Next Article in Journal
Correction: Teslyk et al. Unruh Effect and Information Entropy Approach. Particles 2022, 5, 157–170
Previous Article in Journal
Characterization of RF System for MIR/THz Free Electron Lasers at Chiang Mai University
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Renormalisable Non-Local Quark–Gluon Interaction: Mass Gap, Chiral Symmetry Breaking and Scale Invariance

1
Institute of Physics, University of Tartu, W. Ostwaldi 1, 50411 Tartu, Estonia
2
Independent Researcher, 00176 Rome, Italy
3
Institute of Theoretical Physics, Faculty of Physics, University of Warsaw, ul. Pasteura 5, 02-093 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Particles 2024, 7(2), 392-415; https://doi.org/10.3390/particles7020022
Submission received: 13 March 2024 / Revised: 4 April 2024 / Accepted: 9 April 2024 / Published: 12 April 2024

Abstract

:
We derive a Nambu–Jona-Lasinio (NJL) model from a non-local gauge theory and show that it has confining properties at low energies. In particular, we present an extended approach to non-local QCD and a complete revision of the technique of Bender, Milton and Savage applied to non-local theories, providing a set of Dyson–Schwinger equations in differential form. In the local case, we obtain closed-form solutions in the simplest case of the scalar field and extend it to the Yang–Mills field. In general, for non-local theories, we use a perturbative technique and a Fourier series and show how higher-order harmonics are heavily damped due to the presence of the non-local factor. The spectrum of the theory is analysed for the non-local Yang–Mills sector and found to be in agreement with the local results on the lattice in the limit of the non-locality mass parameter running to infinity. In the non-local case, we confine ourselves to a non-locality mass that is sufficiently large compared to the mass scale arising from the integration of the Dyson–Schwinger equations. Such a choice results in good agreement, in the proper limit, with the spectrum of the local theory. We derive a gap equation for the fermions in the theory that gives some indication of quark confinement in the non-local NJL case as well. Confinement seems to be a rather ubiquitous effect that removes some degrees of freedom in the original action, favouring the appearance of new observable states, as seen, e.g., for quantum chromodynamics at lower energies.
PACS:
11.10.Lm; 21.30.Fe; 75.40.Gb

1. Introduction

A lesson one can take from string theory is that strings are non-local objects [1,2,3,4,5,6,7,8,9,10,11,12]. Indeed, this lesson is the original motivation for the weakly non-local field theory to devise a novel pathway for possible UV-regularised theories inspired by string field theory [1,13], as investigated in several Refs. since the 1990s [14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39]. Later on, this pathway became an alternative to address the divergence and the hierarchy problems in the Standard Model (SM) via generalising the kinetic energy operators that are of second order in the derivative to an infinite series of higher-order derivatives, suppressed by the scale M 2 of non-locality [14,40,41]. Such modifications of the kinetic energy sector in introducing higher-order derivatives are free of ghosts [33] (the unitarity issues are well addressed due to a certain prescription [42,43,44,45,46]) and also cure the Higgs vacuum instability problem of the SM Higgs [47], as analysed by one of the authors. It was shown that the β function reaches a conformal limit, resolving the Landau pole issue in Quantum Field Theory [48]. Therefore, by capturing infinite derivatives by an exponential of an entire function, we obtain a softened ultraviolet (UV) behaviour in the most suitable manner, without the cost of introducing any new degrees of freedom that contribute to the particle mass spectrum, since there are no new poles in the propagators of infinite-derivative extensions. (For the astrophysical implications, dimensional transmutation and dark matter and dark energy phenomenology in these theories, see Refs. [49,50,51]). A bound on the scale of non-locality from observations from LHC and dark matter physics is M O ( 10 ) TeV [41,50]. Moreover, the non-local theory leads to interesting implications for proton decay and Grand Unified Theories [52], as well as for braneworld models [53]. In addition, strongly coupled non-perturbative regimes, exact β functions and the conditions of confinement in higher-derivative non-local theories are being actively investigated [54,55,56,57,58,59]. The results obtained so far show that the effect of the non-locality in the strong coupling limits is the dilution of any mass gaps (that may be present in the theory) in the UV regime that the system generates. (In the context of gravity theories, one can get rid of classical singularities, such as black hole singularities [19,24,60,61,62,63,64,65,66,67,68,69] and cosmological singularities [70,71,72,73,74,75,76]. Recently, false vacuum tunnelling studies were carried out in Ref. [77]. Relations to string theory are found in Refs. [78,79,80] (for an overview, cf. Ref. [81])).
In order to render charge renormalisation finite, infinite-derivative terms should be introduced for the fermion fields as well [48]. An analogous mechanism is the well-known Pauli–Villars regularisation scheme, where a mass dependence Λ is introduced as the cut-off. The infinite-derivative approach is to promote the Pauli–Villars cut-off Λ to the non-local energy scale M of the theory. In the infinite-derivative case, the non-local energy scale M might play the role of the ultraviolet cut-off. Recently, the infinite-derivative model for QED and Yang–Mills has been reconsidered in view of its generalisation to the SM [47,48]. (See also Refs. [41,50] for a few phenomenological applications to LHC physics and dark matter physics). Indeed, this model leads to a theory that is naturally free of quadratic divergencies, thus providing an alternative way to the solution of the hierarchy problem [41]. Higher-dimensional operators containing new interactions naturally appear in higher-derivative theories with a non-Abelian gauge structure. These operators soften quantum corrections in the UV regime and extinguish divergencies in radiative corrections. Nevertheless, as can be easily understood by power-counting arguments, the new higher-dimensional operators do not break renormalisability [48]. This is due to the improved ultraviolet behaviour of the bosonic propagator P ( k ) in the deep Euclidean region, which scales as P ( k ) exp ( k 2 / M 2 ) / k 4 , instead of the usual propagator scaling in the local case as P ( k ) 1 / k 2 for k 2 [48].
Note that the presence of the non-local energy scale M associated with the infinite-derivative term manifestly breaks (at the classical level) the conformal symmetry of the unbroken gauge sector. Therefore, one may wonder whether this term can also trigger (dynamically) chiral symmetry breaking at low energy, or in other words, whether the fermion field could dynamically obtain a mass m satisfying the condition m < M . The aim of the present paper is to investigate this issue by analysing a general class of renormalisable models containing infinite-derivative terms. In this paper, we will show that a non-vanishing mass term for the fermion field can indeed be generated, depending on the kind of interaction at hand, as a solution of the mass gap equation. The fermion mass can be predicted, and it turns out to be a function of the energy scale M. The effect of non-locality is to move a possible violation of micro-causality to the region beyond the non-locality mass scale M, making them possibly unobservable. On the other hand, diagrammatic techniques a la Feynman cannot properly work in this context, which makes our approach through solutions to the Dyson–Schwinger set of equations more appealing.
This paper is organised as follows. In Section 2, we introduce the infinite-derivative S U ( N ) gauge theory that we aim to study. In Section 3, we derive a set of Dyson–Schwinger equations for non-local QCD. We solve these equations in a perturbative manner by noting that higher harmonics are heavily damped by the non-local factor. In Section 4, the spectrum of the theory is analysed for the non-local Yang–Mills sector and found to be in agreement with the local results on the lattice in the limit of the non-local scale running to infinity. In the non-local case, we confine ourselves to a non-local scale that is sufficiently large with respect to the mass scale arising from the integration of the Dyson–Schwinger equations. In Section 5, we derive the gap equation for the fermion in the theory and show that an identical argument to that given in Refs. [58,59] can be applied here, giving some indication of quark confinement in the non-local case as well. Section 6 contains our conclusions. In Appendix A, we derive the Dyson–Schwinger equations for a ϕ 4 theory in differential form. In Appendix B, we calculate the scalar two-point function.

2. Infinite-Derivative SU(N)

The introduction of infinite-derivative terms in the Lagrangian is based on an approach by Lee and Wick [82,83] (cf. also Ref. [84]). The Lee–Wick approach, understood as the first terms in a series expansion, leads to a non-local infinite-derivative approach to S U ( N ) gauge theories. In infinite-derivative S U ( N ) with massless fermions, we start with the Lagrangian [48]:
L = ψ ¯ γ μ ( i μ g s A μ a T a ) ψ 1 4 F μ ν a e f ( D 2 ) F a μ ν 1 2 ξ ( μ A a μ ) e f ( D 2 ) ( ν A a ν ) + j a μ A μ a
where we assume the non-locality to be only in the gauge sector and F μ ν a = μ A ν a ν A μ a g s f a b c A μ b A ν c . Both the Yang–Mills term and the gauge-fixing term are delocalised by the infinite-derivative exponential. A source term j a μ A μ a is added in order to use the Lagrangian for the generating functional. As an example, we can use f ( D 2 ) = D 2 / M 2 , where D μ a b = δ a b μ i g s A μ c ( T c ) a b is the covariant derivative in the adjoint representation. For a large non-local scale M 2 , it was shown in Refs. [54,55] that e f ( D 2 ) can be approximated by e f ( ) , where = 2 . Therefore, one can start with
L = ψ ¯ γ μ ( i μ g s A μ a T a ) ψ 1 4 F μ ν a e f ( ) F a μ ν 1 2 ξ ( μ A a μ ) e f ( ) ( ν A a ν ) + j a μ A μ a .
Dealing with the non-locality, we use a redefinition of fields by employing
A μ a = e 1 2 f ( ) A ̲ μ a
which leads to
L = ψ ¯ γ μ ( i μ g s [ e 1 2 f ( ) A ̲ μ a ] T a ) ψ + L YM + ,
where
L YM + = 1 4 F ̲ μ ν a F ̲ a μ ν 1 2 ξ [ μ A ̲ a μ ] [ ν A ̲ a ν ] + j a μ [ e 1 2 f ( ) A ̲ μ a ] .
The underline stands for the field redefinition. Here and in the following, the square brackets restrict the operational range of the differential operators included. Using the fact that the Lagrangian is determined up to a total divergence, one can use integration by parts to obtain
F μ ν a e f ( ) F a μ ν = F μ ν a e 1 2 f ( ) e 1 2 f ( ) F a μ ν = [ e 1 2 f ( ) F μ ν a ] [ e 1 2 f ( ) F a μ ν ] = : F ̲ μ ν a F ̲ a μ ν
up to a total divergence (named d i v in the following and vanishing in the action integral), where f [ g ] = [ f ] g + d i v , and (by induction) f [ n g ] = [ n f ] g + d i v is used to share the non-locality equally between the two field strength tensors. Explicitly, one has
F ̲ μ ν a = μ A ̲ ν a ν A ̲ μ a + g s f a b c e 1 2 f ( ) [ e 1 2 f ( ) A ̲ μ b ] [ e 1 2 f ( ) A ̲ ν c ] .
For the modified Yang–Mills and gauge-fixing Lagrangian, one obtains
L YM + = 1 4 [ μ A ̲ ν a ν A ̲ μ a ] [ μ A ̲ a ν ν A ̲ a μ ] 1 2 ξ [ μ A ̲ a μ ] [ ν A ̲ a ν ] + j a μ [ e 1 2 f ( ) A ̲ μ a ] ( ( 1 2 g s f a b c [ μ A ̲ ν a ν A ̲ μ a ] e 1 2 f ( ) [ e 1 2 f ( ) A ̲ b μ ] [ e 1 2 f ( ) A ̲ c ν ] ( ( 1 4 g s 2 f a b c f a d e [ e 1 2 f ( ) [ e 1 2 f ( ) A ̲ b μ ] [ e 1 2 f ( ) A ̲ c ν ] ] [ e 1 2 f ( ) [ e 1 2 f ( ) A ̲ μ d ] [ e 1 2 f ( ) A ̲ ν e ] ] = 1 2 [ μ A ̲ ν a ] [ μ A ̲ a ν ν A ̲ a μ ] 1 2 ξ [ μ A ̲ a μ ] [ ν A ̲ a ν ] i j a μ [ e 1 2 f ( ) A ̲ μ a ] ( ( 1 2 g s f a b c [ μ A ̲ ν a ν A ̲ μ a ] e 1 2 f ( ) [ e 1 2 f ( ) A ̲ b μ ] [ e 1 2 f ( ) A ̲ c ν ] ( ( 1 4 g s 2 f a b c f a d e [ e 1 2 f ( ) A ̲ b μ ] [ e 1 2 f ( ) A ̲ c ν ] e f ( ) [ e 1 2 f ( ) A ̲ μ d ] [ e 1 2 f ( ) A ̲ ν e ] + d i v ,
where the first term has been simplified, and integrations by parts have been used to move the exponential derivatives to the central position. This can also be reverted if necessary, i.e., if the derivatives are inappropriate for the variation in the fields. An ordinary integration by parts in the first line for the second part of S = L d 4 x = S f + S YM + gives
S YM + = d 4 x [ 1 2 A ̲ μ a ( x ) ( η μ ν μ ν ) A ̲ ν a ( x ) + 1 2 ξ A ̲ μ a ( x ) μ ν A ̲ ν a ( x ) + j a μ ( x ) [ e 1 2 f ( ) A ̲ μ a ( x ) ] ( ( 1 2 g s f a b c [ μ A ̲ ν a ( x ) ν A ̲ μ a ( x ) ] e 1 2 f ( ) [ e 1 2 f ( ) A ̲ b μ ( x ) ] [ e 1 2 f ( ) A ̲ c ν ( x ) ] ( ( 1 4 g s 2 f a b c f c d e [ e 1 2 f ( ) A ̲ a μ ( x ) ] [ e 1 2 f ( ) A ̲ b ν ( x ) ] e f ( ) [ e 1 2 f ( ) A ̲ μ d ( x ) ] [ e 1 2 f ( ) A ̲ ν e ( x ) ] ] .
The Euler–Lagrange equation δ S / δ A ̲ μ a = 0 for the Yang–Mills field reads
g s e 1 2 f ( ) ψ ¯ γ μ T a ψ = ( η μ ν μ ν ) A ̲ ν a + 1 ξ μ ν A ̲ ν a + e 1 2 f ( ) j a μ g s f a b c e 1 2 f ( ) ν [ e 1 2 f ( ) A ̲ b μ ] [ e 1 2 f ( ) A ̲ c ν ] ( ( g s f a b c e 1 2 f ( ) [ e 1 2 f ( ) A ̲ ν b ] e 1 2 f ( ) ( μ A ̲ c ν ν A ̲ c μ ) + ( ( g s 2 f a b c f c d e e 1 2 f ( ) [ e 1 2 f ( ) A ̲ ν b ] e f ( ) [ e 1 2 f ( ) A ̲ d μ ] [ e 1 2 f ( ) A ̲ e ν ] .
In the following, we use the Feynman gauge ξ = 1 to further simplify the first line. Applying the formalism of Bender, Milton and Savage [85] to the Euler–Lagrange Equation (10) leads to the tower of Dyson–Schwinger equations, which is formulated in terms of multiple-point Green functions. This approach is displayed in Appendices A and B for a local ϕ 4 theory.

3. Solution of Dyson–Schwinger Equations

The creation of the tower of Dyson–Schwinger equations is explained in detail in Appendix B of Ref. [54]. We do not repeat it here to avoid reprinting previously published material. Instead, we give some guidelines. To start with, one takes the expectation value of the Euler–Lagrange Equation (10), weighted by the generating functional
Z [ j ] = [ d A ̲ ] exp i L YM + d 4 x ,
where the current j is already contained in L YM + , indicated by the plus sign in the index. One has e 1 2 f ( x ) A ̲ μ a ( x ) = : Z [ j ] e 1 2 f ( x ) G ̲ 1 μ ( j ) a ( x ) . By either applying an additional partial derivative like e 1 2 f ( x ) ν x A ̲ μ a ( x ) = Z [ j ] e 1 2 f ( x ) ν x G ̲ 1 μ ( j ) a ( x ) or calculating a variation with respect to, e.g., j b ν ( x ) ,
e 1 2 f ( x ) A ̲ μ a ( x ) e 1 2 f ( x ) A ̲ ν b ( x ) = Z [ j ] e 1 2 f ( x ) G ̲ 2 μ ν ( j ) a b ( x , x ) + Z [ j ] e 1 2 f ( x ) G ̲ 1 μ ( j ) a ( x ) e 1 2 f ( x ) G ̲ 1 ν ( j ) b ( x ) ,
one finally obtains
G ̲ 1 μ ( j ) a ( x ) + e 1 2 f ( ) j μ a ( x ) = = g s f a b c { e 1 2 f ( ) ν [ e 1 2 f ( ) G ̲ 2 μ ν ( j ) b c ( x , x ) ] + [ e 1 2 f ( ) G ̲ 1 μ ( j ) b ( x ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) c ( x ) ] ( ( + e 1 2 f ( ) [ e 1 2 f ( ) ( μ G ̲ 2 ν b ( j ) c ν ( x , x ) ν G ̲ 2 μ b ( j ) c ν ( x , x ) ) ] ( ( + e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 1 b ( j ) ν ( x ) ] e f ( ) [ e 1 2 f ( ) ( μ G ̲ 1 ν ( j ) c ( x ) ν G ̲ 1 μ ( j ) c ( x ) ) ] } ( ( + g s 2 f a b c f c d e { e 1 2 f ( ) ( [ e 1 2 f ( ) G ̲ 3 μ ν b ( j ) d e ν ( x , x , x ) ] + [ e 1 2 f ( ) G ̲ 2 μ b ( j ) d ν ( x , x ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) e ( x ) ] ( ( + [ e 1 2 f ( ) G ̲ 1 μ ( j ) d ( x ) ] [ e 1 2 f ( ) G ̲ 2 ν b ( j ) e ν ( x , x ) ] ) ( ( + e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 1 b ( j ) ν ( x ) ] e f ( ) ( [ e 1 2 f ( ) G ̲ 2 μ ν ( j ) d e ( x , x ) ] ( ( + [ e 1 2 f ( ) G ̲ 1 μ ( j ) d ( x ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) e ( x ) ] ) } .
The equation of motion for the two-point function is obtained by variation with respect to j h λ ( y ) :
G ̲ 2 μ λ ( j ) a h ( x , y ) i δ a h η μ λ e 1 2 f ( ) δ ( 4 ) ( x y ) = = g s f a b c { e 1 2 f ( ) ν ( [ e 1 2 f ( ) G ̲ 3 μ ν λ ( j ) b c h ( x , x , y ) ] + [ e 1 2 f ( ) G ̲ 2 μ λ ( j ) b h ( x , y ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) c ( x ) ] ( ( + [ e 1 2 f ( ) G ̲ 1 μ ( j ) b ( x ) ] [ e 1 2 f ( ) G ̲ 2 ν λ ( j ) c h ( x , y ) ] ) ( ( + e 1 2 f ( ) [ e 1 2 f ( ) ( μ G ̲ 3 ν b λ ( j ) c ν h ( x , x , y ) ν G ̲ 3 μ b λ ( j ) c ν h ( x , x , y ) ) ] ( ( + e 1 2 f ( ) ( [ e 1 2 f ( ) G ̲ 2 b λ ( j ) ν h ( x , y ) ] e f ( ) [ e 1 2 f ( ) ( μ G ̲ 1 ν ( j ) c ( x ) ν G ̲ 1 μ ( j ) c ( x ) ) ] ( ( + [ e 1 2 f ( ) G ̲ 1 b ( j ) ν ( x ) ] e f ( ) [ e 1 2 f ( ) ( μ G ̲ 2 ν λ ( j ) c h ( x , y ) ν G ̲ 2 μ λ ( j ) c h ( x , y ) ) ] ) } ( ( + g s 2 f a b c f c d e { e 1 2 f ( ) ( [ e 1 2 f ( ) G ̲ 4 μ ν b λ ( j ) d e ν h ( x , x , x , y ) ] ( ( + [ e 1 2 f ( ) G ̲ 3 μ ν b ( j ) d ν h ( x , x , y ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) e ( x ) ] + [ e 1 2 f ( ) G ̲ 2 μ b ( j ) d ν ( x , x ) ] [ e 1 2 f ( ) G ̲ 2 ν λ ( j ) e h ( x , y ) ] ( ( + [ e 1 2 f ( ) G ̲ 2 μ λ ( j ) d h ( x , y ) ] [ e 1 2 f ( ) G ̲ 2 ν b ( j ) e ν ( x , x ) ] + [ e 1 2 f ( ) G ̲ 1 μ ( j ) d ( x ) ] [ e 1 2 f ( ) G ̲ 3 ν b λ ( j ) e ν h ( x , x , y ) ] ) ( ( + e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 2 b λ ( j ) ν h ( x , y ) ] × ( ( × e f ( ) ( [ e 1 2 f ( ) G ̲ 2 μ ν ( j ) d e ( x , x ) ] + [ e 1 2 f ( ) G ̲ 1 μ ( j ) d ( x ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) e ( x ) ] ) ( ( + e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 1 b ( j ) ν ( x ) ] e f ( ) ( [ e 1 2 f ( ) G ̲ 3 μ ν λ ( j ) d e h ( x , x , y ) ] ( ( + [ e 1 2 f ( ) G ̲ 2 μ λ ( j ) d h ( x , y ) ] [ e 1 2 f ( ) G ̲ 1 ν ( j ) e ( x ) ] + [ e 1 2 f ( ) G ̲ 1 μ ( j ) d ( x ) ] [ e 1 2 f ( ) G ̲ 2 ν λ ( j ) e h ( x , y ) ] ) } ,
with δ G ̲ 1 μ ( j ) a ( x ) / δ j h λ ( y ) = G ̲ 2 μ λ ( j ) a h ( x , y ) . The system of equations of motion for the complete set of components of the Green functions for the Yang–Mills field cannot be treated exactly. Instead, we use a mapping to the scalar case. The mapping theorem introduced in Refs. [86,87] is based on Andrei Smilga’s solution for the problem that “the Yang–Mills system is not exactly solvable, … in contrast to … some early hopes” (cf. Sec. 1.2 in Ref. [88]). Indeed, for more than one independent component, in solving the system of equations, one observes chaotic behaviour (cf. Refs. [89,90,91]). Even though the importance of such observations for macroscopic behaviour is questionable, the safe path is to use a mapping to a single scalar function that we dub ϕ ̲ ( x ) , representing an exact solution for the 1P-correlation function in the scalar case. This is realized here by using G ̲ 1 μ ( 0 ) a ( x ) = η μ a ϕ ̲ ( x ) and G ̲ 2 μ ν ( 0 ) a b ( x , y ) = η μ ν δ a b G ̲ 2 ( x y ) for j μ a ( x ) set to zero, where η μ a represents the components of the polarisation vector and η μ ν denotes the components of the Minkowski metric. The two-point Green functions from x to x become constants with a vanishing derivative, while n-point Green functions with n > 2 can be set to zero if at least two arguments coincide. One obtains
η μ a ϕ ̲ ( x ) = g s f a b c { e 1 2 f ( ) ν η μ b η ν c [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 ( ( + e 1 2 f ( ) η b ν [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ( η ν c μ ϕ ̲ ( x ) η μ c ν ϕ ̲ ( x ) ) ] } ( ( + g s 2 f a b c f c d e { e 1 2 f ( ) ( η μ ν δ b d η ν e + η ν ν δ b e η μ d ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] ( ( + e 1 2 f ( ) η b ν [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) ( × η μ ν δ d e [ e 1 2 f ( ) G ̲ 2 ( x x ) ] + η μ d η ν e [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 } = g s f a b c { e 1 2 f ( ) η μ b η c ν ν [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 ( ( + e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ( η b ν η ν c μ ϕ ̲ ( x ) η b ν η μ c ν ϕ ̲ ( x ) ) ] } ( ( + g s 2 f a b c f c d e { ( δ b d η μ e + D δ b e η μ d ) e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] ( ( + η b ν η μ d η ν e e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 }
and
δ a h η μ λ G ̲ 2 ( x y ) i δ a h η μ λ e 1 2 f ( ) δ ( 4 ) ( x y ) = = g s f a b c { e 1 2 f ( ) ( η μ λ δ b h η ν c + η ν λ η μ b δ c h ) ν [ e 1 2 f ( ) ϕ ̲ ( x ) ] [ e 1 2 f ( ) G ̲ 2 ( x y ) ] ( ( + e 1 2 f ( ) δ b h η λ ν [ e 1 2 f ( ) G ̲ 2 ( x y ) ] e f ( ) [ e 1 2 f ( ) ( η ν c μ ϕ ̲ ( x ) η μ c ν ϕ ̲ ( x ) ) ] ( ( + e 1 2 f ( ) η b ν δ c h [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ( η ν λ μ η μ λ ν ) G ̲ 2 ( x y ) ] ( ( + g s 2 f a b c f c d e { η μ λ ( δ b d δ e h + δ b e δ d h η ν ν ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) G ̲ 2 ( x y ) ] ( ( + η λ ν δ b h e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 2 ( x y ) ] ( × e f ( ) η μ ν δ d e [ e 1 2 f ( ) G ̲ 2 ( x x ) ] + η μ d η ν e [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 ( ( + η b ν ( η μ λ δ d h η ν e + η ν λ δ e h η μ d ) [ e 1 2 f ( ) G ̲ 2 ( x y ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 } .
Contracting with η a μ and δ a h η μ λ and using the orthogonality and completeness relations
η μ a η b μ = δ b a , η μ a η ν a = ( N c 2 1 ) η μ ν / D ,
where ( η μ ν ) = d i a g ( 1 , 1 , 1 , 1 ) ( D = η μ λ η μ λ is the space–time dimension), one obtains
( N c 2 1 ) ϕ ̲ ( x ) = g s 2 f a b c f c d e { ( δ b d δ a e + D δ b e δ a d ) e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] ( ( + δ b e δ a d e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 } = N c ( N c 2 1 ) g s 2 { ( D 1 ) e 1 2 f ( ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] ( ( + e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 }
and
D ( N c 2 1 ) G ̲ 2 ( x y ) i e 1 2 f ( ) δ ( 4 ) ( x y ) = = g s 2 f a b c f c d e { D ( δ b d δ a e + D δ b e δ a d ) [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) G ̲ 2 ( x y ) ] ( ( ( D δ a d δ b e + δ a e δ b d ) [ e 1 2 f ( ) G ̲ 2 ( x y ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 } = ( D 1 ) N c ( N c 2 1 ) g s 2 D [ e 1 2 f ( ) G ̲ 2 ( x x ) ] [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 [ e 1 2 f ( ) G ̲ 2 ( x y ) ] ,
where f a b c f a b d = N c δ c d for two and, therefore, f a b c f a b c = N c ( N c 2 1 ) for three summed indices have been used. The common factor cancels, and one obtains
( + Δ m G 2 ) ϕ ̲ ( x ) + λ e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 = 0
and
( + Δ m G 2 ) G ̲ 2 ( x y ) + D 1 D λ [ e 1 2 f ( ) ϕ ̲ ( x ) ] 2 [ e 1 2 f ( ) G ̲ 2 ( x y ) ] = i e 1 2 f ( ) δ ( 4 ) ( x y ) ,
where λ = N c g s 2 and Δ m G 2 = ( D 1 ) e 1 2 f ( x ) λ G 2 ( x x ) | x = x .
Looking at Equation (20), the solution for the corresponding local one-point function, given by ϕ ̲ 0 ( x ) = ϕ ^ 0 ( k x ) = μ sn ( k x + θ | κ ) [86], obeys the dispersion relation k 2 = Δ m G 2 + λ μ 2 / 2 , with κ = ( Δ m G 2 k 2 ) / k 2 and θ = ( 1 + 4 N ) K ( κ ) . The newly introduced constant μ needs some explanation. This is an integration constant coming from the solution of the Dyson–Schwinger equation for the one-point correlation function, with the dimension of energy. It sets the scale for our non-perturbative solution. Therefore, it should be related to the constant Λ QCD that emerges from dimensional transmutation in perturbative computations, setting the scale for the confined phase of Yang–Mills theory. The solution ϕ ^ 0 ( k x ) can be expanded in a Fourier series,
ϕ ^ 0 ( k x ) = 2 η κ 2 k 2 κ λ n = 0 q n + 1 / 2 1 q 2 n + 1 sin 2 n + 1 2 η ( k x + θ ) = 2 i η 2 k 2 λ n = 0 ( 1 ) n q n + 1 / 2 1 q 2 n + 1 cos 2 n + 1 2 η k x = i μ η n = 0 ( 1 ) n q n + 1 / 2 1 q 2 n + 1 ( e ( n + 1 / 2 ) i η k x + e ( n + 1 / 2 ) i η k x ) = i μ η n = b 2 n + 1 e ( n + 1 / 2 ) i η k x = i μ η m odd b m e i m η k x / 2 ,
with η = π / K ( κ ) . For non-negative values of n, one obviously has
b 2 n + 1 = ( 1 ) n q n + 1 / 2 1 q 2 n + 1 .
However, this can also be extended to negative values, such as for
b 2 n 1 = ( 1 ) n 1 q n 1 / 2 1 q 2 n 1 = ( 1 ) n q 2 n + 1 q n 1 / 2 q 2 n + 1 ( 1 q 2 n 1 ) = ( 1 ) n q n + 1 / 2 q 2 n + 1 1 = ( 1 ) n q n + 1 / 2 1 q 2 n + 1 = b 2 n + 1 .
First-order non-locality effects are taken into account by solving
( + Δ m G 2 ) ϕ ̲ 1 ( x ) = λ e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] 2 .
For the solution ϕ ̲ 1 ( x ) = ϕ ^ 1 ( k x ) , we can use a similar ansatz:
ϕ ^ 1 ( k x ) = i η μ m odd b m e i m η k x / 2 ,
where the b m values are to be computed. Substituting ϕ ̲ 0 ( x ) = ϕ ^ 0 ( k x ) on the right-hand side and ϕ ̲ 1 ( x ) = ϕ ^ 1 ( k x ) on the left-hand side, on the right-hand side, one calculates, step by step,
[ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] 2 = η 2 μ 2 m 2 , m 3 odd b m 2 b m 3 e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) e i ( m 2 + m 3 ) η k x / 2 , e f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] 2 = η 2 μ 2 m 2 , m 3 odd b m 2 b m 3 e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) ( ( × e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) e i ( m 2 + m 3 ) η k x / 2 , [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] 2 = i η 3 μ 3 m 1 , m 2 , m 3 odd b m 1 b m 2 b m 3 ( ( × e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) e i ( m 1 + m 2 + m 3 ) η k x / 2 , e 1 2 f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] e f ( ) [ e 1 2 f ( ) ϕ ̲ 0 ( x ) ] 2 = i η 3 μ 3 m 1 , m 2 , m 3 odd b m 1 b m 2 b m 3 e 1 2 f ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 ) ( ( × e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) e i ( m 1 + m 2 + m 3 ) η k x / 2 .
Therefore, the coefficients b m are determined by the following system of equations:
b m ( m 2 η 2 k 2 / 4 + Δ m G 2 ) = λ η 2 μ 2 m 1 , m 2 , m 3 odd m 1 + m 2 + m 3 = m b m 1 b m 2 b m 3 ( ( × e 1 2 f ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) .
One can solve for b m to obtain
b m = λ η 2 μ 2 m 1 , m 2 , m 3 odd m 1 + m 2 + m 3 = m b m 1 b m 2 b m 3 ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 + Δ m G 2 ) 1 ( ( × e 1 2 f ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) ,
and after inserting it back into ϕ ̲ 1 ( x ) , as the first iteration, one has
ϕ ̲ 1 ( x ) = i λ η 3 μ 3 m 1 , m 2 , m 3 odd b m 1 b m 2 b m 3 e i ( m 1 + m 2 + m 3 ) η k x / 2 ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 Δ m G 2 ( ( × e 1 2 f ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) .
However, this is not the physical field. Instead, going back to physics, we have to revert the field redefinition, obtaining ϕ 1 ( x ) = e 1 2 f ( ) ϕ ̲ 1 ( x ) or
ϕ 1 ( x ) = i λ η 3 μ 3 m 1 , m 2 , m 3 odd b m 1 b m 2 b m 3 e i ( m 1 + m 2 + m 3 ) η k x / 2 ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 Δ m G 2 ( ( × e f ( ( m 1 + m 2 + m 3 ) 2 η 2 k 2 / 4 ) e f ( ( m 2 + m 3 ) 2 η 2 k 2 / 4 ) e 1 2 f ( m 1 2 η 2 k 2 / 4 ) e 1 2 f ( m 2 2 η 2 k 2 / 4 ) e 1 2 f ( m 3 2 η 2 k 2 / 4 ) .
Due to the denominator factor, for f ( ) = / M 2 , the series turns out to be convergent.
The same is performed for Equation (21) in the case of D = 4 . After inserting ϕ ̲ ( x ) = ϕ ̲ 0 ( x ) , in momentum space, one has
( p 2 Δ m G 2 e 1 2 f ( p 2 ) ) G ˜ ̲ 2 ( p ) i e 1 2 f ( p 2 ) ( = 3 4 λ μ 2 η 2 m , n odd b m b n e 1 2 f ( m 2 η 2 k 2 / 4 ) e 1 2 f ( n 2 η 2 k 2 / 4 ) e 1 2 f ( ( p + 1 2 ( m + n ) η k ) 2 ) G ˜ ̲ 2 ( p + 1 2 ( m + n ) η k ) .
The localised equation is solved in Appendix B and, according to the Källén–Lehmann theorem, provides a spectrum of harmonics, known as glueball states, which can be understood as a form of the glueball operators in Yang–Mills theory according to our solution. Gaussian factors can be used to restrict the solution to the lowest harmonics m , n = ± 1 , resulting in
( p 2 Δ m G 2 e 1 2 f ( p 2 ) ) G ˜ ̲ 2 ( p ) i e 1 2 f ( p 2 ) = m G 2 e 1 2 f ( p 2 ) G ˜ ̲ 2 ( p ) + 1 2 e 1 2 f ( ( p η k ) 2 ) G ˜ ̲ 2 ( p η k ) + 1 2 e 1 2 f ( ( p + η k ) 2 ) G ˜ ̲ 2 ( p + η k ) ,
where
m G 2 = 3 2 λ μ 2 η 2 b 1 2 e f ( η 2 k 2 / 4 ) = 3 κ η 2 k 2 b 1 2 e f ( η 2 k 2 / 4 )
The equation can be solved for G ˜ ̲ 2 ( p ) , resulting in
G ˜ ̲ 2 ( p ) = Δ ˜ ̲ ( p ) 1 + i 2 m G 2 e 1 2 f ( p 2 ) e 1 2 f ( ( p η k ) 2 ) G ˜ ̲ 2 ( p η k ) + e 1 2 f ( ( p + η k ) 2 ) G ˜ ̲ 2 ( p + η k ) ,
where
Δ ˜ ̲ ( p ) : = i e 1 2 f ( p 2 ) p 2 ( m G 2 + Δ m G 2 ) e 1 2 f ( p 2 ) .
Equation (35) can be solved iteratively up to arbitrary orders of m G 2 , as shown in Ref. [56]. For simplicity, we take only the leading-order approximation G ˜ ̲ 2 ( p ) Δ ˜ ̲ ( p ) .

4. The Mass Gap Equation for the Glueball

The mass gap equation obtained from Δ m G 2 = ( D 1 ) λ e 1 2 f ( x ) G 2 ( x x ) | x = x reads
Δ m G 2 = d 4 p ( 2 π ) 4 3 i λ e f ( p 2 ) p 2 ( m G 2 + Δ m G 2 ) e 1 2 f ( p 2 ) = i d 4 p E ( 2 π ) 4 3 i λ e f ( p E 2 ) p E 2 ( m G 2 + m G 2 ) e 1 2 f ( p E 2 ) = 3 λ ( 4 π ) 2 0 ρ 3 e f ( ρ 2 ) d ρ ρ 2 + ( m G 2 + Δ m G 2 ) e 1 2 f ( ρ 2 ) .
Note that both m G 2 as defined before and Δ m G 2 = ( 1 + κ ) k 2 contain the common factor k 2 . This common factor sets the scale of our solution, as it is proportional to μ 2 , the scale of confinement in Yang–Mills theory, as discussed above. Therefore, if we were to also consider quarks, μ would be related to Λ QCD as the scale of asymptotic freedom. The lattice community will recognise this common factor as what they call the “string tension”. At this stage, we do not want to evaluate the inter-quark potential to see the full relation between all three of these experimental constants. What we need to compare with lattice data are just the pure numbers arising from the ratio with such general confinement scales. While the integral is regular at the lower limit, it has to be regularised for the upper limit. We emphasise that the fact that there is a non-local factor does not imply at all that all the integrals in the theory are UV-finite. What such an approach grants is a physical cut-off scale, M, given by the theory, working like a horizon to keep such integrals finite. Dimensional regularisation is not applicable in this case, as the integral cannot be computed analytically, in contrast to the local limit M . One might think of ρ 2 = M 2 as an upper cut-off. In this case, the integral will diverge as M 2 , and a truncation of this highest power is necessary in order to obtain a finite local limit. Instead, our proposal is to use an upper cut-off Λ 2 fixed up to the scale k 2 , which is determined by matching the results to values obtained on the lattice for different values of N c = 2 , 3 , 4 , 5 , 6 , 8 , 10 , 12 [92]. Results from an older Ref. [93] were analysed in the local case in Refs. [94,95], with excellent agreement. The idea is to fix the cut-off scale for the mass gap integral using the results of the local theory (given for M ) in order to prevent changes to the physics with the introduction of the non-locality. Such changes, if any, do not appear experimentally, and they should be really tiny. In order to compare with the lattice, instead of λ as the input, one has to use λ = N c g s 2 = 2 N c 2 / β with β = 2 N c / g s 2 . The values for β are taken from Ref. [92] and listed in the second column of Table 1.
(A recent paper about the Casimir effect in non-Abelian gauge theories on the lattice [96] has shown that the ground state of local Yang–Mills theory is not the same as that found in lattice computations (cf. Table 1). The authors of Ref. [96] obtained M G lat / k 2 = 1.0 ( 1 ) , which, for us, corresponds to the choice n = 0 in the glueball spectrum displayed in Appendix B and Refs. [94,95]. This is also in agreement with the picture of the Casimir effect in local Yang–Mills theory presented in Ref. [97]). The agreement is excellent. For the fit of the mass values in the local limit, we used the method of least squares by minimising
σ ( Λ ) = 1 N i = 1 N ( M G ( i ) est ( Λ ) M G ( i ) lat ) 2 ,
with M G 2 = m G 2 + Δ m G 2 . The result is given by Λ 2 = 177.16 ( 3 ) k 2 . As this result suggests, the agreement between the lattice values and our estimates is very good, as these are found in the remaining lines of Table 1. In Figure 1, we show the solution of the mass gap Equation (37) for M G 2 for the upper limit Λ 2 = 177.16 k 2 in relation to the non-local scale M 2 , together with the dependence of the parameter κ . This could possibly be of the order of TeV 2 . It is interesting to point out that, in this way, we can obtain a physical understanding of the non-local scale and its proper order of magnitude as compared to local physics.

5. The Mass gap Equation of the Quark

The solution obtained in the previous section is input for the determination of the dynamical quark mass. Namely, the Green function G ̲ 2 μ ν ( 0 ) a b ( x , y ) = η μ ν δ a b G ̲ 2 ( x y ) , with G ̲ 2 ( x y ) satisfying the differential Equation (21), has to be convoluted with the left-hand side of the equation of motion (10) in order to obtain
A ̲ a ν ( x ) = i g s G ̲ 2 ( x y ) e 1 2 f ( y ) ψ ¯ ( y ) γ ν T a ψ ( y ) d 4 y .
This result is inserted into the Euler–Lagrange equation for the quark field,
0 = L ψ ¯ μ L ( μ ψ ¯ ) = L ψ ¯ = γ μ i μ g s T a e 1 2 f ( ) A ̲ μ a ψ
to obtain
0 = i γ μ μ ψ ( x ) + i g s 2 γ μ T a ψ ( x ) e 1 2 f ( ) G ̲ 2 ( x y ) e 1 2 f ( y ) ψ ¯ ( y ) γ μ T a ψ ( y ) d 4 y .
This equation of motion can be understood as the equation of motion of a non-local Nambu–Jona-Lasinio (NJL) model. After integration by parts, we derive the NJL Lagrangian
L NJL = ψ ¯ ( x ) i γ μ μ ψ ( x ) + i g s 2 ψ ¯ ( x ) γ μ T a ψ ( x ) e 1 2 f ( ) [ e 1 2 f ( y ) G ̲ 2 ( x y ) ] ψ ¯ ( y ) γ μ T a ψ ( y ) d 4 y .
The standard procedure to arrive at the mass gap equation is taken from Ref. [98] and consists of several steps. First of all, a Fierz rearrangement leads to the action integral S [ ψ , ψ ¯ ] = S 0 [ ψ , ψ ¯ ] + S int [ ψ , ψ ¯ ] with
S 0 [ ψ , ψ ¯ ] = d 4 x ψ ¯ i ( x ) i γ μ μ ψ i ( x ) a n d S int [ ψ , ψ ¯ ] = i g s 2 G ̲ 2 f ( x y ) ψ ¯ i ( x ) Γ α i j ψ j ( y ) ψ ¯ j ( y ) Γ j i α ψ i ( x ) d 4 y = i g s 2 d 4 w d 4 z ψ ¯ ( w + z 2 ) Γ α ψ ( w z 2 ) G ̲ 2 f ( z ) ψ ¯ ( w z 2 ) Γ α ψ ( w + z 2 )
with G ̲ 2 f ( z ) = e f ( ) G ̲ 2 ( z ) . This action integral is an ingredient for the functional integral
Z = D ψ D ψ ¯ exp ( i S [ ψ , ψ ¯ ] ) .
In the following, we use i G ̲ 2 f ( z ) = G C ̲ 2 ( z ) / 2 with C ̲ 2 ( z ) d 4 z = 1 . The quartic interaction term can be removed by introducing the meson field ( ϕ α ) = ( σ , π ) via the factor
D ϕ exp i 2 G d 4 z C ̲ 2 ( z ) d 4 w ϕ α * ( w ) ϕ α ( w ) ,
into the functional integral, interchanging the integrations over w and z and performing the functional “shift”:
ϕ α ( w ) ϕ α ( w ) + g s G ψ ¯ ( w z 2 ) Γ α ψ ( w + z 2 ) .
In doing so and returning in part to x and y, one ends up with the action functional
S [ ψ , ψ ¯ , ϕ ] = 1 2 G d 4 z C ̲ 2 ( z ) d 4 w ϕ α * ( w ) ϕ α ( w ) ( ( + d 4 x d 4 y ψ ¯ ( x ) δ ( 4 ) ( x y ) i γ μ μ g s C ̲ 2 ( x y ) Re ϕ α x + y 2 Γ α ψ ( y ) ,
The functional integral reads
Z = D ψ D ψ ¯ D ϕ exp ( i S [ ψ , ψ ¯ , ϕ ] ) .
Performing a Fourier transform and integrating out the fermionic degrees of freedom, one ends up with the functional determinant and the bosonised action
S bos [ ϕ ] = 1 2 G d 4 q ( 2 π ) 4 ϕ α * ( q ) ϕ α ( q ) ( + i ln det ( 2 π ) 4 δ ( 4 ) ( p p ) γ μ p μ 1 2 g s C ˜ ̲ 2 p + p 2 ( ϕ α ( p p ) + ϕ α * ( p p ) ) Γ α ,
where the logarithm of the determinant is understood to be between states p | | p in momentum space. While in the mean field approximation, the first term gives σ ¯ 2 / 2 G , the determinant (for N f = 2 flavours) leads to
det ( γ μ p μ M q ( p ) 1 l ) = ( p 2 M q 2 ( p ) ) 4 N c ,
with M q ( p ) = g s C ˜ ̲ 2 ( p ) σ ¯ . The variation in the action integral with respect to the mean value σ ¯ leads to
σ ¯ = 8 i N c g s G d 4 p ( 2 π ) 4 C ˜ ̲ 2 ( p ) M q ( p ) p 2 M q 2 ( p ) ,
and the insertion into the definition of M q ( p ) finally leads to the mass gap equation
M q ( p ) = 8 i N c g s 2 G C ˜ ̲ 2 ( p ) d 4 q ( 2 π ) 4 C ˜ ̲ 2 ( q ) M q ( q ) q 2 M q 2 ( q ) = 16 λ G ˜ 2 ( p ) G ˜ ̲ 2 ( 0 ) d 4 q ( 2 π ) 4 G ˜ ̲ 2 ( q ) M q ( q ) q 2 M q 2 ( q ) ,
where we insert C ˜ ̲ 2 ( p ) = 2 i G ˜ ̲ 2 ( p ) / G and use
G ̲ 2 f ( z ) = d 4 q ( 2 π ) 4 e f ( ) G ˜ ̲ 2 ( q ) e i q z
to obtain
G 2 = i G ̲ 2 f ( z ) d 4 z = i d 4 q ( 2 π ) 4 e f ( q 2 ) G ˜ ̲ 2 ( q ) e i q z d 4 z = i G ˜ ̲ 2 ( 0 ) .
In order to solve the gap equation for the quark, we assume that the mass gap M q ( p ) does not depend explicitly on the momentum. In this case, one has to solve the equation
M q = 16 λ d 4 q ( 2 π ) 4 e f ( q 2 ) G ˜ ̲ 2 ( q ) M q q 2 M q 2 .
In principle, M q can be cancelled on both sides, leading to a gap equation similar to Equation (3.7) in the original NJL publication [99]. However, in order to solve this equation iteratively, it is more appropriate to instead multiply it with M q in order to determine M q 2 from
M q 2 = 16 N c g s 2 d 4 q ( 2 π ) 4 e f ( q 2 ) G ˜ ̲ 2 ( q ) M q 2 q 2 M q 2 = d 4 q ( 2 π ) 4 16 i λ M q 2 e 3 2 f ( q 2 ) ( q 2 M G 2 e 1 2 f ( q 2 ) ) ( q 2 M q 2 ) = d 4 q E ( 2 π ) 4 16 λ M q 2 e 3 2 f ( q E 2 ) ( q E 2 M G 2 e 1 2 f ( q E 2 ) ) ( q E 2 M q 2 ) = 8 λ ( 4 π ) 2 0 Λ 2 M q 2 ρ 2 e 3 2 f ( ρ 2 ) d ρ 2 ( ρ 2 + M G 2 e 1 2 f ( ρ 2 ) ) ( ρ 2 + M q 2 ) .
The upper limit is taken to be the same as in the previous section, and the solution M G 2 of the mass gap equation for the glueball is used. The fixed-point problem converges, and one obtains a dependence on the non-local scale M 2 , which is displayed in Figure 2. We can see that chiral symmetry breaking appears in non-local QCD as well.
It is shown in Refs. [58,59] that an educated guess for confinement is given by the threshold M q / M G > 0.39 , above which two real pole solutions for p 2 in the equation
p 2 = M q 2 G ˜ 2 ( p ) G ˜ 2 ( 0 ) 2 = M q 2 M G 4 ( p 2 M G 2 ) 2
change to imaginary pole solutions. According to Figure 3, this is the case for the non-local scale M 2 below 150 k 2 .

6. Discussion and Conclusions

The spectrum of the Yang–Mills field is analysed for the non-local Yang–Mills sector and found to be in agreement with the local results on the lattice in the limit of the non-local scale running to infinity (cf. Table 1). The agreement with the lattice is astonishingly good, as is the agreement with previous computations [94,95]. Because of this, this spectrum becomes our reference for the non-local case. We point out that recent lattice results on the Casimir effect in Yang–Mills theory seem to challenge the determination of the ground state of the theory on the lattice. In our case, this would be a further confirmation of our approach (for details, see Ref. [96]). It is a relevant result of our investigations that scale invariance is badly broken by interactions, a fact that can be taken as a possible clue of confinement. Indeed, the solution of the gap equation for the fermion shows some indication of quark confinement in the non-local case as well. This result is really important, as it seems to point to the fact that confinement could be a ubiquitous effect in nature that removes degrees of freedom in a theory to favour others. A rigorous mathematical proof of confinement is beyond the scope of the present manuscript. In general, further studies are needed to improve these results. Still, the results obtained here appear to be a sound confirmation of previous work with a different technique and with the important extension to non-local QCD.

Author Contributions

Conceptualisation, M.F. and A.G.; methodology, M.F. and S.G.; formal analysis, A.C. and S.G.; writing, M.F., A.G. and S.G.; supervision, M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Estonian Research Council under Grant No. IUT2-27 and by the European Regional Development Fund under Grant No. TK133.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Dyson–Schwinger Equations

In this appendix, we show how to derive the Dyson–Schwinger equations for a ϕ 4 theory in differential form, as exemplified in Refs. [100,101] using the technique devised in Ref. [85] by Bender, Milton and Savage. Such a technique was initially conceived for a PT -invariant non-Hermitian theory and properly extended to more general cases by one of us (M.F.). The starting point is the (massless) Lagrangian
L = 1 2 ( ϕ ) 2 1 4 λ ϕ 4 ,
leading to the generating functional
Z [ j ] = [ d ϕ ] exp i ( L + j ϕ ) d 4 x .
First of all, after integration by parts, the variation in this generating functional with respect to ϕ leads to the Dyson–Schwinger master equation
1 Z [ j ] i δ Z [ j ] δ ϕ ( x ) = 2 ϕ ( x ) j λ ϕ 3 ( x ) j + j ( x ) = 0 ,
where
O ( x ) j = [ d ϕ ] O ( x ) exp i ( L + j ϕ ) d 4 x [ d ϕ ] exp i ( L + j ϕ ) d 4 x | j = 0 = [ d ϕ ] O ( x ) exp i L d 4 x [ d ϕ ] exp i L d 4 x .
Following the notation used, e.g., by Abbott [102], the application of functional derivatives with respect to the current j to the effective action W [ j ] = ln Z [ j ] on the one hand and to the generating functional Z [ j ] on the other hand leads to the tower of Green functions
G 1 j ( x ) = i δ δ j ( x ) W [ j ] = 1 Z [ j ] i δ δ j ( x ) Z [ j ] = ϕ ( x ) j , G 2 j ( x , y ) = i δ δ j ( x ) i δ δ j ( y ) W [ j ] = i δ δ j ( x ) 1 Z [ j ] i δ δ j ( y ) Z [ j ] = 1 Z [ j ] 2 i δ δ j ( x ) Z [ j ] i δ δ j ( y ) Z [ j ] + 1 Z [ j ] i δ δ j ( x ) i δ δ j ( y ) Z [ j ] = ϕ ( x ) j ϕ ( y ) j + ϕ ( x ) ϕ ( y ) j , G 3 j ( x , y , z ) = 2 ϕ ( x ) j ϕ ( y ) j ϕ ( z ) j ϕ ( x ) j ϕ ( y ) ϕ ( z ) j ϕ ( y ) j ϕ ( z ) ϕ ( x ) j ( ( ϕ ( z ) j ϕ ( x ) ϕ ( y ) j + ϕ ( x ) ϕ ( y ) ϕ ( z ) j ,
Inverting this system step by step, one obtains the tower
ϕ ( x ) j = G 1 j ( x ) , ϕ ( x ) ϕ ( y ) j = G 2 j ( x , y ) + G 1 j ( x ) G 1 j ( y ) , ϕ ( x ) ϕ ( y ) ϕ ( z ) j = G 3 j ( x , y , z ) + G 2 j ( x , y ) G 1 j ( z ) + G 2 j ( y , z ) G 1 j ( x ) + G 2 j ( z , x ) G 1 j ( y ) ( ( + G 1 j ( x ) G 1 j ( y ) G 1 j ( z ) ,
Particles 07 00022 i001
This tower is finished by the expectation value of the Dyson–Schwinger master equation, which, via the insertion of the tower, leads to
j ( x ) = 2 ϕ ( x ) j + λ ϕ 3 ( x ) j = 2 G 1 j ( x ) + λ G 3 j ( x , x , x ) + 3 G 2 j ( x , x ) G 1 j ( x ) + G 1 j ( x ) 3 .
In obtaining equations for the Green functions by setting j = 0 , at the same time, we use translation invariance to write the Green functions in the form
G 1 ( x ) : = G 1 j ( x ) , G k ( x 1 x 2 , x 2 x 3 , , x k 1 x k ) : = G k j ( x 1 , x 2 , x 3 , , x k 1 , x k )
to obtain
2 G 1 ( x ) + λ G 3 ( 0 , 0 ) + 3 G 2 ( 0 ) G 1 ( x ) + G 1 ( x ) 3 = 0 .
As G 2 ( 0 ) and G 3 ( 0 , 0 ) are constants, this equation is an ordinary (though non-linear) differential equation for G 1 ( x ) . For 3 λ G 2 ( 0 ) = Δ m G 2 and G 3 ( 0 , 0 ) = 0 , one has 2 G 1 + Δ m G 2 G 1 + λ G 1 3 = 0 . As shown in Ref. [101], this non-linear differential equation is solved by
G 1 ( x ) = μ sn ( k · x + θ | κ ) , μ = 2 ( k 2 Δ m G 2 ) λ
with
k 2 = Δ m G 2 + 1 2 λ μ 2 , κ = Δ m G 2 k 2 k 2 ,
where μ and θ are integration constants, and sn ( ζ | κ ) , cn ( ζ | κ ) and dn ( ζ | κ ) are Jacobi elliptic functions. Even for Δ m G = 0 , i.e., for the absence of a Green function or gap mass, one obtains a nontrivial solution:
G 1 ( x ) = 2 k 2 λ sn ( k · x + θ | 1 ) , k 2 = 1 2 λ μ 2 .
If we take the functional derivative i δ / δ j ( y ) of Equation (A7) before setting j = 0 , one obtains the differential equation
i δ 4 ( x y ) = 2 G 2 j ( x , y ) ( ( + λ G 4 j ( x , x , x , y ) + 3 G 3 j ( x , x , y ) G 1 j ( x ) + 3 G 2 j ( x , x ) G 2 j ( x , y ) + 3 G 1 j ( x ) 2 G 2 j ( x , y ) .
Again, we can choose G 4 j ( x , x , x , y ) = G 3 j ( x , x , y ) = 0 and 3 λ G 2 j ( x , x ) = Δ m G 2 to obtain
2 G 2 j ( x , x ) + Δ m G 2 G 2 j ( x , x ) + 3 λ G 1 ( x ) 2 G 2 j ( x , x ) = i δ 4 ( x x ) .
Note that the Green function defined by this differential equation is not translational-invariant. Therefore, we cannot use G 2 j ( x , x ) = G 2 ( x x ) at this point. However, as shown in Appendix B, we can restore the translational invariance. Inserting Equation (A10), one obtains
2 G 2 j ( x , x ) + Δ m G 2 G 2 j ( x , x ) 6 κ k 2 sn 2 ( k · x + θ | κ ) G 2 j ( x , x ) = i δ 4 ( x x )
(note that λ μ 2 = 2 ( k 2 Δ m G 2 ) = 2 κ k 2 ).
In order to show that our solution for G 1 is meaningful, we show that the theory has a zero mode. The Hamiltonian of the system is given by
H = d 3 x 1 2 π 2 ( x ) + 1 2 ϕ ( x ) 2 + λ 4 ϕ 4 ( x ) .
We expand this Hamiltonian around the classical solution, given by G 1 ( x ) in Equation (A10),
ϕ ( x ) = G 1 ( x ) + δ ϕ ( x ) , π ( x ) = t G 1 ( x ) + t δ ϕ ( x ) ,
yielding
H = H 0 + d 3 x 1 2 t δ ϕ ( x ) 2 + 1 2 δ ϕ ( x ) 2 + 3 2 λ G 1 2 ( x ) δ ϕ 2 ( x ) + O δ ϕ 3 ,
where H 0 is the contribution coming from the classical solution. The linear part vanishes after integration by parts and the application of the equations of motion for the classical solution. The quadratic part can be diagonalised with a Fourier series, provided that we are able to obtain the eigenvalues and the eigenvectors of the operator
L μ 0 2 = 0 = + 3 λ G 1 2 ( x ) .
It is not difficult to realise that there is a zero mode. We give the solutions for both the zero and non-zero modes. The spectrum is continuous with eigenvalues of 0 and 3 μ 2 λ / 2 , where μ varies continuously from zero to infinity. The zero-mode solution has the form
χ 0 ( x , μ ) = a 0 cn ( k · x + θ | 1 ) dn ( k · x + θ | 1 ) ,
where a 0 is a normalisation constant. Non-zero modes are given by
χ ( x , μ ) = a sn ( k · x + θ | 1 ) dn ( k · x + θ | 1 ) ,
where a is again a normalisation constant. These equations hold on-shell, that is, when k 2 = μ 2 λ / 2 . Since the spectrum is continuous, the eigenfunctions are not normalisable. Therefore, we note that there is a doubly degenerate set of zero modes spontaneously breaking translational invariance and the Z 2 symmetry of the theory. For the zero mode, this results in
χ 0 ( x , μ ) = 2 a 0 sn ( k · x | 1 ) dn 2 ( k · x | 1 ) .
For a given parameter μ , Z 2 symmetry is spontaneously broken through this zero mode. The mode disappears when μ = 0 , as it should, and one gets back to a standard textbook solution.

Appendix B. The 2P-Correlation Function for the Scalar Field

Introducing a parameter ε , the differential equation
t 2 G 2 j ( x , x ) + Δ m G 2 G 2 j ( x , x ) 6 κ k 2 sn 2 ( k · x + θ | κ ) G 2 j ( x , x ) = i δ 4 ( x x ) + ε 2 G 2 j ( x , x )
is iteratively solved for G 2 j ( x , x ) by using the gradient expansion in ε ,
G 2 j ( x , x ) = n = 0 ε n G 2 ( n ) ( x , x )
where ε = 1 is set in the end. One obtains
t 2 G 2 ( 0 ) ( x , x ) + Δ m G 2 G 2 ( 0 ) ( x , x ) 6 κ k 2 sn 2 ( k · x + θ | κ ) G 2 ( 0 ) ( x , x ) = i δ 4 ( x x ) , t 2 G 2 ( 1 ) ( x , x ) + Δ m G 2 G 2 ( 1 ) ( x , x ) 6 κ k 2 sn 2 ( k · x + θ | κ ) G 2 ( 1 ) ( x , x ) = 2 G 2 ( 0 ) ( x , x ) , t 2 G 2 ( 2 ) ( x , x ) + Δ m G 2 G 2 ( 2 ) ( x , x ) 6 κ k 2 sn 2 ( k · x + θ | κ ) G 2 ( 2 ) ( x , x ) = 2 G 2 ( 1 ) ( x , x ) ,
In order to perform the calculation, we start with the rest frame of the motion. As the first equation contains no spatial derivative, one has G 2 ( 0 ) ( x , x ) = δ 3 ( x x ) G ¯ 2 ( t , t ) , which simplifies the first differential equation (for the choice x = 0 and, accordingly, x = 0 ) to
G ¯ 2 ( t , t ) + Δ m G 2 G ¯ 2 ( t , t ) 6 κ k 2 sn 2 ( ω t + θ | κ ) G ¯ 2 ( t , t ) = i δ ( t t ) .
The corresponding homogeneous differential equation is solved by G ¯ 2 ( t ) : = G ¯ 2 ( t , t ) = G ^ 2 ( ω t + θ ) , which does not depend on t at all. Therefore, we can set t = 0 in the following. The differential equation for the Green function is then solved by G ¯ 2 ( t ) = C Θ ( t ) G ^ 2 ( ω t + θ ) with the Heaviside step function Θ ( t ) , where the amplitude C and the phase θ get fixed as well, and this also holds if the mass term is not skipped. This can be seen by calculating the derivatives to obtain
G ¯ 2 ( t ) = C Θ ( t ) G ^ 2 ( ω t + θ ) , t G ¯ 2 ( t ) = C δ ( t ) G ^ 2 ( ω t + θ ) + C Θ ( t ) ω G ^ 2 ( ω t + θ ) , t 2 G ¯ 2 ( t ) = C δ ( t ) G ^ 2 ( ω t + θ ) + 2 C δ ( t ) ω G ^ 2 ( ω t + θ ) + C Θ ( t ) ω 2 G ^ 2 ( ω t + θ ) = C δ ( t ) ω G ^ 2 ( ω t + θ ) + C Θ ( t ) ω 2 G ^ 2 ( ω t + θ )
(where we used δ ( t ) f ( t ) = δ ( t ) f ( t ) ) and inserting
C δ ( t ) ω G ^ 2 ( ω t + θ ) ( ( + C Θ ( t ) ω 2 G ^ 2 ( ω t + θ ) + Δ m G 2 G ^ 2 ( ω t + θ ) 6 κ k 2 sn 2 ( ω t + θ | κ ) G ^ 2 ( ω t + θ ) = i δ ( t ) .
For Δ m G = 0 , the homogeneous equation is solved by G ^ 2 ( ζ ) = G ^ 1 ( ζ ) with G ^ 1 ( ζ ) = sn ( ζ | 1 ) . However, using the more general ansatz G ^ 1 ( ζ ) = sn ( ζ | κ ) , this also holds for Δ m G 0 . Namely, one obtains G ^ 2 ( ζ ) = dn ( ζ | κ ) cn ( ζ | κ ) ,
G ^ 2 ( ζ ) = sn ( ζ | κ ) dn 2 ( ζ | κ ) + κ cn 2 ( ζ | κ ) , G ^ 2 ( ζ ) = dn ( ζ | κ ) cn ( ζ | κ ) dn 2 ( ζ | κ ) + κ cn 2 ( ζ | κ ) 4 κ sn 2 ( ζ | κ ) .
Therefore, with ζ = ω t + θ and ω 2 = k 2 , for the rest frame, one obtains
ω 2 G ^ 2 ( ζ ) + Δ m G 2 G ^ 2 ( ζ ) 6 κ k 2 sn 2 ( ζ | κ ) G ^ 2 ( ζ ) = = k 2 dn 2 ( ζ | κ ) + κ cn 2 ( ζ | κ ) 4 κ sn 2 ( ζ | κ ) + Δ m G 2 6 κ k 2 sn 2 ( ζ | κ ) G ^ 2 ( ζ ) = k 2 ( 1 + κ ) + Δ m G 2 G ^ 2 ( ζ ) = 0 ,
where we have used cn 2 ( ζ | κ ) + sn 2 ( ζ | κ ) = 1 and dn 2 ( ζ | κ ) + κ 2 sn 2 ( ζ | κ ) = 1 . One can start with sn ( θ | κ ) = 1 , which is satisfied if θ = ( 1 + 4 N ) K ( κ ) , and as a consequence of this, C ω ( 1 κ ) = i , which is solved by C = i / ω ( 1 κ ) . Therefore, we end up with
G ¯ 2 ( t ) = i Θ ( t ) ω ( 1 κ ) sn ( ω t + θ | κ ) , θ = ( 1 + 4 N ) K ( κ ) .
As Jacobi’s elliptic functions sn ( ζ | κ ) , cn ( ζ | κ ) and dn ( ζ | κ ) are periodic functions, it is possible to expand them in a Fourier series. Using the nome q = exp ( π K * ( κ ) / K ( κ ) ) , where K * ( κ ) = K ( 1 κ ) and
K ( κ ) = 0 π / 2 d θ 1 κ sin 2 θ = F ( π / 2 | κ ) ,
one has
sn ( ζ | κ ) = 2 π K ( κ ) κ n = 0 q n + 1 / 2 1 q 2 n + 1 sin ( 2 n + 1 ) π ζ 2 K ( κ ) .
Inserting this Fourier series into our solution for G ¯ 2 ( t ) leads to
G ¯ 2 ( t ) = i Θ ( t ) π 2 ω ( 1 κ ) K ( κ ) 2 κ n = 0 ( 2 n + 1 ) q n + 1 / 2 1 q 2 n + 1 cos ( 2 n + 1 ) π ( ω t + θ ) 2 K ( κ )
with θ = ( 1 + 4 N ) K ( κ ) . Inserting these values for θ , one has
cos ( 2 n + 1 ) π ω t 2 K ( κ ) + π 2 + 2 π N = ( 1 ) n sin ( 2 n + 1 ) π ω t 2 K ( κ ) .
Our result for the Fourier series of the zeroth-order Green function is therefore given by
G ¯ 2 ( t ) = i Θ ( t ) π 2 ω ( 1 κ ) K ( κ ) 2 κ n = 0 ( 2 n + 1 ) ( 1 ) n q n + 1 / 2 1 q 2 n + 1 sin ( 2 n + 1 ) π ω t 2 K ( κ ) .
We can define mass states m n = ( 2 n + 1 ) m 0 with m 0 = π ω / ( 2 K ( κ ) ) , forming the spectrum of the gluonic part of the QCD Lagrangian, i.e., the spectrum of glueballs. The Fourier transform leads to terms of the kind
Θ ( t ) sin ( m n t ) e i E t d t = 0 1 2 i e i m n t e i m n t e i E t d t = 1 2 i e i ( E m n ) t i ( E m n ) e i ( E + m n ) t i ( E + m n ) t = 0 = 1 2 1 E m n 1 E + m n = m n E 2 m n 2 .
Therefore, in using the Feynman propagator convention, we end up with
G ¯ ˜ 2 ( E ) = i n = 0 B n ( 1 + Δ m G 2 / k 2 ) E 2 m n 2 + i ϵ , B n ( κ ) : = ( 2 n + 1 ) 2 π 3 2 ( 1 κ ) K ( κ ) 3 κ ( 1 ) n q n + 1 / 2 1 q 2 n + 1 .
In the massless case, i.e., κ = 1 , the situation is simplified to K ( 1 ) = 1.31103 and K * ( 1 ) = K ( 2 ) = 1.31103 ( 1 i ) . Therefore, q = e π ( 1 i ) = e π . For the Fourier coefficient, one obtains
e ( n + 1 / 2 ) π ( 1 i ) 1 e ( 2 n + 1 ) π ( 1 i ) = i ( 1 ) n e ( n + 1 / 2 ) π 1 + e ( 2 n + 1 ) π ,
and the factor i cancels against κ = 1 = i in the denominator. Therefore, one ends up with
B n ( 1 ) : = ( 2 n + 1 ) 2 π 3 4 K ( 1 ) 3 e ( 2 n + 1 ) π / 2 1 + e ( 2 n + 1 ) π .
Conversely, one has
G ¯ 2 ( t t ) = d p 0 2 π G ¯ ˜ 2 ( p 0 ) e i p 0 ( t t )
and
G 2 ( 0 ) ( x , x ) = δ ( 3 ) ( x x ) d p 0 2 π G ¯ ˜ 2 ( p 0 ) e i p 0 ( t t ) = d 4 p ( 2 π ) 4 G ¯ ˜ 2 ( p 0 ) e i p ( x x ) .
Having G 2 ( 0 ) ( x , x ) at hand, the system (A25) can be solved iteratively by
G 2 ( 1 ) ( x , x ) = i G 2 ( 0 ) ( x , x 1 ) 2 G 2 ( 0 ) ( x 1 , x ) d 4 x 1 , G 2 ( 2 ) ( x , x ) = i G 2 ( 0 ) ( x , x 2 ) 2 G 2 ( 1 ) ( x 2 , x ) d 4 x 2 = = G 2 ( 0 ) ( x , x 2 ) 2 G 2 ( 0 ) ( x 2 , x 1 ) 2 G 2 ( 0 ) ( x 1 , x ) d 4 x 1 d 4 x 2 ,
and G 2 ( x , x ) , the sum of all of these (for ε = 1 ). In momentum space, this is just a Dyson series, where the propagators are given by Equation (A38), and the vertices can be derived from
2 δ ( x x ) = 2 d 3 p ( 2 π ) 3 e i p ( x x ) = d 3 p ( 2 π ) 3 ( p 2   ) e i p ( x x )
to be p 2   in momentum space. The Dyson series can be resummed to [100].
G ˜ 2 ( E ; p ) = n = 0 i B n E 2 m n 2 + i ϵ + n = 0 i B n E 2 m n 2 + i ϵ ( i p 2   ) m = 0 i B m E 2 m m 2 + i ϵ ( ( + n = 0 i B n E 2 m n 2 + i ϵ ( i p 2   ) m = 0 i B m E 2 m m 2 + i ϵ ( i p 2   ) l = 0 i B l E 2 m l 2 + i ϵ + = n = 0 i B n E 2 m n 2 + i ϵ 1 m = 0 p 2   B m E 2 m m 2 + i ϵ 1 = n = 0 i B n E 2 m n 2 + i ϵ p 2   ( E 2 m n 2 + i ϵ ) m = 0 B m E 2 m m 2 + i ϵ 1 .
Because of the sum rule, this restores the Lorentz invariance of the Green function for high energies E m 0 = π ω / ( 2 K ( κ ) ) ,
G ˜ 2 ( E ; p ) = n = 0 i B n E 2 p 2   m n 2 + i ϵ = n = 0 i B n p 2 m n 2 + i ϵ = G ˜ 2 ( p )
in momentum space and, therefore, the translational invariance in configuration space. Finally, we consider the correlation function for scalar glueballs that is given by [103,104]
O ( x ) = F a μ ν ( x ) F μ ν a ( x ) F b ρ η ( 0 ) F ρ η b ( 0 )
and whose poles are the physical glueballs. Using the technique explained above and explained in Ref. [101], one can see that, according to Ref. [105], the four-point correlator (A47), defining the correlation function of the glueball, can be reduced to integrated products of one- and two-point functions. As the one-point function has no poles but zeros, the poles of the glueball four-point correlator (A47) are given by the poles of the two-point correlator. Therefore, these poles represent true colourless glueball states. Identifying the lowest glueball mass state m 0 with the σ resonance f 0 ( 500 ) , one can fix the scale k 2 to be k 2 = ω = 2 m 0 K ( 1 ) / π 417 MeV .

References

  1. Witten, E. Noncommutative Geometry and String Field Theory. Nucl. Phys. B 1986, 268, 253–294. [Google Scholar] [CrossRef]
  2. Freund, P.G.O.; Olson, M. Nonarchimedean Strings. Phys. Lett. B 1987, 199, 186–190. [Google Scholar] [CrossRef]
  3. Freund, P.G.O.; Witten, E. Adelic String Amplitudes. Phys. Lett. B 1987, 199, 191. [Google Scholar] [CrossRef]
  4. Brekke, L.; Freund, P.G.O.; Olson, M.; Witten, E. Nonarchimedean String Dynamics. Nucl. Phys. B 1988, 302, 365–402. [Google Scholar] [CrossRef]
  5. Frampton, P.H.; Okada, Y. Effective Scalar Field Theory of P-adic String. Phys. Rev. D 1988, 37, 3077–3079. [Google Scholar] [CrossRef]
  6. Kostelecky, V.A.; Samuel, S. The Static Tachyon Potential in the Open Bosonic String Theory. Phys. Lett. B 1988, 207, 169–173. [Google Scholar] [CrossRef]
  7. Kostelecky, V.A.; Samuel, S. On a Nonperturbative Vacuum for the Open Bosonic String. Nucl. Phys. B 1990, 336, 263–296. [Google Scholar] [CrossRef]
  8. Tseytlin, A.A. On singularities of spherically symmetric backgrounds in string theory. Phys. Lett. B 1995, 363, 223–229. [Google Scholar] [CrossRef]
  9. Siegel, W. Stringy gravity at short distances. arXiv 2003, arXiv:hep-th/0309093. [Google Scholar]
  10. Biswas, T.; Grisaru, M.; Siegel, W. Linear Regge trajectories from worldsheet lattice parton field theory. Nucl. Phys. B 2005, 708, 317–344. [Google Scholar] [CrossRef]
  11. Calcagni, G.; Modesto, L. Nonlocality in string theory. J. Phys. A 2014, 47, 355402. [Google Scholar] [CrossRef]
  12. Calcagni, G.; Modesto, L. Nonlocal quantum gravity and M-theory. Phys. Rev. D 2015, 91, 124059. [Google Scholar] [CrossRef]
  13. Taylor, W.; Zwiebach, B. D-branes, tachyons, and string field theory. arXiv, 2004; arXiv:hep-th/0311017. [Google Scholar]
  14. Moffat, J.W. Finite nonlocal gauge field theory. Phys. Rev. D 1990, 41, 1177–1184. [Google Scholar] [CrossRef]
  15. Evens, D.; Moffat, J.W.; Kleppe, G.; Woodard, R.P. Nonlocal regularizations of gauge theories. Phys. Rev. D 1991, 43, 499–519. [Google Scholar] [CrossRef]
  16. Kleppe, G.; Woodard, R.P. Nonlocal Yang-Mills. Nucl. Phys. B 1992, 388, 81–112. [Google Scholar] [CrossRef]
  17. Tomboulis, E.T. Superrenormalizable gauge and gravitational theories. arXiv 1997, arXiv:hep-th/9702146. [Google Scholar]
  18. Clayton, M.A. Nonlocal regularization of Abelian models with spontaneous symmetry breaking. Mod. Phys. Lett. A 2001, 16, 1117–1127. [Google Scholar] [CrossRef]
  19. Biswas, T.; Gerwick, E.; Koivisto, T.; Mazumdar, A. Towards singularity and ghost free theories of gravity. Phys. Rev. Lett. 2012, 108, 031101. [Google Scholar] [CrossRef]
  20. Modesto, L. Super-renormalizable Quantum Gravity. Phys. Rev. D 2012, 86, 044005. [Google Scholar] [CrossRef]
  21. Modesto, L. Super-renormalizable Higher-Derivative Quantum Gravity. arXiv 2012, arXiv:1202.0008. [Google Scholar]
  22. Modesto, L. Super-renormalizable Multidimensional Quantum Gravity. Astron. Rev. 2013, 8, 4–33. [Google Scholar]
  23. Modesto, L. Finite Quantum Gravity. arXiv 2013, arXiv:1305.6741. [Google Scholar]
  24. Biswas, T.; Conroy, A.; Koshelev, A.S.; Mazumdar, A. Generalized ghost-free quadratic curvature gravity. Class. Quant. Grav. 2014, 31, 015022, Erratum in Class. Quant. Grav. 2014, 31, 159501. [Google Scholar] [CrossRef]
  25. Modesto, L. Multidimensional finite quantum gravity. arXiv 2014, arXiv:1402.6795. [Google Scholar]
  26. Modesto, L.; Rachwał, L. Super-renormalizable and finite gravitational theories. Nucl. Phys. B 2014, 889, 228–248. [Google Scholar] [CrossRef]
  27. Talaganis, S.; Biswas, T.; Mazumdar, A. Towards understanding the ultraviolet behavior of quantum loops in infinite-derivative theories of gravity. Class. Quant. Grav. 2015, 32, 215017. [Google Scholar] [CrossRef]
  28. Modesto, L.; Rachwał, L. Universally finite gravitational and gauge theories. Nucl. Phys. B 2015, 900, 147–169. [Google Scholar] [CrossRef]
  29. Modesto, L.; Piva, M.; Rachwał, L. Finite quantum gauge theories. Phys. Rev. D 2016, 94, 025021. [Google Scholar] [CrossRef]
  30. Tomboulis, E.T. Nonlocal and quasilocal field theories. Phys. Rev. D 2015, 92, 125037. [Google Scholar] [CrossRef]
  31. Modesto, L.; Rachwał, L.; Shapiro, I.L. Renormalization group in super-renormalizable quantum gravity. Eur. Phys. J. C 2018, 78, 555. [Google Scholar] [CrossRef]
  32. Hashi, M.N.; Isono, H.; Noumi, T.; Shiu, G.; Soler, P. Higgs Mechanism in Nonlocal Field Theories. JHEP 2018, 8, 64. [Google Scholar] [CrossRef]
  33. Buoninfante, L.; Lambiase, G.; Mazumdar, A. Ghost-free infinite derivative quantum field theory. Nucl. Phys. B 2019, 944, 114646. [Google Scholar] [CrossRef]
  34. Modesto, L. Nonlocal Spacetime-Matter. arXiv 2021, arXiv:2103.04936. [Google Scholar]
  35. Modesto, L. The Higgs mechanism in nonlocal field theory. JHEP 2021, 6, 49. [Google Scholar] [CrossRef]
  36. Boos, J.; Carone, C.D. Asymptotic nonlocality. Phys. Rev. D 2021, 104, 015028. [Google Scholar] [CrossRef]
  37. Boos, J.; Carone, C.D. Asymptotic nonlocality in gauge theories. Phys. Rev. D 2021, 104, 095020. [Google Scholar] [CrossRef]
  38. Boos, J.; Carone, C.D. Asymptotic nonlocality in non-Abelian gauge theories. Phys. Rev. D 2022, 105, 035034. [Google Scholar] [CrossRef]
  39. Boos, J.; Carone, C.D. Asymptotically nonlocal gravity. arXiv 2022, arXiv:2212.00861. [Google Scholar] [CrossRef]
  40. Krasnikov, N.V. Nonlocal Gauge Theories. Theor. Math. Phys. 1987, 73, 1184–1190. [Google Scholar] [CrossRef]
  41. Biswas, T.; Okada, N. Towards LHC physics with nonlocal Standard Model. Nucl. Phys. B 2015, 898, 113–131. [Google Scholar] [CrossRef]
  42. Pius, R.; Sen, A. Cutkosky rules for superstring field theory. JHEP 2016, 10, 024, Erratum in JHEP 2018, 9, 122. [Google Scholar] [CrossRef]
  43. Briscese, F.; Modesto, L. Cutkosky rules and perturbative unitarity in Euclidean nonlocal quantum field theories. Phys. Rev. D 2019, 99, 104043. [Google Scholar] [CrossRef]
  44. Chin, P.; Tomboulis, E.T. Nonlocal vertices and analyticity: Landau equations and general Cutkosky rule. JHEP 2018, 6, 014. [Google Scholar] [CrossRef]
  45. Briscese, F.; Modesto, L. Non-unitarity of Minkowskian non-local quantum field theories. Eur. Phys. J. C 2021, 81, 730. [Google Scholar] [CrossRef]
  46. Koshelev, A.S.; Tokareva, A. Unitarity of Minkowski nonlocal theories made explicit. Phys. Rev. D 2021, 104, 025016. [Google Scholar] [CrossRef]
  47. Ghoshal, A.; Mazumdar, A.; Okada, N.; Villalba, D. Stability of infinite derivative Abelian Higgs models. Phys. Rev. D 2018, 97, 076011. [Google Scholar] [CrossRef]
  48. Ghoshal, A.; Mazumdar, A.; Okada, N.; Villalba, D. Nonlocal non-Abelian gauge theory: Conformal invariance and β-function. Phys. Rev. D 2021, 104, 015003. [Google Scholar] [CrossRef]
  49. Buoninfante, L.; Ghoshal, A.; Lambiase, G.; Mazumdar, A. Transmutation of nonlocal scale in infinite derivative field theories. Phys. Rev. D 2019, 99, 044032. [Google Scholar] [CrossRef]
  50. Ghoshal, A. Scalar dark matter probes the scale of nonlocality. Int. J. Mod. Phys. A 2019, 34, 1950130. [Google Scholar] [CrossRef]
  51. Frasca, M.; Ghoshal, A.; Koshelev, A.S. Quintessence dark energy from strongly-coupled higgs mass gap: Local and non-local higher-derivative non-perturbative scenarios. Eur. Phys. J. C 2022, 82, 1108. [Google Scholar] [CrossRef]
  52. Krasnikov, N.V. Nonlocal SU(5) GUT. Mod. Phys. Lett. A 2021, 36, 2150104. [Google Scholar] [CrossRef]
  53. Nortier, F. Extra Dimensions and Fuzzy Branes in String-inspired Nonlocal Field Theory. Acta Phys. Polon. B 2023, 54, 6-A2. [Google Scholar] [CrossRef]
  54. Frasca, M.; Ghoshal, A. Diluted mass gap in strongly coupled non-local Yang-Mills. JHEP 2020, 21, 226. [Google Scholar] [CrossRef]
  55. Frasca, M.; Ghoshal, A.; Okada, N. Confinement and renormalization group equations in string-inspired nonlocal gauge theories. Phys. Rev. D 2021, 104, 096010. [Google Scholar] [CrossRef]
  56. Frasca, M.; Ghoshal, A. Mass gap in strongly coupled infinite derivative non-local Higgs: Dyson–Schwinger approach. Class. Quant. Grav. 2021, 38, 17. [Google Scholar] [CrossRef]
  57. Frasca, M.; Ghoshal, A.; Koshelev, A.S. Non-perturbative Lee-Wick gauge theory: Towards Confinement & RGE with strong couplings. Class. Quant. Grav. 2024, 41, 015014. [Google Scholar]
  58. Frasca, M.; Ghoshal, A.; Groote, S. Confinement in QCD and generic Yang-Mills theories with matter representations. Phys. Lett. B 2023, 846, 138209. [Google Scholar] [CrossRef]
  59. Frasca, M.; Ghoshal, A.; Groote, S. Quark confinement in QCD in the ’t Hooft limit. Nucl. Part. Phys. Proc. 2023, 324–329, 85–89. [Google Scholar] [CrossRef]
  60. Frolov, V.P.; Zelnikov, A.; de Paula Netto, T. Spherical collapse of small masses in the ghost-free gravity. JHEP 2015, 6, 107. [Google Scholar] [CrossRef]
  61. Frolov, V.P.; Zelnikov, A. Head-on collision of ultrarelativistic particles in ghost-free theories of gravity. Phys. Rev. D 2016, 93, 064048. [Google Scholar] [CrossRef]
  62. Koshelev, A.S.; Marto, J.; Mazumdar, A. Schwarzschild 1/r-singularity is not permissible in ghost free quadratic curvature infinite derivative gravity. Phys. Rev. D 2018, 98, 064023. [Google Scholar] [CrossRef]
  63. Koshelev, A.S.; Mazumdar, A. Do massive compact objects without event horizon exist in infinite derivative gravity? Phys. Rev. D 2017, 96, 084069. [Google Scholar] [CrossRef]
  64. Buoninfante, L.; Koshelev, A.S.; Lambiase, G.; Mazumdar, A. Classical properties of non-local, ghost- and singularity-free gravity. JCAP 2018, 9, 34. [Google Scholar] [CrossRef]
  65. Cornell, A.S.; Harmsen, G.; Lambiase, G.; Mazumdar, A. Rotating metric in nonsingular infinite derivative theories of gravity. Phys. Rev. D 2018, 97, 104006. [Google Scholar] [CrossRef]
  66. Buoninfante, L.; Koshelev, A.S.; Lambiase, G.; Marto, J.; Mazumdar, A. Conformally-flat, non-singular static metric in infinite derivative gravity. JCAP 2018, 6, 14. [Google Scholar] [CrossRef]
  67. Buoninfante, L.; Harmsen, G.; Maheshwari, S.; Mazumdar, A. Nonsingular metric for an electrically charged point-source in ghost-free infinite derivative gravity. Phys. Rev. D 2018, 98, 084009. [Google Scholar] [CrossRef]
  68. Abel, S.; Buoninfante, L.; Mazumdar, A. Nonlocal gravity with worldline inversion symmetry. JHEP 2020, 1, 3. [Google Scholar] [CrossRef]
  69. Buoninfante, L.; Lambiase, G.; Miyashita, Y.; Takebe, W.; Yamaguchi, M. Generalized ghost-free propagators in nonlocal field theories. Phys. Rev. D 2020, 101, 084019. [Google Scholar] [CrossRef]
  70. Biswas, T.; Mazumdar, A.; Siegel, W. Bouncing universes in string-inspired gravity. JCAP 2006, 3, 9. [Google Scholar] [CrossRef]
  71. Biswas, T.; Brandenberger, R.; Mazumdar, A.; Siegel, W. Non-perturbative Gravity, Hagedorn Bounce & CMB. JCAP 2007, 12, 11. [Google Scholar]
  72. Biswas, T.; Koivisto, T.; Mazumdar, A. Towards a resolution of the cosmological singularity in non-local higher derivative theories of gravity. JCAP 2010, 11, 8. [Google Scholar] [CrossRef]
  73. Biswas, T.; Koshelev, A.S.; Mazumdar, A.; Vernov, S.Y. Stable bounce and inflation in non-local higher derivative cosmology. JCAP 2012, 8, 24. [Google Scholar] [CrossRef]
  74. Koshelev, A.S.; Vernov, S.Y. On bouncing solutions in non-local gravity. Phys. Part. Nucl. 2012, 43, 666–668. [Google Scholar] [CrossRef]
  75. Koshelev, A.S.; Marto, J.; Mazumdar, A. Towards resolution of anisotropic cosmological singularity in infinite derivative gravity. JCAP 2019, 2, 20. [Google Scholar] [CrossRef]
  76. Kolář, I.; Torralba, F.J.M.; Mazumdar, A. New nonsingular cosmological solution of nonlocal gravity. Phys. Rev. D 2022, 105, 044045. [Google Scholar] [CrossRef]
  77. Ghoshal, A.; Nortier, F. Fate of the false vacuum in string-inspired nonlocal field theory. JCAP 2022, 8, 47. [Google Scholar] [CrossRef]
  78. Boos, J.; Frolov, V.P.; Zelnikov, A. Gravitational field of static p-branes in linearized ghost-free gravity. Phys. Rev. D 2018, 97, 084021. [Google Scholar] [CrossRef]
  79. Boos, J. Gravitational Friedel oscillations in higher-derivative and infinite-derivative gravity? Int. J. Mod. Phys. D 2018, 27, 1847022. [Google Scholar] [CrossRef]
  80. Boos, J. Angle deficit and nonlocal gravitoelectromagnetism around a slowly spinning cosmic string. Int. J. Mod. Phys. D 2020, 29, 2043027. [Google Scholar] [CrossRef]
  81. Boos, J. Effects of Non-locality in Gravity and Quantum Theory. arXiv 2020, arXiv:2009.10856. [Google Scholar]
  82. Lee, T.D.; Wick, G.C. Negative Metric and the Unitarity of the S Matrix. Nucl. Phys. B 1969, 9, 209–243. [Google Scholar] [CrossRef]
  83. Lee, T.D.; Wick, G.C. Finite Theory of Quantum Electrodynamics. Phys. Rev. D 1970, 2, 1033–1048. [Google Scholar] [CrossRef]
  84. Gabrielli, E. On the dynamical breaking of chiral symmetry: A New mechanism. Phys. Rev. D 2008, 77, 055020. [Google Scholar] [CrossRef]
  85. Bender, C.M.; Milton, K.A.; Savage, V.M. Solution of Schwinger-Dyson equations for PT symmetric quantum field theory. Phys. Rev. D 2000, 62, 085001. [Google Scholar] [CrossRef]
  86. Frasca, M. Infrared Gluon and Ghost Propagators. Phys. Lett. B 2008, 670, 73–77. [Google Scholar] [CrossRef]
  87. Frasca, M. Mapping a Massless Scalar Field Theory on a Yang-Mills Theory: Classical Case. Mod. Phys. Lett. A 2009, 24, 2425–2432. [Google Scholar] [CrossRef]
  88. Smilga, A.V. Lectures on Quantum Chromodynamics; World Scientific: Singapore, 2001. [Google Scholar]
  89. Matinyan, S.G.; Savvidy, G.K.; Ter-Arutunian, S.N.G. Stochasticity of Classical Yang-Mills Mechanics and Its Elimination by Higgs Mechanism. JETP Lett. 1981, 34, 590–593. (In Russian) [Google Scholar]
  90. Savvidy, G.K. Yang–Mills Classical Mechanics as a Kolmogorov K System. Phys. Lett. B 1983, 130, 303–307. [Google Scholar] [CrossRef]
  91. Savvidy, G.K. Classical and Quantum Mechanics of Nonabelian Gauge Fields. Nucl. Phys. B 1984, 246, 302–334. [Google Scholar] [CrossRef]
  92. Athenodorou, A.; Teper, M. SU(N) gauge theories in 3+1 dimensions: Glueball spectrum, string tensions and topology. JHEP 2021, 12, 82. [Google Scholar] [CrossRef]
  93. Lucini, B.; Teper, M.; Wenger, U. Glueballs and k-strings in SU(N) gauge theories: Calculations with improved operators. JHEP 2004, 6, 12. [Google Scholar] [CrossRef]
  94. Frasca, M. Spectrum of a four-dimensional Yang-Mills theory. arXiv 2017, arXiv:1705.07647. [Google Scholar]
  95. Frasca, M. Spectrum of Yang-Mills theory in 3 and 4 dimensions. Nucl. Part. Phys. Proc. 2018, 294–296, 124. [Google Scholar] [CrossRef]
  96. Chernodub, M.N.; Goy, V.A.; Molochkov, A.V.; Tanashkin, A.S. Boundary states and Non-Abelian Casimir effect in lattice Yang-Mills theory. Phys. Rev. D 2023, 108, 014515. [Google Scholar] [CrossRef]
  97. Frasca, M. Casimir effect in Yang-Mills theories. arXiv 2011, arXiv:1108.6299. [Google Scholar]
  98. Hell, T.; Rößner, S.; Cristoforetti, M.; Weise, W. Dynamics and thermodynamics of a non-local Polyakov–Nambu–Jona-Lasinio model with running coupling. Phys. Rev. D 2009, 79, 014022. [Google Scholar] [CrossRef]
  99. Nambu, Y.; Jona-Lasinio, G. Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. 1. Phys. Rev. 1961, 122, 345. [Google Scholar] [CrossRef]
  100. Frasca, M. Scalar field theory in the strong self-interaction limit. Eur. Phys. J. C 2014, 74, 2929. [Google Scholar] [CrossRef]
  101. Frasca, M. Quantum Yang-Mills field theory. Eur. Phys. J. Plus 2017, 132, 38, Erratum in Eur. Phys. J. Plus 2017, 132E, 242. [Google Scholar] [CrossRef]
  102. Abbott, L.F. Introduction to the Background Field Method. Acta Phys. Polon. B 1982, 13, 33. [Google Scholar]
  103. Narison, S. QCD as a Theory of Hadrons: From Partons to Confinement. Camb. Monogr. Part. Phys. Nucl. Phys. Cosmol. 2007, 17, 1–812. [Google Scholar]
  104. Narison, S. Di-gluonium sum rules, I=0 scalar mesons and conformal anomaly. Nucl. Phys. A 2022, 1017, 122337. [Google Scholar] [CrossRef]
  105. Windisch, A.; Huber, M.Q.; Alkofer, R. On the analytic structure of scalar glueball operators at the Born level. Phys. Rev. D 2013, 87, 065005. [Google Scholar] [CrossRef]
Figure 1. The solution of the mass gap Equation (37) for different values of N c in relation to the non-local scale M 2 . A second pair of curves displays the dependence of the dynamically determined parameter −10,000/ κ on M 2 , indicating a nontrivial solution of the gap equation for κ 1 .
Figure 1. The solution of the mass gap Equation (37) for different values of N c in relation to the non-local scale M 2 . A second pair of curves displays the dependence of the dynamically determined parameter −10,000/ κ on M 2 , indicating a nontrivial solution of the gap equation for κ 1 .
Particles 07 00022 g001
Figure 2. The result of the mass gap determination for the quark for different values of N c in relation to the non-local scale M 2 . The number of flavours is set to N f = 2 .
Figure 2. The result of the mass gap determination for the quark for different values of N c in relation to the non-local scale M 2 . The number of flavours is set to N f = 2 .
Particles 07 00022 g002
Figure 3. The quark–glueball mass ratio M q / M G in relation to M 2 , as read off from Figure 1 and Figure 2. Shown is the threshold value M q / M G = 0.39 .
Figure 3. The quark–glueball mass ratio M q / M G in relation to M 2 , as read off from Figure 1 and Figure 2. Shown is the threshold value M q / M G = 0.39 .
Particles 07 00022 g003
Table 1. Values for the dynamical glueball masses M G lat on the lattice in units of k 2 for different values of the number N c of colours (and corresponding values of β ), compared to our estimates M G est from the solution of the mass gap Equation (37) in the local limit M .
Table 1. Values for the dynamical glueball masses M G lat on the lattice in units of k 2 for different values of the number N c of colours (and corresponding values of β ), compared to our estimates M G est from the solution of the mass gap Equation (37) in the local limit M .
N c β M G lat / k 2 M G est / k 2 Error
2 2.427 3.781 ( 23 ) 3.56137 5.8 %
3 6.235 3.405 ( 21 ) 3.25530 4.4 %
4 11.02 3.271 ( 27 ) 3.26827 0.1 %
5 17.61 3.156 ( 31 ) 3.21857 + 2.0 %
6 25.35 3.102 ( 32 ) 3.21930 + 3.8 %
8 45.50 3.099 ( 26 ) 3.19833 + 3.2 %
10 71.38 3.102 ( 37 ) 3.18957 + 2.8 %
12 103.03 3.156 ( 33 ) 3.18443 + 1.1 %
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chatterjee, A.; Frasca, M.; Ghoshal, A.; Groote, S. Renormalisable Non-Local Quark–Gluon Interaction: Mass Gap, Chiral Symmetry Breaking and Scale Invariance. Particles 2024, 7, 392-415. https://doi.org/10.3390/particles7020022

AMA Style

Chatterjee A, Frasca M, Ghoshal A, Groote S. Renormalisable Non-Local Quark–Gluon Interaction: Mass Gap, Chiral Symmetry Breaking and Scale Invariance. Particles. 2024; 7(2):392-415. https://doi.org/10.3390/particles7020022

Chicago/Turabian Style

Chatterjee, Arpan, Marco Frasca, Anish Ghoshal, and Stefan Groote. 2024. "Renormalisable Non-Local Quark–Gluon Interaction: Mass Gap, Chiral Symmetry Breaking and Scale Invariance" Particles 7, no. 2: 392-415. https://doi.org/10.3390/particles7020022

Article Metrics

Back to TopTop