relativistic particles and describing such a system in
a purely three-dimensional way.
Wave function (3) can describe the fermion–antifermion bound state both with spin 1 and with spin 0.

We restrict ourselves to only considering the pseudoscalar state with the mass M = P 2 for which the
wave function is
(+)5 (+)
 (−) (k1 ) = v̄1 (k1 )γ v2 (k2 ) ϕP (k1 ),
Φ (5)
P
2εk2
where ϕP (k1 ) is the scalar part of the wave function, which satisfies the quasipotential equation [5] in the
case where m1 = m2 = m,

1
2εk1 (2εk1 − M )ϕP (k1 ) = d3 ωk1 (2εk1 εk1 − m2 )V (k1 ; k1 |P, M )ϕP (k1 ) (6)
m2

with the wave function normalization condition



 2
M= d3 ωk1 2εk1 ϕP (k1 ) +
 
1 ∗ ∂
+ 2 3
d ωk1 d3 ωk1 (2εk1 εk1 − m2 )ϕP (k1 ) V (k1 ; k1 |P, M )ϕP (k1 ), (7)
m ∂M

where d3 ωk = d3 k/(2π)3 2k 0 is the invariant three-dimensional volume element in the momentum space

on the mass hyperboloid. We set λ = P/ P 2 in formulas (6) and (7), i.e., direct the vector λ along
the total system four-momentum. This corresponds to the so-called Markov–Yukawa gauge, which fixes
the arbitrariness in choosing the direction of λ in a physically meaningful way because the only selected
direction in the Minkowski space is the total four-momentum of the system under consideration. In this
case, we have
k1 + k2 − (εk1 + εk2 )λ = k1 + k2 − (εk1 + εk2 )λ = 0, (8)

whereas in the case with equal fermion masses, we have εk1 = εk2 and εk1 = εk2 .
 √
The kernel of three-dimensional dynamic equation (6) (the quasipotential V k1 ; k1 |P, s , s = P 2 ) is
the relativistic generalization of the quantum mechanical potential, but the quasipotential differs from the
potential because it is complex, is nonlocal, and depends on the system energy. This reflects the purely

1121
relativistic effects of particle creation and interaction retarding. In particular, the necessity of taking the
energy dependence of the quasipotential into account can be clearly seen in normalization condition (7),
which explicitly contains the quasipotential derivative w.r.t. the total energy. The important quasipotential
property, which follows from the unitarity condition for the S-matrix, is the nonnegativity of its mean
imaginary part [6],

 
1 ∗  √ 1
d3 ωk1 d3 ωk1 (2εk1 εk1 − m2 )ϕP (k1 ) Im V k1 ; k1 |P, s ϕP (k1 ) = H(p1 ; p1 |P ) ≥ 0, (9)
m2 2
where ϕP (k1 ) is the wave function corresponding to the elastic fermion–antifermion scattering, P = p1 + p2 ,
and H(p1 ; p1 |P ) is the diagonal contribution of all possible inelastic channels in the unitarity condition.
In addition to the sign-definiteness, formula (9) also implies that the presence of the imaginary part of the
quasipotential reflects the presence of the particle creation and annihilation processes in relativistic physics.

3. Electromagnetic interaction quasipotential


To use integral equation (6) and normalization condition (7) to find wave eigenfunctions and energies,
we must know the explicit form of the quasipotential. There are various methods for constructing the
interaction quasipotential. For example, in the weak interaction case where we can apply the perturbation
theory, the regular method for constructing a local quasipotential was found based on the two-time Green’s
function method and dispersion relations for scattering amplitudes in the momentum transfer [7]. This
method for constructing a local quasipotential was successfully used to solve a series of electrodynamic
problems, in particular, to investigate energy levels of positronium and hydrogen-like atoms [8], and to
study the convergence of coupling-constant expansions in nonrenormalizable theories [9].
It was shown [8], [10] that the quasipotential of the electromagnetic interaction of electrons and
positrons in the pseudoscalar state in the lowest approximation of the one-photon exchange can be written
as
 √ 4πα(2m)2
V k1 ; k1 |P, s =  √ , (10)
q q − s + εk1 + εk1 − i0

where q = (εk1 − εk1 )2 − (k1 − k1 )2 . In the rest frame of reference, we obtain

 √  4πα(2m)2
V k; k |P, s P=0 =  √ , (11)
|k − k ||k − k | − s + k 0 + k  0 − i0

where k 0 = k2 + m2 . This quasipotential is the relativistic generalization of the Coulomb potential
obtained in the lowest order of the perturbation theory in the quantum electrodynamic framework.
We now turn to the properties of quasipotential (11), which distinguish it from the standard Coulomb
potential used in classical physics or in the nonrelativistic quantum mechanics for describing charged particle
interaction. First, quasipotential (11) is nonlocal, i.e., it depends not only on the momentum difference
k − k but also on the momenta themselves. Second, we see that already in the lowest approximation,

quasipotential (11) has a domain where its imaginary part is nonzero for s > 2m,
 √   √ 2 
Im V k; k |P, s P=0 = (4πm)2 αδ (k − k )2 − s − k0 − k 0 . (12)

Eventually, quasipotential (11) depends explicitly on the energy s of the system under consideration.
This last property drastically changes the formulation of the eigenvalue problem for integral equation (6)
because the binding energy E = M − 2m enters the kernel of this equation explicitly in this case.

1122
It was shown in [10] that at sufficiently small particle momenta, quasipotential (11) can be set approx-
imately local and can be investigated in the configuration space of relative distances between particles. We
then find that the quasipotential oscillates for E > 0, while the oscillation amplitude decrease is governed by
the Coulomb law and the oscillation frequency is equal to E/(c), i.e., vanishes in the nonrelativistic limit
c → ∞. When the value of E is negligibly small, the quasipotential becomes the standard nonrelativistic
Coulomb potential.
We also note that the picture of scattering on oscillating quasipotentials, which also depend on the
energy, can differ substantially from the customary picture of potential scattering, which is manifested, in
particular, in the appearance of discrete levels embedded in the continuous spectrum [11].

4. Relativistic configuration space


In the rest frame of reference for the pseudoscalar bound state P = 0, quasipotential equation (6)
becomes

1
2k 0 (2k 0 − M )ϕM (k) = 2 d3 ωk (2k 0 k  0 − m2 )V (k; k |P, M )ϕM (k ). (13)
m
Passing to the configuration space using the standard three-dimensional Fourier transformation for this
relativistic equation becomes obviously senseless because the momentum space in this case is non-Euclidean
(it is the mass hyperboloid). This is why we need the functions that realize the Lorentz group representation
as the group of motions in the Lobachevsky space on the particle mass hyperboloid for describing the
corresponding transformation in the relativistic case. These functions, found in [12], are

−1−imr
kn
ξ(r; k) = , (14)
m

where n = (1, r/r) is a lightlike four-vector. These functions have the necessary properties of completeness
and orthogonality. Expanding the quasipotential over these functions, we obtain

 
 √ ∗  √
V k; k |P, s = −(2m)2 d r3
d3 r ξ(k; r)V r; r | s ξ(r ; k ). (15)

If we now introduce the local quasipotential

 √  √
V r; r | s = V r| s δ (3) (r − r ), (16)
 √
where the function V r| s is the analogue of the local spherically symmetric potential in the relativistic

configuration space and depends on the total state energy s , then substituting (16) in formula (15), we
obtain

 √  √ ∗  √
V k; k |P, s = V q| s = −(2m)2 d3 r ξ(q; r)V r| s , (17)

where

k (kk )
q = k(−)k ≡ k − k0 −  0 (18)
m k +m
is the difference of two vectors in the curved Lobachevsky space realized on the particle mass hyperboloid.
The customary squared four-momentum transfer vector is then


t ≡ (k − k  )2 = −2m q2 + m2 − m .

1123
To write quasipotential equation (13) in the relativistic configuration space, we must construct the
operator of the free Hamiltonian satisfying the equation

 (0) rξ(r; k) = 2k 0 rξ(r; k)


H (19)
r

in this space. This operator, found in [12], has the form in the spherical coordinates

 2

r(0) = 2m cosh iλ ∂ + λ ∆θ,ϕ exp iλ ∂
H , (20)
∂r 2r(2) ∂r

where r(2) = r(r + iλ) and λ = m−1 is the fermion Compton wave length, i.e., the operator is of the finite-
difference type. As a result, the quasipotential equation in the relativistic configuration space becomes

  √ (0)  √
H  (0) rϕM (r) = 1 H
 (0) M − H  (0) V r| s H
 rϕM (r) − 2mV r| s rϕM (r), (21)
r r r r
m
where

ϕM (r) = 2m d3 ωk ξ(r; k)ϕM (k). (22)

It is easy to show that the nonrelativistic limit λ → 0 of Eq. (21) yields the differential Schrödinger equation.
It is clear that the finite-difference operators appear in Eq. (21) because of purely relativistic effects—
the dispersion law (the relation between particle energy and momentum) and the nonlocality of the kernel
of quasipotential equation (13). Some time ago, the necessity of introducing finite-difference operations
resulted in the heuristic idea of quantizing the relativistic configuration space. But this idea has not been
developed in depth.
We can expand wave function (22) over partial waves in the coordinate space,


 (l)
rϕM (r) = (2l + 1)ϕM (r)Pl (cos θr ), (23)
l=0

as well as in the momentum space,


 (l)
kϕM (k) = 4π (2l + 1)Pl (cos θk )ϕM (χ). (24)
l=0

Here, we introduce the rapidity variable χ = log(k 0 + k)/m, which is convenient for parameterizing the
particle energy and momentum on the mass hyperboloid because k 0 = m cosh χ and k = m sinh χ. Moreover,
this variable χ, not the momentum k, is conjugate to the relativistic coordinate r, which most explicitly
follows from the relation between the s-state partial wave functions,

 ∞
(0) (0)
ϕM (χ) = dr sin(mrχ)ϕM (r). (25)
0

5. Strong interaction quasipotential and elastic hadron scattering


In Sec. 3, we investigated charged particle interaction quasipotential (10) constructed in the lowest
order of the perturbation theory in the framework of quantum electrodynamics. But such a construction
is possible only in those cases where the coupling constant is small. Such an approach becomes completely

1124
incorrect if we want to describe the hadron interaction because we deal with the strong interaction in this
case. The quasipotential approach for describing hadrons scattering at high energies is therefore based on
a phenomenological choice of local two-particle quasipotential (16). We must then start from the general
principles of local field theory, such as the relativistic invariance, the analyticity, the unitarity, and the
crossing symmetry. Already in the first papers devoted to the quasipotential approach in quantum field
theory, it was shown how starting from dispersion relations on the momentum transfer, one can construct
a local quasipotential in the form of a continuous superposition of the Yukawa potentials with an energy-
dependent density [7],
 ∞ √
 √ σ(µ2 | s )
V q| s = dµ2 . (26)
4m2 µ2 − t
Quasipotentials of such type were used to study the asymptotic behavior of amplitudes in quantum field
theory. In particular, these quasipotentials were used to study the nature of the Regge asymptotic behavior
of the two-particle scattering amplitude, and the possibility that branching points of partial amplitudes
appear in the complex plane of the angular moment was indicated [9].
In addition to the general restrictions, the experimentally observed empirical rules for two-particle
scattering at high energies must be satisfied when constructing a quasipotential. The hypothesis about the
smoothness of a local quasipotential introduced in [13] turns out to be rather helpful. The smoothness
property of a local quasipotential reflects the important dynamic property of two-particle interactions at
high energies, which means, imaginatively speaking, that hadrons behave themselves as “crumbly” spread
objects of finite sizes when colliding at high energies. This hypothesis explains many principal laws of elastic
scattering of high-energy particles at small and large angles [14]. In particular, the validity of the eikonal
approximation for the scattering amplitude, which correctly reproduces both the experimentally observed
diffraction picture of elastic scattering at small angles and the exponential decrease of differential sections
with the transferred momentum increase, was proved for smoothed potentials.
The quasipotential equation for the elastic scattering amplitude of two identical particles (e.g., protons)
in the center-of-mass frame P = 0 is
  √  √
 √  √ V k; k |P, s T k ; p|P, s
T k; p|P, s = V k; p|P, s + 3
d ωk  √ , (27)
2k  0 2k  0 − s − i0
 √ √
while the amplitude T k; p|P, s is outside the energy shell, i.e., 2k 0 = 2p0 = s . We can seek solutions
of this equation in the domain of high energies and bounded transferred momenta, |t|/s  1, by iterations.
In the first Born approximation, the amplitude obviously coincides with the quasipotential. Because the
observed differential scattering cross section has the form of a diffraction peak in the domain of small
transferred momentum, we can write the local quasipotential in the form (neglecting particle spins)

 √   
V q| s = isgeat = isg exp −2ma q2 + m2 − m . (28)

By virtue of (17), we have



 √ 1  √ isg 2
V r| s = − d3 ωq ξ(r; q)V q| s = − e2m a Kimr (2m2 a) (29)
2m (4π)2 am

in the configuration space. The second iteration of Eq. (27) on the energy shell with the use of expression (28)
now results in
  √  √   
V k; k |P, s V k ; p|P, s exp 2a(k  0 − p0 )2 − 2a(k − l)2
3
d ωk  √ 2
= (isg) e at/2 3
d ωk  √ , (30)
2k  0 2k  0 − s − i0 2k  0 2k  0 − s − i0

1125
where l = (k + p)/2. Integrating over angle variables in the spherical coordinates, we obtain

 ∞
 
(isg)2 at/2   exp 2a(k  0 − p0 )2 − 2a(k  − l)2
e dk k  √ , (31)
(4π)2 al −∞ (2k  0 )2 2k  0 − s − i0
√ √ √
where 2l = s − 4m2 + t = −u and 2p = s − 4m2 . As a result, the imaginary part of the amplitude
in the two-iteration approximation and in the domain of small transferred momenta |t|/s  1 is

 √  g at/2 
Im T k; p|P, s  sg eat − e + ··· . (32)
32πa
We note that a linear growth of the Pomeron Regge trajectory can be attained by choosing the dependence
of the parameter a on the collision energy to be

s
a(s) = a0 + α log . (33)
s0

Expansion (32) implies the applicability conditions

g
 1, a|t| ≤ 1 (34)
32πa
for the second Born approximation. It is clear that the first of these conditions may fail when the total
cross section increases sufficiently rapidly. Then, as before, the solution of quasipotential equation (27) for
the scattering amplitude for |t|/s  1 can be found as an infinite series of Born approximations,

eat/n  g n
∞
 √
Im T k; p|P, s = −8πas − . (35)
n=1
n n! 8πa

This series converges absolutely at any fixed momentum transfer and hence determines an entire analytic
function in the variable t. We note that the sign-alternation of the Born series results in the appearance of
diffraction minimums in the differential cross section observed in proton–proton scattering experiments at
an intermediate momentum transfer.
Summing series (35) at t = 0 and using the optical theorem, we obtain

 g  g 
σtot (s) = 8πa log + C − Ei − . (36)
8πa 8πa
As a result, the asymptotic behavior of the total cross section as g/(8πa) → ∞ is

g
σtot (s)  8πa log . (37)
8πa
When the function g(s) has a power-law increase, we use dependence (33) to find that this asymptotic
behavior coincides with the absolute Froissart bound σtot (s) ∼ log2 (s/s̄).
It is easy to see that as |t| increases, the leading contribution to the scattering amplitude comes from
higher and higher terms of series (35). For a|t|  1, the sum of series (35) can be estimated using the
saddle-point method,
 √ 8πas 
Im T k; p|P, s  − exp − 2a|t| log 2a|t| . (38)
2a|t|
Therefore, the differential cross section at large t becomes less steep, and its oscillations become less distinct.

1126
The so-called eikonal approximation is often used to describe elastic scattering at high energies. It
is based on the semiclassical picture of high-energy scattering when wavelengths of colliding particles are
much less than the effective size of the interaction domain. The eikonal representation for the scattering
amplitude can be easily derived if we seek the solution of the quasipotential equation in the form of iterations
for smooth quasipotentials (35).
Substituting
  
eat/n 1 nρ2
= d ρ exp iρ∆T −
2
(39)
n 4πa 4a

in series (35) (where ρ is the two-dimensional impact vector, t = −∆2T , and ∆T = p − k is changed in the
plane perpendicular to the vector l), we obtain


 √ 2 iρ∆T e2iδ(ρ) − 1
T k; p|P, s = 4s d ρe , (40)
2i

where
ig −ρ2 /(4a)
δ(ρ) = e . (41)
16πa
The phase function δ(ρ) (the eikonal) is related to the quasipotential by


m2 +∞  √
δ(ρ) = dz V ρ2 + z 2 | s , (42)
s −∞

which holds for a wide class of smooth quasipotentials [14], [15].

6. Describing meson spectra

In the quasipotential approach, hadrons (analogously to the nonrelativistic theory) are considered
bound states of relativistic quarks whose interactions are described by quasipotentials that admit space-
localized solutions of the quasipotential equation at some discrete values of the energy. We note that
the customarily used derivation of the quasipotential equation in the quark interaction case can be only
formal because the confinement property dictates that quarks do not admit physical asymptotic states.
But if we consider this derivation as a mathematical procedure, it becomes quite correct and results in
a physical picture analogous to the picture in the nonrelativistic quantum mechanics with an infinitely
growing potential. In other words, we admit the existence of asymptotic states, but in an infinitely distant
domain, whose influence on actual processes is negligibly small.
The nonrelativistic potential theory based on the Schrödinger equation has long been widely used to
describe mesons as quark–antiquark bound states (see, e.g., [16]). But the mass spectra and widths of
meson decay obtained within the quasipotential approach, as we show below, can differ substantially from
the corresponding quantum mechanical quantities, i.e., taking the relativistic effects into account is often
essential.
When describing the bound quarks, we again deal with the strong interaction and cannot use the
perturbation theory to reconstruct the quasipotential. In this case, the quasipotential of the quark–quark
interaction can be constructed somewhat empirically. First, we use the analogy between the quasipotential
and the standard potential and also the similarity between the three-dimensional relativistic interaction de-
scription based on the quasipotential equation and the interaction description in the nonrelativistic quantum
mechanics.

1127
In the strong interaction theory—quantum chromodynamics (QCD)—the interaction between quarks
 √
is due to a massless gluon exchange, and it is therefore natural to assume that the quasipotential V r| s
has the Coulomb form in the relativistic configuration space,

 √ αs
V r| s = − , (43)
r
where the coupling constant is, generally speaking, energy dependent. Substituting this quasipotential in
transformation (17), we obtain

 √ 16παs m
V q| s =   (44)
|q| log 1 + q2 /m2 + |q|/m

in the momentum space [17].


We now consider the asymptotic behavior of the quasipotential in the domain of small distances, or as
Q → ∞,
2

 √ 32παs
V q| s  . (45)
(Q /m ) log(Q2 /m2 )
2 2

We see that the asymptotic behavior of the quasipotential must be then consistent with the known asymp-
totically free behavior of the one-gluon contribution in QCD. Moreover, we assume nothing about the
dependence of αs on the transferred momentum, i.e., from the beginning in this approach, we do not need
to introduce a “running” coupling constant. It turns out that the role of the scale constant well known in
QCD is played by the constituent quark mass here. For simplicity, we do not consider the color degrees of
freedom intrinsic for the QCD here, but we can show that introducing such degrees of freedom does not
alter the qualitative conclusion that the QCD interaction at small distances is reproduced by the standard
Coulomb potential but in the relativistic configuration space.
Calculating the spectrum of bound states in relativistic Coulomb potential (43) using Eq. (21) belongs
to the class of exactly solvable problems [12]. In the s-state case, the mass spectrum of the bound-state
radial excitations is
 
1
Mn = 2m 1 +
2 2
. (46)
1 + α2s /n2
It is obvious that in the case with a small coupling constant αs , the binding energy must be small (Mn →
2m), and spectrum (46) becomes close to the standard nonrelativistic Coulomb spectrum

α2s m
En = − . (47)
4n2
In contrast, relativistic spectrum (46) is essentially non-Coulomb in the case with strong fields. This
spectrum then manifests a linear increase for low-lying levels (n  αs ),

 √ nm
En = − 2 − 2 m + √ , (48)
2 αs

while this spectrum for n  αs is close to nonrelativistic spectrum (47). An important property of this
 √
spectrum is that even in very strong fields, the ground-state binding energy cannot exceed 2 − 2 m. This
is presumably because the creation of new particles from the vacuum must occur above this threshold; hence,
such a state already cannot be described as a two-particle system, which was assumed when formulating
the problem.

1128
Of course, to describe actual spectra of mesons represented as bound states of quarks and antiquarks,
we must introduce a “confining” potential that prevents these particles from escaping. As shown in a
series of papers, the experimental data on mass spectra and widths of meson decay are well described using
combined potentials of the “funnel” type, for instance,

 √ αs
V r| s = − + σr. (49)
r
But it is impossible to obtain an exact solution of three-dimensional quasipotential equation (21) in this
case, and the semiclassical approximation method was used [18]. It is instructive to consider the s-wave
case here. Quasipotential equation (21) for the partial wave with l = 0 can then be written as


∂ ∂ (0) ∂ ∂ (0) (0)
cosh iλ a − cosh iλ ϕM (r) = 2 cosh iλ v(r) cosh iλ ϕM (r) − v(r)ϕM (r), (50)
∂r ∂r ∂r ∂r

where a = M/2m and v(r) = V (r|M )/2m.


(0)
In the semiclassical approximation, we seek the solution of Eq. (21) for the partial wave function ϕM (r)
in the form
 
(0) i
ϕM (r) ∼ exp g(r) , (51)
λ
where g(r) is expanded in the series

2
λ λ
g(r) = g0 (r) + g1 (r) + g2 (r) + . . . . (52)
i i

Substituting this expansion in Eq. (21) and equating the coefficients of equal powers of λ, we obtain the
expressions for the first two terms of the expansion,

g0 (r) = ±χ(r), (53)

1    
g1 (r) = − log sinh χ(r) 2 2v(r) + 1 cosh χ(r) − a , (54)
2
where

 
4v(r) 2v(r) + 1 + a2 + a
cosh χ(r) =   . (55)
2 2v(r) + 1
The classical turning point r+ is obviously determined by the relation cosh χ(r+ ) = 1, which, as can be
easily seen, is equivalent to the equality v(r+ ) = a − 1. We consider the case a > 1 here. Then the
semiclassical solution of the quasipotential equation in the classically accessible domain left of the turning
point is
  r+ 
(0) C 1 π
ϕM (r) =
    sin χ(r) + . (56)
sinh χ(r) 2 2v(r) + 1 cosh χ(r) − a λ r 4

Another turning point r− < r+ is determined by the equality 2v(r− ) + 1 = 0. At this point, cosh χ(r)
becomes infinite, and the wave function must vanish. Hence, the solution to the right of this turning point
must be
  r 
(0) C 1
ϕM (r) =
    sin χ(r) . (57)
sinh χ(r) 2 2v(r) + 1 cosh χ(r) − a λ r−

1129
The semiclassical quantization condition is therefore given by the formula

 r+
3
χ(r) = π n + λ. (58)
r− 4

The turning point r− is of purely relativistic origin, and we have r−  αs λ in the domain of moderate
values of constants in our approximation.
We note that the function χ(r) introduced here has the sense of the rapidity of a particle moving in
the field v(r). We therefore conclude that, as in the momentum space consideration, the description in

terms of rapidities, which are direct generalizations of nonrelativistic momenta p(r) = m a − 1 − v(r) in
the semiclassical approach, is most adequate in the relativistic case.
Using the semiclassical approximation for the quasipotential equation in the relativistic configuration
space, we can obtain satisfactory results for the mass spectra and widths of meson decay (see, e.g., [19]).

7. Form factors of meson decay


In the quasipotential approach, the amplitude of decay of the meson, composed of a quark and antiquark
and described by the wave function Φ  (−) , into two particles can be written as
P
 
M(p1 |P ) = d3 ωk1 T (p1 ; k1 |P, M ) (0) (k1 ; k1 |P, M ) ×
d3 ωk1 G

 (−) −1    (−) 
× 
d3 ωk1 G  (k1 ),
k1 ; k1 |P, M Φ (59)
P

where p1 is the momentum of one decay product particle, G (0) (P, M ) and G
(−) (P, M ) are the respective
Green’s functions of systems of free and interacting quarks and antiquarks, and T (P, M ) is the amplitude of
transition of the quark–antiquark system into the two particles into which the meson decays. If we neglect
the interaction between quarks and antiquarks in the intermediate state (the QCD corrections), then [20]


M(p1 |P ) =  (−) (k1 ).
d3 ωk1 T (p1 ; k1 |P, M )Φ (60)
P

We now consider the process of decay of the pseudoscalar meson into a lepton–antilepton pair, for
instance, π − → eν̄e . Starting from the standard model Lagrangian, we can calculate the amplitude T (P, M )
in the leading order in a weak coupling constant,

G
T (p1 ; k1 |P, M ) = √ v̄e(+) (p1 )γ µ (1 − γ 5 )vν(+) (p2 ) ×
2 (MW − M 2 )
2

(−)
× (MW
2
gµν − Pµ Pν )v̄d (k2 )γ ν (1 − γ 5 )vu(−) (k1 ), (61)

where G/ 2 = g 2 /(8MW
2
). Substituting expressions (61) and (5) in (60), we obtain

G
M(p1 |P ) = √ fπ v̄e(+) (p1 )P(1 − γ 5 )γ 5 vν(+) (p2 ), (62)
2

where the decay constant is expressed through the wave function,


4mq
fπ = d3 ωk1 ϕπM (k1 ), (63)
M

1130
and we set mu = md = mq . It is now easy to calculate the decay width

2
3G2 M m2e 2 m2
Γπ− →eν̄e = fπ 1 − e2 , (64)
8π M

where me is the electron mass and the multiplier 3 is the color factor.
We now consider the decay of the vector meson into an electron–positron pair. For this, we again use
formula (60), but expression (5) for the quasipotential wave function must now be replaced with

(+) (+)
 (−) (k1 ) = v̄q (k1 )êvq (k2 ) ϕV (k1 ),
Φ (65)
P P
2εk2

where eµ is the polarization four-vector of the vector meson satisfying the conditions (P e) = 0 and
∗
ee = −1. Starting from the quantum electrodynamics Lagrangian, we can calculate the amplitude of
the annihilation of the quark–antiquark pair into the electron–positron pair in the leading order of the
perturbation theory,

4παeq (+)
T (p1 ; k1 |P, M ) = − v̄ (p1 )γµ ve(+) (p2 )v̄q(−) (k2 )γ µ vq(−) (k1 ), (66)
M2 e
where eq is the quark charge. Substituting this amplitude and wave function (65) in formula (60), we obtain

M(p1 |P ) = 4παfV v̄e(+) (p1 )êve(+) (p2 ), (67)

where the form factor of the decay process V → e− e+ in the rest frame is

  
2eq (ek1 )2
fV = 3
d k1 1− 2 ϕVM (k1 ). (68)
(2π)3 M 2 k1 + m2q

The decay width is then

1/2
4m2e 2m2e
ΓV →e− e+ = 4πα 2
M fV2 1− 1+ . (69)
M2 M2

We now consider the so-called static limit where we can neglect the internal motion of quarks in the
meson. In this case, formula (68) becomes


2eq 2eq V
fV = d3 k1 ϕVM (k1 ) =  (0),
ϕ (70)
(2π)3 M 2 M 3/2 M

where ϕVM (r) is the wave function normalized by unity in the coordinate space. Substituting this expression
in formula (69) for the decay width, we obtain

16πα2 e2q  V 2
ΓV →e− e+ = M (0) ,
ϕ (71)
M 2

where we neglect the electron mass. This classic formula was first obtained in [21] and then in [22]. If we
√  √
turn to actual vector mesons, then we must set e2q to be (e2u − e2d )/ 2 = 1/ 3 2 for the ρ-meson, e2c = 4/9
for the J/ψ-meson, and e2b = 1/9 for the Υ-meson.

1131
For decay width (64) of the π-meson in the static approximation, we obviously have

3G2 m2e  π 2
Γπ− →eν̄e = M (0) .
ϕ (72)

The quasipotential approach can be applied to another classic decay process π 0 → γγ, whose amplitude
structure in the rest frame is
∗ ∗
M(p1 |M ) = 4παf0 i8ijk ei1 ej2 pk1 , (73)

where p1 is the momentum of one of the photons, e1 and e2 are photon polarization vectors, and the indices
i, j, and k take the values 1, 2, 3. In the quasipotential approach, the form factor of this decay is

 

−1
8mq e2q
f0 = d3 ωk1 (p1 − k1 )2 + m2q (p1 − k1 )2 + m2q + k10 − p01 − i0 ϕπM (k1 ), (74)
(2π)3
 √
where e2q = 1/ 3 2 as in the ρ-meson case. It is now easy to calculate the decay width in the static
approximation,

3πα2 M 2 32πα2  π 2
Γπ0 →γγ = f0 =  M (0) ,
2 ϕ (75)
4 3M 2 (zπ2 + 1) zπ2 + 1 + zπ − 1

where the parameter zπ = 2mq /M .


Now supposing that the interaction of u- and d-quarks is universal, we can assume that the same
wave function ϕ πM (0) enters both formula (72) and formula (75). In this case, excluding this function
from both these formulas and using the experimental data for the decay widths Γπ− →eν̄e and Γπ0 →γγ [23],
we can easily calculate the value of the parameter zπ . It approximately equals six, i.e., we have a huge
mass defect. This means that the quark interaction inside the meson is essentially relativistic, and it is
incorrect to use the static approximation, at least for these lightest mesons. In the relativistic case, we
must therefore know the meson wave functions, which are solutions of the quasipotential equation, and
substituting which in formulas (63), (68), and (74), we can calculate the corresponding decay constants.
Applications of the quasipotential approach to semilepton decays of B-mesons using quasipotential wave
functions can be found, e.g., in [24].

8. Hadron structure functions


The quasipotential approach can be easily generalized to interaction of three and more particles. This
permits using it successfully to describe inclusive processes and hadron structure functions. The quasipo-
tential equation for a system of n particles is [25]
 √
2εkn εk1 + · · · + εkn − s ΨP (k1 , . . . , kn−1 ) =

 √
= d3 ωk1 . . . d3 ωkn−1 V k1 , . . . , kn−1 ; k1 , . . . , kn−1 |P, s ΨP (k1 , . . . , kn−1 ). (76)

It is assumed in the quasipotential approach for describing inclusive processes that a hadron comprises a
large number of partons (quarks), which are initially bound inside it, but interaction results in decay of the
hadron into its constituents. Amplitudes of inclusive reactions can be then expressed through the single-time
wave functions describing hadrons as quark bound states and through amplitudes of subprocesses with the
participation of quarks, gluons, and other particles [26]. The presence of the probability interpretation of

1132
quasipotential wave functions allows using these wave functions to construct the probability densities, which
have the sense of the quark momentum distribution functions inside hadrons, i.e., the sense of structure
functions.
For instance, when considering the process of deep inelastic scattering of a lepton on a hadron, the
emission amplitude for a virtual photon that is afterward absorbed by the lepton can be written analogously
to decay amplitude (60):

M(q, k1 , . . . , kn−1 |P, M ) =




= d3 ωk1 . . . d3 ωkn−1 T q, k1 , . . . , kn−1 ; k1 , . . . , kn−1 |P, M ΨP (k1 , . . . , kn−1 ), (77)

where T (P, M ) is the amplitude of the radiation transition of a system with n constituents with the emission
of a virtual photon with the four-momentum q. As in the case of elastic high-energy scattering, we here
consider the case of identical particles and neglect spins. The amplitude T (P, M ) can be represented as the
sum of the amplitudes of interaction of the virtual photon with each parton. Then, taking only the terms
of coherent interaction into account and using the parton identity, we can write [27] the probability of the
photon emission in the form

w(ν, Q2 ) = 4παne2c (2m)2 ×



 2
× d3 ωk1 . . . d3 ωkn (2π)4 δ (4) (q + k1 + · · · + kn − P )ΨP (k1 , . . . , kn−1 ) , (78)

where ec is the parton charge, ν = εq , and Q2 = −q 2 . Performing the change of variables kn = k − q and
setting (P − k1 − · · · − kn−1 )2 = m2 , we obtain

2
w(ν, Q ) = 4παne2c (2m)2 d3 ωk 2πδ(2kq − Q2 )I(εk ), (79)

where

 2
I(εk ) = d3 ωk1 · · · d3 ωkn−1 (2π)3 2k 0 δ (3) (k1 + · · · + kn−1 + k − P)ΨP (k1 , . . . , kn−1 ) (80)

is the relativistic-invariant function, which is completely determined by the squared quasipotential wave
function of the composite hadron.
We now pass to the rest frame for the hadron P = 0. We can then integrate over spherical angles in
formula (79) obtaining

αne2c (2m)2 ∞ dk k
w(ν, Q2 ) = I(k 0 ), (81)
2ν  kmin k 0

where ν  = ν 2 + Q2 and, neglecting parton masses, we have 2kmin = ν  − ν. The wave function normal-
ization condition yields
 ∞  
1 ∂V (P, M )
dk k 2
I(k 0
) = 2M − . (82)
2π 2 0 ∂M
This expression contains the quasipotential derivative w.r.t. the total energy of the system averaged over
wave functions.
We now introduce the Nachtmann scale variable ξ = (ν  − ν)/M . The integral in formula (81) then
depends on only this variable, which varies within the limits
 
Q2 4M 2
0 ≤ ξ ≤ ξmax = 1+ −1 . (83)
2M 2 Q2

1133
Introducing the hadron structure function F (ξ, Q2 ), we can write probability (81) in the form

8παM
w(ν, Q2 ) = F (ξ, Q2 ), (84)
ν

where the structure function is normalized in a way to ensure the relation F (ξ, Q2 ) = 2πξδ(ξmax − ξ) when
passing to the pointlike hadron. Comparing (84) and (81), we obtain

 kmax
ne2c m2
F (ξ, Q2 ) = dk I(k 0 ), (85)
4πM Mξ/2

where the upper limit of the integration w.r.t. k is introduced because the structure function must vanish
on the threshold of the physical domain and we must set kmax = M ξmax /2. The hadron structure function
therefore depends on only the scale variable ξ if we neglect the weak increase of this function with increasing
Q2 because of changing the upper limit of the integration. We note that in the large-Q2 limit, Q2  4M 2 ,
the variable ξ coincides with the standard Björken variable x = Q2 /2M ν, which ranges exactly the interval
[0, 1].
We now calculate the moments of structure function (85). Changing the integration order, we obtain

 ξmax N  kmax
N −1 2 ne2c m2 2
MN −1 = dξ ξ F (ξ, Q ) = dk k N I(k 0 ). (86)
0 4πN M M 0

In particular, we have
  
πne2c m2 ∂V (P, M )
M1 = 2M − (87)
M3 ∂M
for the first moment by virtue of formula (82).
We note that semi-inclusive distributions, as well as the relation between moments of the semi-inclusive
distribution and structure functions, were studied in detail in [26] in the framework of the quasipotential
approach.

9. Conclusion
In this paper, we considered some consequences of the application of the quasipotential approach to
relativistic quark models of hadrons. Special attention was paid to various possible methods for constructing
the quasipotential in each specific model. In a single journal paper, it is apparently impossible to cover the
entire spectrum of results achieved in the framework of the quasipotential approach during forty years of
its existence. But we think that already the problems stated in this paper demonstrate the effectiveness
and self-consistency of such an approach in relativistic particle physics.
The quasipotential approach provides the grounds for theoretical descriptions of relativistic composite
systems in the framework of the general principles and methods of a local quantum field theory. It is
important that we can then preserve tools and methods of nonrelativistic quantum mechanics because the
main object of our investigation is the system wave function, which satisfies a definite dynamic equation
and physical boundary conditions.
But the quasipotential itself, in contrast to a classical potential, becomes nonlocal and complex and
explicitly depends on the total system energy even in the case of a purely electromagnetic interaction. For
small particle momenta, it can be approximately reduced to a local potential and can be investigated in the
configuration space of relative distances between particles. We then find that a quasipotential oscillates at
positive binding energies, while the oscillation frequency is determined by the value of the binding energy.

1134
In the case with a negligibly small binding energy, the quasipotential becomes the standard nonrelativistic
Coulomb potential.
In this paper, we briefly described the quasipotential approach in the relativistic configuration space
related to the momentum space by a transformation that uses functions realizing the motion group rep-
resentation in the Lobachevsky space. For constructing a quasipotential interaction of two hadrons in the
configuration space, the hypothesis about the smoothness of a local quasipotential, which reflects the prop-
erty of extendedness and “crumbliness” of hadrons participating in high-energy collisions, is important.
This hypothesis permits explaining many basic regularities of the elastic scattering of high-energy particles
at small and large angles.
We also investigated the problem of particle interaction in the relativistic configuration space. For
this, we used the empirical Coulomb quasipotential. We showed that this quasipotential simulates the
so-called asymptotic freedom property (the property of the “running” coupling constant) in QCD. The
important conclusion is that the spectrum of relativistic bound states in the Coulomb quasipotential with
a large coupling constant drastically differs from the nonrelativistic Coulomb spectrum. We can use the
semiclassical approximation to calculate realistic mass spectra and widths of meson decay if we switch on a
“confining” quasipotential. The relativistic nature of the particle motion drastically changes the boundary
of the classically accessible domain and therefore the quantization rules.
The quasipotential approach also turns out to be effective for describing meson decay form factors.
In this case, the necessity of accounting for the relativistic nature of the wave function of the composite
system, which manifests a huge mass defect, is most apparent.
The quasipotential approach can be easily generalized to the case of interactions of three and more
particles. Here we demonstrated the general solution method with the simplest example of deep inelastic
scattering of a lepton on a hadron to show how the observed meson structure functions can be expressed
through the meson wave function.

REFERENCES

1. N. N. Bogoliubov, A. A. Logunov, A. I. Oksak, and I. T. Todorov, General Principles of Quantum Field Theory
[in Russian], Nauka, Moscow (1987); English transl., Kluwer, Dordrecht (1990).
2. E. E. Salpeter and H. A. Bethe, Phys. Rev., 84, 1232 (1951).
3. A. A. Logunov and A. N. Tavkhelidze, Nuovo Cimento, 29, 380 (1963).
4. W. Macke, Z. Naturforsch, 8a, 599 (1953); W. Zimmermann, Suppl. Nuovo Cimento, 11, 1 (1954); W. Kro-
likowski and J. Rzewuski, Nuovo Cimento, 2, 203 (1955); R. Blankenbecler and R. Sugar, “Single linear integral
equation for relativistic scattering,” in: Proc. 12th Annual Intl. Conf. on HEP (Ya. A. Smorodinskii, ed.), Vol. 1,
Atomizdat, Moscow (1966), p. 225.
5. V. G. Kadyshevsky, Nucl. Phys. B, 6, 125 (1968); A. A. Logunov, V. I. Savrin, N. E. Tyurin, and O. A. Khrustalev,
Theor. Math. Phys., 6, 113 (1971); R. N. Faustov, Ann. Phys., 78, 176 (1973); A. A. Arkhipov and V. I. Savrin,
Fiz. Elem. Chast. At. Yadra, 16, 1091 (1985).
6. P. N. Bogolyubov, “On the sign of the imaginary part of a quasipotential [in Russian],” Preprint JINR P2-5021,
Joint Inst. Nucl. Res., Dubna (1970).
7. A. A. Logunov, A. N. Tavkhelidze, I. T. Todorov, and O. A. Khrustalev, Nuovo Cimento, 30, 134 (1963).
8. R. N. Faustov, Fiz. Elem. Chast. At. Yadra, 3, 238 (1972).
9. B. A. Arbuzov, A. A. Logunov, A. N. Tavkhelidze, R. N. Faustov, and A. T. Filippov, JETP, 17, 950 (1963);
B. A. Arbuzov, A. A. Logunov, A. T. Filippov, and O. A. Khrustalev, JETP, 19, 861 (1964).
10. V. N. Kapshai, V. I. Savrin, and N. B. Skachkov, Theor. Math. Phys., 69, 1226 (1986).
11. B. A. Arbuzov, É. É. Boos, V. I. Savrin, and S. A. Shichanin, JETP Letters, 50, 262 (1989).
12. V. G. Kadyshevsky, R. M. Mir-Kasimov, and N. B. Skachkov, Nuovo Cimento A, 55, 233 (1968); V. G. Kady-
shevsky, R. M. Mir-Kasimov, and N. B. Skachkov, Fiz. Elem. Chast. At. Yadra, 2, 635 (1972).

1135
13. S. P. Alliluyev, S. S. Gershtein, and A. A. Logunov, Phys. Lett., 18, 195 (1965).
14. V. R. Garsevanishvili, V. A. Matveev, and L. A. Slepchenko, Fiz. Elem. Chast. At. Yadra, 1, 91 (1970); V. R. Gar-
sevanishvili, V. A. Matveev, L. A. Slepchenko, and A. N. Tavkhelidze, Phys. Rev. D, 1, 849 (1971).
15. A. A. Logunov and O. A. Khrustalev, Fiz. Elem. Chast. At. Yadra, 1, 71 (1970).
16. V. A. Novikov, L. B. Okun, M. A. Shifman, A. I. Vainshtein, M. V. Voloshin, and V. I. Zakharov, Phys. Rep. C,
41, 1 (1978).
17. V. I. Savrin and N. B. Skachkov, Lett. Nuovo Cimento, 29, 363 (1980).
18. A. D. Donkov, V. G. Kadyshevskii, and M. D. Mateev, Theor. Math. Phys., 50, 236 (1982).
19. V. I. Savrin, A. V. Sidorov, and N. B. Skachkov, Hadronic J., 4, 1642 (1981).
20. V. I. Savrin and N. B. Skachkov, “Form factors and hadron structure functions in single-time formulation
of quantum field theory [in Russian],” in: “Problems of High-Energy Physics and Field Theory” (5th Intl.
Workshop), Vol. II, Inst. for High-Energy Physics, Protvino (1982), p. 229.
21. V. A. Matveev, B. V. Struminskii, and A. N. Tavkhelidze, “Some effects in a quark model [in Russian],” JINR
Communication P-252, Joint Inst. Nucl. Res., Dubna (1965).
22. R. Van Royen and V. F. Weisskopf, Nuovo Cimento A, 50, 617 (1967).
23. D. E. Groom et al. (Particle Data Group), Eur. Phys. J. C, 15, 1 (2000).
24. R. N. Faustov, V. O. Galkin, and A. Yu. Mishurov, Phys. Lett. B, 536, 516 (1995).
25. V. A. Matveev, R. M. Muradyan, and A. N. Tavkhelidze, “Relativistic covariant wave equations for N particles in
quantum field theory [in Russian],” Preprint JINR P2-3900, Joint Inst. Nucl. Res., Dubna (1968); A. A. Arkhipov
and V. I. Savrin, Theor. Math. Phys., 16, 871 (1973).
26. V. A. Matveev, A. N. Sisakyan, and L. A. Slepchenko, Sov. J. Nucl. Phys., 23, 227 (1976); A. N. Kvinikhidze,
A. N. Sisakyan, L. A. Slepchenko, and A. N. Tavkhelidze, Fiz. Elem. Chast. At. Yadra, 8, 478 (1977).
27. M. N. Dubinin, V. I. Savrin, and N. B. Skachkov, Sov. J. Nucl. Phys., 52, 808 (1990).

1136

You might also like