This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Confining potential under the gauge field condensation in the SU(2) Yang-Mills theory

Hirohumi Sawayanagi National Institute of Technology, Kushiro College, Kushiro, 084-0916, Japan
 E-mail:[email protected]
Abstract

QQ¯Q\bar{Q} potential is studied in the SU(2) gauge theory. Based on the nonlinear gauge of the Curci-Ferrari type, the possibility of a gluon condensation Aμ+Aμ\langle A_{\mu}^{+}A_{\mu}^{-}\rangle in low-energy region has been considered at the one-loop level. Instead of the magnetic monopole condensation, this condensation makes classical gluons massive, and can yield a linear potential. We show this potential consists of the Coulomb plus linear part and an additional part. Comparing with the Cornell potential, we study this confining potential in detail, and find that the potential has two implicit scales rcr_{c} and R~0\tilde{R}_{0}. The meanings of these scales are clarified. We also show that the Cornell potential that fits well to this confining potential is obtained by taking these scales into account.

\subjectindex

B0,B3,B6

1 Introduction

In the study of quarkonia, QCD potential is often used. Although there are some phenomenological potentials (see, e.g., bali ), the Cornell potential VCL(r)V_{CL}(r) cor ; cor2 is simple but workable. This potential has the Coulomb plus linear form as

VCL(r)=Kr+σr,V_{CL}(r)=-\frac{K}{r}+\sigma r, (1.1)

where KK and σ\sigma, that is called the string tension, are constants. The Coulomb part is expected from the perturbative one-gluon exchange, and the linear part represents the confinement.

Is it possible to derive VCL(r)V_{CL}(r) from QCD? Using the dual Ginzburg-Landau model (see, e.g., rip ), the following Yukawa plus linear potential was obtained suz ; mts ; sst ; sst2 :

VYL(r)=Q24πemrr+(Q2m28πlnm2+mχ2m2)r,V_{YL}(r)=-\frac{Q^{2}}{4\pi}\frac{e^{-mr}}{r}+\left(\frac{Q^{2}m^{2}}{8\pi}\ln\frac{m^{2}+m_{\chi}^{2}}{m^{2}}\right)r, (1.2)

where QQ is the static quark charge and mχm_{\chi} is the momentum cut-off. In this model, the mass mm is related to the vacuum expectation value (VEV) of the monopole field. In Ref. hs , based on the SU(2) gauge theory in the non-linear gauge of the Curci-Ferrari type, we also derived the potential VYL(r)V_{YL}(r). In this case, the mass mm comes from the gauge field condensation Aμ+Aμ\langle A_{\mu}^{+}A_{\mu}^{-}\rangle.

In this paper, in the framework of Refs. hs ; hs1 , we restudy the confining potential. In the next section, we briefly review Refs. hs ; hs1 , and present the potential between the static charges QQ and Q¯\bar{Q}. In Sect. 3, the equation to determine an ultraviolet cut-off Λc\Lambda_{c} is derived. In Sect. 4, using this cut-off, we show that the QQ¯Q\bar{Q} potential becomes the confining potential Vc(r)=VCL(r)+V3(r)V_{c}(r)=V_{CL}(r)+V_{3}(r), where V3(r)V_{3}(r) is the additional potential. The potential Vc(r)V_{c}(r) has several parameters. Comparing Vc(r)V_{c}(r) with VCL(r)V_{CL}(r), and choosing the appropriate values of KK and σ\sigma, the parameters in Vc(r)V_{c}(r) are determined in Sect. 5. In this process, we find a scale rc0.2fmr_{c}\approx 0.2\ \mathrm{fm}. In the intermediate region, the scale R00.5fmR_{0}\approx 0.5\ \mathrm{fm} has been proposed so . The meanings of the scales rcr_{c} and R0R_{0} for Vc(r)V_{c}(r) are clarified in Sect. 6. We also propose a scale R~0\tilde{R}_{0} that is related to R0R_{0}. Based on this analysis, we obtain VCL(r)V_{CL}(r) that fits well to Vc(r)V_{c}(r). Section 7 is devoted to a summary and comments. In Appendix A, the propagator for the off-diagonal gluons is presented. The equations in Sect. 5, that determine the values of the parameters in Vc(r)V_{c}(r), are solved in Appendix B.

2 Condensate Aμ+Aμ\langle A_{\mu}^{+}A_{\mu}^{-}\rangle and QQ¯Q\bar{Q} potential

In this section, we review Refs. hs ; hs1 briefly.

2.1 Ghost condensation

We consider the SU(2) gauge theory in Euclidean space. The Lagrangian in the nonlinear gauge of the Curci-Ferrari type cf is

\displaystyle\mathcal{L} =inv+NL,inv=14Fμν2,\displaystyle=\mathcal{L}_{inv}+\mathcal{L}_{NL},\quad\mathcal{L}_{inv}=\frac{1}{4}F_{\mu\nu}^{2},
NL\displaystyle\mathcal{L}_{NL} =BμAμ+ic¯μDμcα12B2α22B¯2Bw,\displaystyle=B\cdot\partial_{\mu}A_{\mu}+i\bar{c}\cdot\partial_{\mu}D_{\mu}c-\frac{\alpha_{1}}{2}B^{2}-\frac{\alpha_{2}}{2}\bar{B}^{2}-B\cdot w, (2.1)

where BB is the Nakanishi-Lautrup field, cc (c¯\bar{c}) is the ghost (antighost), B¯=B+igc¯×c\bar{B}=-B+ig\bar{c}\times c, α1\alpha_{1} and α2\alpha_{2} are gauge parameters, and ww is a constant to keep the BRS symmetry. Introducing the auxiliary field φ\varphi, which represents α2B¯\alpha_{2}\bar{B}, NL\mathcal{L}_{NL} is rewritten as hs2

φ=α12B2+B(μAμ+φw)+ic¯(μDμ+gφ×)c+φ22α2.\mathcal{L}_{\varphi}=-\frac{\alpha_{1}}{2}B^{2}+B\cdot(\partial_{\mu}A_{\mu}+\varphi-w)+i\bar{c}\cdot(\partial_{\mu}D_{\mu}+g\varphi\times)c+\frac{\varphi^{2}}{2\alpha_{2}}. (2.2)

In Ref. hs3 , by integrating out c¯\bar{c} and cc with momentum μpΛ\mu\leq p\leq\Lambda, we studied the one-loop effective potential for φ\varphi, and showed that gφAg\varphi^{A} acquires the VEV gφ0δA3g\varphi_{0}\delta^{A3} under an energy scale μ0\mu_{0}. The scale μ0\mu_{0} and the VEV v=gφ0v=g\varphi_{0} are

μ0=Λe4π2/(α2g2),v={μ04μ41e16π2/(α2g2)}1/2.\mu_{0}=\Lambda e^{-4\pi^{2}/(\alpha_{2}g^{2})},\ v=\left\{\frac{\mu_{0}^{4}-\mu^{4}}{1-e^{-16\pi^{2}/(\alpha_{2}g^{2})}}\right\}^{1/2}. (2.3)

At the one-loop order, it is shown that α2=β0/2\alpha_{2}=\beta_{0}/2 is the ultraviolet fixed point, where β0=22/3\beta_{0}=22/3 is the first coefficient of the renormalization group function β\beta for SU(2). So, when Λμ0μ\Lambda\gg\mu_{0}\gg\mu, the relation

μ0=ΛQCD,vΛQCD2\mu_{0}=\Lambda_{\mathrm{QCD}},\quad v\simeq\Lambda_{\mathrm{QCD}}^{2} (2.4)

holds hs3 , where ΛQCD\Lambda_{\mathrm{QCD}} is the QCD scale parameter.

2.2 Condensate Aμ+Aμ\langle A_{\mu}^{+}A_{\mu}^{-}\rangle

When v0v\neq 0, the ghost Lagrangian becomes

ia=12(c¯acab=12vϵabc¯acb)+ic¯3c3.i\sum_{a=1}^{2}\left(\bar{c}^{a}\Box c^{a}-\sum_{b=1}^{2}v\epsilon_{ab}\bar{c}^{a}c^{b}\right)+i\bar{c}^{3}\Box c^{3}. (2.5)

Because of the vv term in Eq.(2.5), the two-point function

Aμ+(y)Aν(x)=Gμν(y,x)\langle A_{\mu}^{+}(y)A_{\nu}^{-}(x)\rangle=G_{\mu\nu}(y,x)

shows the tachyonic behavior. In fact, the ghost loop depicted in Fig. 1(b) yields the tachyonic mass (g2v)/(64π)(-g^{2}v)/(64\pi) for Aμa(a=1,2)A_{\mu}^{a}\ (a=1,2) in the low momentum limit p0p\to 0. In the same way, Aμ3A_{\mu}^{3} has the tachyonic mass g2v/(32π)-g^{2}v/(32\pi) in this limit.

Refer to caption
Figure 1: The diagrams that contribute to the inverse propagator for Aμ±A_{\mu}^{\pm}. Fig. 1(b) yields the tachyonic mass in the limit p0p\to 0, and Fig. 1(c) brings about the VEV Gμμ(0)(x,x)G_{\mu\mu}^{(0)}(x,x).

To remove the tachyonic mass terms

12(g2v64π){a=12(Aμa)2+2(Aμ3)2},\frac{1}{2}\left(\frac{-g^{2}v}{64\pi}\right)\left\{\sum_{a=1}^{2}(A_{\mu}^{a})^{2}+2(A_{\mu}^{3})^{2}\right\}, (2.6)

we introduce the source term

KAμ+Aμ,K=K(0)+K(1)+,K(n)=O(n).KA_{\mu}^{+}A_{\mu}^{-},\quad K=K^{(0)}+K^{(1)}+\cdots,\quad K^{(n)}=O(\hslash^{n}).

Although the source KK may depend on the momentum scale, for simplicity, we treat it as constant, and write K(0)=M2K^{(0)}=M^{2}. To consider the inverse propagator for Aμ±A_{\mu}^{\pm} depicted in Fig. 1 in the limit p0p\to 0, we write the free part of the propagator Gμν(y,x)G_{\mu\nu}(y,x) as Gμν(0)(y,x)G_{\mu\nu}^{(0)}(y,x). Then the diagram in Fig. 1(c) gives the VEV g2Gμμ(0)(x,x)g^{2}G_{\mu\mu}^{(0)}(x,x). If K(1)K^{(1)} subtracts divergent terms of O()O(\hslash) in this limit, the condition

Gμμ(0)(x,x)=v64π,G_{\mu\mu}^{(0)}(x,x)=\frac{v}{64\pi}, (2.7)

removes the tachyonic mass for Aμ±A_{\mu}^{\pm}. Because of the interaction g2(Aμ+Aμ)(Aν3)2g^{2}(A_{\mu}^{+}A_{\mu}^{-})(A_{\nu}^{3})^{2} in Fμν2/4F_{\mu\nu}^{2}/4, this VEV also removes the tachyonic mass for Aμ3A_{\mu}^{3}.

2.3 Inclusion of a classical solution

To introduce a classical solution, the gauge field AμAA_{\mu}^{A} is divided into the classical part bμA=bμδA3b_{\mu}^{A}=b_{\mu}\delta^{A3} and the quantum part aμAa_{\mu}^{A}, i.e., Aμ±=aμ±,Aμ3=bμ+aμ3A_{\mu}^{\pm}=a_{\mu}^{\pm},\ A_{\mu}^{3}=b_{\mu}+a_{\mu}^{3}. As the tachyonic masses come from the ghost loops with the VEV vv, it is expected that bμb_{\mu} acquires no tachyonic mass. To see it, we divide the gauge transformation δAμ=Dμ(A)ε\delta A_{\mu}=D_{\mu}(A)\varepsilon as

δaμ=Dμ(a)ε,δbμ=gbμ×ε,\delta a_{\mu}=D_{\mu}(a)\varepsilon,\quad\delta b_{\mu}=gb_{\mu}\times\varepsilon,

where Dμ(A)=(μ+gAμ×)D_{\mu}(A)=(\partial_{\mu}+gA_{\mu}\times). Then, using the gauge-fixing function G=μaμ+φwG=\partial_{\mu}a_{\mu}+\varphi-w, the ghost Lagrangian

B[μaμ+φw]+ic¯[μDμ(a)+gφ×]cB\cdot[\partial_{\mu}a_{\mu}+\varphi-w]+i\bar{c}\cdot[\partial_{\mu}D_{\mu}(a)+g\varphi\times]c

is obtained. If AμA_{\mu} is replaced by aμa_{\mu}, Eq.(2.2) becomes this Lagrangian. So, the tachyonic mass terms for aμAa_{\mu}^{A} are

12(g2v64π){a=12(aμa)2+2(aμ3)2},\frac{1}{2}\left(\frac{-g^{2}v}{64\pi}\right)\left\{\sum_{a=1}^{2}(a_{\mu}^{a})^{2}+2(a_{\mu}^{3})^{2}\right\},

and bμb_{\mu} acquires no tachyonic mass. 111In Ref. hs1 , using the background covariant gauge-fixing, we showed, although the ghost Lagrangian contains bμb_{\mu}, it does not acquire the tachyonic mass.

Next, we consider the effect of the VEV Gμμ(0)(x,x)G_{\mu\mu}^{(0)}(x,x). As in the previous subsection, these tachyonic mass terms for aμAa_{\mu}^{A} are removed by the VEV in Eq.(2.7). In addition, the interaction g2(Aμ+Aμ)(aν3+bν)2g^{2}(A_{\mu}^{+}A_{\mu}^{-})(a_{\nu}^{3}+b_{\nu})^{2} in Fμν2/4F_{\mu\nu}^{2}/4 generates the mass term

m22(bν)2,m2=2g2Gμμ(0)(x,x).\frac{m^{2}}{2}(b_{\nu})^{2},\quad m^{2}=2g^{2}G_{\mu\mu}^{(0)}(x,x). (2.8)

Thus, after integrating out cc and c¯\bar{c}, an effective low energy Lagrangian becomes

=\displaystyle\mathcal{L}= 14(b)μν(b)μν+m22bμbμ+14(a3)μν(a3)μν\displaystyle\frac{1}{4}(\partial\wedge b)_{\mu\nu}(\partial\wedge b)_{\mu\nu}+\frac{m^{2}}{2}b_{\mu}b_{\mu}+\frac{1}{4}(\partial\wedge a^{3})_{\mu\nu}(\partial\wedge a^{3})_{\mu\nu}
+12(A+)μν(A)μν+M2Aμ+Aμ+,\displaystyle+\frac{1}{2}(\partial\wedge A^{+})_{\mu\nu}(\partial\wedge A^{-})_{\mu\nu}+M^{2}A^{+}_{\mu}A^{-}_{\mu}+\cdots, (2.9)

where (A)μν=μAννAμ(\partial\wedge A)_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}. Namely, although the quantum part aμ3a_{\mu}^{3} is massless, the classical part bμb_{\mu} has the mass mm. The off-diagonal components Aμ±A_{\mu}^{\pm} have the mass MM determined by the equation

v64π=Gμμ(0)(x,x).\frac{v}{64\pi}=G_{\mu\mu}^{(0)}(x,x). (2.10)

2.4 QQ¯Q\bar{Q} potential

Now we consider the confining potential. As the classical field bμb_{\mu}, we choose the dual electric potential μ\mathcal{B}_{\mu}, that describes the electric monopole solution hs . The color electric current jμj_{\mu} is incorporated by the replacement

()μν()μν+ϵμναβ(n)1nαjβ,(\partial\wedge\mathcal{B})_{\mu\nu}\to(\partial\wedge\mathcal{B})_{\mu\nu}+\epsilon_{\mu\nu\alpha\beta}(n\cdot\partial)^{-1}n_{\alpha}j_{\beta},

where the space-like vector nμn_{\mu} zwa is chosen as nμ=(0,𝒏)n_{\mu}=(0,\mbox{\boldmath$n$}) with 𝒏𝒏=1\mbox{\boldmath$n$}\cdot\mbox{\boldmath$n$}=1, and n=nμμn\cdot\partial=n_{\mu}\partial_{\mu}. We note this is the Zwanziger’s dual field strength Fd=(B)+(n)1(nje)dF^{d}=(\partial\wedge B)+(n\cdot\partial)^{-1}(n\wedge j_{e})^{d} in Ref. zwa . Thus the classical part of \mathcal{L} in Eq.(2.9) becomes

14[()μν+ϵμναβ(n)1nαjβ]2+m22μμ.\frac{1}{4}\left[(\partial\wedge\mathcal{B})_{\mu\nu}+\epsilon_{\mu\nu\alpha\beta}(n\cdot\partial)^{-1}n_{\alpha}j_{\beta}\right]^{2}+\frac{m^{2}}{2}\mathcal{B}_{\mu}\mathcal{B}_{\mu}. (2.11)

The equation of motion for μ\mathcal{B}_{\mu} is

(Dm1)μνν=ϵμραβ(n)1nραjβ,(Dm1)μν=(+m2)δμν+μν,(D_{m}^{-1})_{\mu\nu}\mathcal{B}_{\nu}=\epsilon_{\mu\rho\alpha\beta}(n\cdot\partial)^{-1}n_{\rho}\partial_{\alpha}j_{\beta},\quad(D_{m}^{-1})_{\mu\nu}=(-\square+m^{2})\delta_{\mu\nu}+\partial_{\mu}\partial_{\nu},

and μ\mathcal{B}_{\mu} is solved as

μ=(Dm)μνϵνραβ(n)1nραjβ,(Dm)μν=δμνμν/+m2+μνm2.\mathcal{B}_{\mu}=(D_{m})_{\mu\nu}\epsilon_{\nu\rho\alpha\beta}(n\cdot\partial)^{-1}n_{\rho}\partial_{\alpha}j_{\beta},\quad(D_{m})_{\mu\nu}=\frac{\delta_{\mu\nu}-\partial_{\mu}\partial_{\nu}/\square}{-\square+m^{2}}+\frac{\partial_{\mu}\partial_{\nu}}{m^{2}\square}. (2.12)

If we use Eq.(2.12), Eq.(2.11) becomes

jj=12jμ1+m2jμ12jμm2+m2nn(n)2(δμνnμnνnn)jν.\mathcal{L}_{jj}=\frac{1}{2}j_{\mu}\frac{1}{-\square+m^{2}}j_{\mu}-\frac{1}{2}j_{\mu}\frac{m^{2}}{-\square+m^{2}}\frac{n\cdot n}{(n\cdot\partial)^{2}}\left(\delta_{\mu\nu}-\frac{n_{\mu}n_{\nu}}{n\cdot n}\right)j_{\nu}. (2.13)

To derive the static potential between the color electric charges QQ and Q-Q, the static current

jμ(x)=Qδμ0{δ(𝒙𝒂)δ(𝒙𝒃)}j_{\mu}(x)=Q\delta_{\mu 0}\{\delta(\mbox{\boldmath$x$}-\mbox{\boldmath$a$})-\delta(\mbox{\boldmath$x$}-\mbox{\boldmath$b$})\} (2.14)

is substituted into jj\mathcal{L}_{jj}. Then it leads to the potential

V(r)=\displaystyle V(r)= VY(r)+VL(r),VY=Q2d3q(2π)31cos𝒒𝒓q2+m2,\displaystyle V_{Y}(r)+V_{L}(r),\quad V_{Y}=Q^{2}\int\frac{d^{3}q}{(2\pi)^{3}}\frac{1-\cos\mbox{\boldmath$q$}\cdot\mbox{\boldmath$r$}}{q^{2}+m^{2}},
VL=Q2d3q(2π)3(1cos𝒒𝒓)m2(q2+m2)qn2,\displaystyle V_{L}=Q^{2}\int\frac{d^{3}q}{(2\pi)^{3}}(1-\cos\mbox{\boldmath$q$}\cdot\mbox{\boldmath$r$})\frac{m^{2}}{(q^{2}+m^{2})q_{n}^{2}}, (2.15)

where 𝒓=𝒂𝒃\mbox{\boldmath$r$}=\mbox{\boldmath$a$}-\mbox{\boldmath$b$}, q=|𝒒|q=|\mbox{\boldmath$q$}| and qn=𝒒𝒏q_{n}=\mbox{\boldmath$q$}\cdot\mbox{\boldmath$n$}. Although the term VYV_{Y} (VLV_{L}) becomes the Yukawa (linear) potential in Eq.(1.2) usually, we restudy Eq.(2.15) in Sect. 4.

3 Cut-off Λc\Lambda_{c} and the mass MM

In this section, we study Eq.(2.10). The free propagator Gμν(0)(p)G_{\mu\nu}^{(0)}(p) is calculated in Appendix A as

Gμν(0)(p)=1p2+M2PμνT+1M2PμνL,PμνT=δμνpμpνp2,PμνL=pμpνp2.G_{\mu\nu}^{(0)}(p)=\frac{1}{p^{2}+M^{2}}P^{T}_{\mu\nu}+\frac{1}{M^{2}}P^{L}_{\mu\nu},\quad P^{T}_{\mu\nu}=\delta_{\mu\nu}-\frac{p_{\mu}p_{\nu}}{p^{2}},\quad P^{L}_{\mu\nu}=\frac{p_{\mu}p_{\nu}}{p^{2}}. (3.1)

Assuming that MM exists below a cut-off Λc\Lambda_{c}, we obtain

Gμμ(0)(x,x)=0Λcd4p(2π)4Gμμ(0)(p)=1(4π)2{Λc42M23M2ln(1+Λc2M2)},G_{\mu\mu}^{(0)}(x,x)=\int_{0}^{\Lambda_{c}}\frac{d^{4}p}{(2\pi)^{4}}G_{\mu\mu}^{(0)}(p)=\frac{1}{(4\pi)^{2}}\left\{\frac{\Lambda_{c}^{4}}{2M^{2}}-3M^{2}\ln\left(1+\frac{\Lambda_{c}^{2}}{M^{2}}\right)\right\}, (3.2)

where the MM-independent term Λc2\Lambda_{c}^{2} is subtracted.

The VEV vv depends on the momentum scale μ\mu. From Eqs.(2.3) and (2.4), we find vv disappears above the scale μ0=ΛQCD\mu_{0}=\Lambda_{\mathrm{QCD}}. When 0μΛQCD0\leq\mu\leq\Lambda_{\mathrm{QCD}}, vv behaves as

v0(μΛQCD),vΛQCD2(μ0).v\to 0\quad(\mu\to\Lambda_{\mathrm{QCD}}),\quad v\to\Lambda_{\mathrm{QCD}}^{2}\quad(\mu\to 0).

Namely the maximal value of vv is ΛQCD2\Lambda_{\mathrm{QCD}}^{2}, and the left hand side of Eq.(2.10) satisfies

0v64πΛQCD264π.0\leq\frac{v}{64\pi}\leq\frac{\Lambda_{\mathrm{QCD}}^{2}}{64\pi}.

On the other hand, the VEV Gμμ(0)(x,x)G_{\mu\mu}^{(0)}(x,x) only depends on the constants MM and Λc\Lambda_{c}. Since the tachyonic mass should be removed in the overall momentum region completely, using the maximal value of vv, we interpret Eq.(2.10) as 222If we consider the scale dependent MM, Eq.(3.2) should be replaced by the integral 0Λcd4p(2π)4{PμνTp2+M(p)2+PμνLM(p)2}.\int_{0}^{\Lambda_{c}}\frac{d^{4}p}{(2\pi)^{4}}\left\{\frac{P_{\mu\nu}^{T}}{p^{2}+M(p)^{2}}+\frac{P_{\mu\nu}^{L}}{M(p)^{2}}\right\}. As M(p)M(p) is unknown, we cannot calculate it. But it is also μ\mu-independent.

ΛQCD264πGμμ(0)(x,x).\frac{\Lambda_{\mathrm{QCD}}^{2}}{64\pi}\simeq G_{\mu\mu}^{(0)}(x,x). (3.3)

From Eqs.(3.2) and (3.3), we obtain

π2ΛQCD2M2Λc4M46ln(1+Λc2M2).\frac{\pi}{2}\frac{\Lambda_{\mathrm{QCD}}^{2}}{M^{2}}\simeq\frac{\Lambda_{c}^{4}}{M^{4}}-6\ln\left(1+\frac{\Lambda_{c}^{2}}{M^{2}}\right). (3.4)

We make a comment. Eq.(3.4) determines the relation of MM and Λc\Lambda_{c} so as to give the maximal value of vv. Using Eq.(3.3), we find m2=2g2Gμμ(0)(x,x)g2ΛQCD2/(32π)m^{2}=2g^{2}G_{\mu\mu}^{(0)}(x,x)\simeq g^{2}\Lambda_{\mathrm{QCD}}^{2}/(32\pi). Although the VEV vv depends on the scale μ\mu, and disappears above ΛQCD\Lambda_{\mathrm{QCD}}, Gμμ(0)(x,x)G_{\mu\mu}^{(0)}(x,x) does not. So the mass mm can survive above the scale ΛQCD\Lambda_{\mathrm{QCD}}.

4 Confining potential

Usually, the potential VV in Eq.(2.15) is calculated as follows. Let us divide the momentum 𝒒q in VLV_{L} into qn=𝒒𝒏q_{n}=\mbox{\boldmath$q$}\cdot\mbox{\boldmath$n$} and 𝒒T\mbox{\boldmath$q$}_{T} that satisfies 𝒒T𝒏=0\mbox{\boldmath$q$}_{T}\cdot\mbox{\boldmath$n$}=0. Then the integral of qnq_{n} has the infrared divergence, and the integral of qT=|𝒒T|q_{T}=|\mbox{\boldmath$q$}_{T}| has the ultraviolet divergence. The former divergence is removed by the choice 𝒓𝒏\mbox{\boldmath$r$}\parallel\mbox{\boldmath$n$} sst ; hs , and the latter divergence is avoided by the cut-off mχm_{\chi} as suz ; mts ; sst ; sst2

VL=Q2m2(2π)2𝑑qn0mχ𝑑qTqT1cosqnr(qn2+qT2+m2)qn2.V_{L}=\frac{Q^{2}m^{2}}{(2\pi)^{2}}\int_{-\infty}^{\infty}dq_{n}\int_{0}^{m_{\chi}}dq_{T}q_{T}\frac{1-\cos q_{n}r}{(q_{n}^{2}+q_{T}^{2}+m^{2})q_{n}^{2}}. (4.1)

Eq.(4.1) becomes the linear term in Eq.(1.2). The term VYV_{Y} has the integral of q=|𝒒|q=|\mbox{\boldmath$q$}| over the region of 0q<0\leq q<\infty, and the Yukawa potential in Eq.(1.2) is obtained.

To introduce a cut-off in a different way, we write the potential as

V(r)=𝑑qW(𝒒,m,r).V(r)=\int dqW(\mbox{\boldmath$q$},m,r). (4.2)

As we stated in Sect. 3, the mass mm can exist above ΛQCD\Lambda_{\mathrm{QCD}}. From Fig. 1(c), after vv disappears, it is expected that m2m^{2} contributes to the correction for M2M^{2}. Since Aμ±A_{\mu}^{\pm} are considered to be massless above Λc\Lambda_{c}, we assume that the cut-off for mm is Λc\Lambda_{c}, and define Eq.(4.2) as

V(r)=0Λc𝑑qW(𝒒,m,r)+Λc𝑑qW(𝒒,0,r).V(r)=\int_{0}^{\Lambda_{c}}dqW(\mbox{\boldmath$q$},m,r)+\int_{\Lambda_{c}}^{\infty}dqW(\mbox{\boldmath$q$},0,r). (4.3)

Eq.(4.3) is rewritten as

V(r)=0𝑑qW(𝒒,0,r)+0Λc𝑑q{W(𝒒,m,r)W(𝒒,0,r)}.V(r)=\int_{0}^{\infty}dqW(\mbox{\boldmath$q$},0,r)+\int_{0}^{\Lambda_{c}}dq\left\{W(\mbox{\boldmath$q$},m,r)-W(\mbox{\boldmath$q$},0,r)\right\}. (4.4)

The first term becomes

0𝑑qW(𝒒,0,r)=V1(r)=Q2Dd3q(2π)31cos𝒒𝒓q2,D={q| 0q<},\int_{0}^{\infty}dqW(\mbox{\boldmath$q$},0,r)=V_{1}(r)=Q^{2}\int_{D_{\infty}}\frac{d^{3}q}{(2\pi)^{3}}\frac{1-\cos\mbox{\boldmath$q$}\cdot\mbox{\boldmath$r$}}{q^{2}},\quad D_{\infty}=\{q\ |\ 0\leq q<\infty\}, (4.5)

and the second term leads to

0Λc𝑑q{W(𝒒,m,r)W(𝒒,0,r)}=V2(r)+V3(r),\displaystyle\int_{0}^{\Lambda_{c}}dq\left\{W(\mbox{\boldmath$q$},m,r)-W(\mbox{\boldmath$q$},0,r)\right\}=V_{2}(r)+V_{3}(r),
V2(r)=Q2DΛcd3q(2π)3(1cos𝒒𝒓)m2(q2+m2)qn2,\displaystyle V_{2}(r)=Q^{2}\int_{D_{\Lambda_{c}}}\frac{d^{3}q}{(2\pi)^{3}}(1-\cos\mbox{\boldmath$q$}\cdot\mbox{\boldmath$r$})\frac{m^{2}}{(q^{2}+m^{2})q_{n}^{2}},
V3(r)=Q2DΛcd3q(2π)3(1cos𝒒𝒓)m2(q2+m2)q2,DΛc={q| 0q<Λc},\displaystyle V_{3}(r)=-Q^{2}\int_{D_{\Lambda_{c}}}\frac{d^{3}q}{(2\pi)^{3}}(1-\cos\mbox{\boldmath$q$}\cdot\mbox{\boldmath$r$})\frac{m^{2}}{(q^{2}+m^{2})q^{2}},\quad D_{\Lambda_{c}}=\{q\ |\ 0\leq q<\Lambda_{c}\}, (4.6)

where V2(r)V_{2}(r) (V3(r)V_{3}(r)) comes from VLV_{L} (VY(m2)VY(m2=0)V_{Y}(m^{2})-V_{Y}(m^{2}=0)). Eq.(4.5) gives the usual Coulomb potential

V1(r)=Kcr,Kc=Q24π.V_{1}(r)=-\frac{K_{c}}{r},\quad K_{c}=\frac{Q^{2}}{4\pi}. (4.7)

Now we consider V2(r)V_{2}(r). To satisfy 0qΛc0\leq q\leq\Lambda_{c}, the domain of integration is not (<qn<,0qTmχ)(-\infty<q_{n}<\infty,0\leq q_{T}\leq m_{\chi}) in Eq.(4.1), but (εqnε,0qTΛc2ε2)(-\varepsilon\leq q_{n}\leq\varepsilon,0\leq q_{T}\leq\sqrt{\Lambda_{c}^{2}-\varepsilon^{2}}) with 0<ε10<\varepsilon\ll 1, i.e.,

V2(r)=Q2m2(2π)20Λc2ε2𝑑qTqTεε𝑑qn1eiqnrqn2(qn2+qT2+m2).V_{2}(r)=\frac{Q^{2}m^{2}}{(2\pi)^{2}}\int_{0}^{\sqrt{\Lambda_{c}^{2}-\varepsilon^{2}}}dq_{T}q_{T}\int_{-\varepsilon}^{\varepsilon}dq_{n}\frac{1-e^{iq_{n}r}}{q_{n}^{2}(q_{n}^{2}+q_{T}^{2}+m^{2})}.

Since the integrand is singular at qn=0q_{n}=0, we choose the anticlockwise path Γε\Gamma_{\varepsilon} in Fig. 2, and take the limit ε0\varepsilon\to 0. Then we obtain

V2(r)=Q2m2(2π)20Λc𝑑qTqTπrqT2+m2=σcr,σc=Q2m28πln(Λc2+m2m2).V_{2}(r)=\frac{Q^{2}m^{2}}{(2\pi)^{2}}\int_{0}^{\Lambda_{c}}dq_{T}q_{T}\frac{\pi r}{q_{T}^{2}+m^{2}}=\sigma_{c}r,\quad\sigma_{c}=\frac{Q^{2}m^{2}}{8\pi}\ln\left(\frac{\Lambda_{c}^{2}+m^{2}}{m^{2}}\right). (4.8)

If the cut-off Λc\Lambda_{c} is replaced by mχm_{\chi}, V2(r)V_{2}(r) becomes the linear term in Eq.(1.2).

Refer to caption
Figure 2: The integration path Γε\Gamma_{\varepsilon} for qnq_{n}.

Finally, neglecting additive constants, we find V3(r)V_{3}(r) becomes

V3(r)=Q2m22π20Λc𝑑qsinqrqr1q2+m2.V_{3}(r)=\frac{Q^{2}m^{2}}{2\pi^{2}}\int_{0}^{\Lambda_{c}}dq\frac{\sin qr}{qr}\frac{1}{q^{2}+m^{2}}. (4.9)

Thus the confining potential we propose is

Vc(r)=k=13Vk(r)=Kcr+σcr+Q2m22π20Λc𝑑qsinqrqr1q2+m2.V_{c}(r)=\sum_{k=1}^{3}V_{k}(r)=-\frac{K_{c}}{r}+\sigma_{c}r+\frac{Q^{2}m^{2}}{2\pi^{2}}\int_{0}^{\Lambda_{c}}dq\frac{\sin qr}{qr}\frac{1}{q^{2}+m^{2}}. (4.10)

In addition to the Coulomb plus linear part, there is the term V3(r)V_{3}(r).

5 Determination of parameters

Although we presented the potential Vc(r)=k=13Vk(r)V_{c}(r)=\sum_{k=1}^{3}V_{k}(r), the values of the parameters Q2Q^{2} and mm are unknown. To determine them, let us expand a potential V(r)V(r) as

V(r)=V(rc)+V(rc)(rrc)+V′′(rc)2(rrc)2+V(3)(rc)3!(rrc)3+.V(r)=V(r_{c})+V^{\prime}(r_{c})(r-r_{c})+\frac{V^{\prime\prime}(r_{c})}{2}(r-r_{c})^{2}+\frac{V^{(3)}(r_{c})}{3!}(r-r_{c})^{3}+\cdots.

We assume there is a true confining potential VT(r)V_{T}(r). Then we require the Cornell potential VCL(r)V_{CL}(r) fits well to VT(r)V_{T}(r) at a point r=rcr=r_{c}. This is achieved by choosing (K,σ)(K,\sigma) and the constant V(rc)V(r_{c}) appropriately. 333To determine the three parameters, it is natural to choose appropriate three points rk(k=1,2,3)r_{k}\ (k=1,2,3). However, in the next step, we must determine VcV_{c} so as to the difference between VcV_{c} and VCLV_{CL} becomes minimum. Since V3V_{3} in VcV_{c} contains rr in the integrand, it is difficult to determine rkr_{k}. We impose the conditions

VT(n)(rc)=VCL(n)(rc),(n=0,1,2),V_{T}^{(n)}(r_{c})=V_{CL}^{(n)}(r_{c}),\quad(n=0,1,2), (5.1)

and determine (K,σ)(K,\sigma) and V(rc)V(r_{c}).

Next, we require that the potential Vc(r)V_{c}(r) fits well to this VCL(r)V_{CL}(r) at r=rcr=r_{c}, and use the conditions

VCL(n)(rc)=Vc(n)(rc),(n=0,1,2,3).V_{CL}^{(n)}(r_{c})=V_{c}^{(n)}(r_{c}),\quad(n=0,1,2,3). (5.2)

We note, to determine the parameters Q2Q^{2} and mm, the two conditions with n=1,2n=1,2 are necessary. However, to determine rcr_{c}, the condition with n=3n=3 is required.

From Eqs.(1.1) and (4.10), we have

r32VCL′′(r)=K,12r(r2VCL(r))=σ,\displaystyle-\frac{r^{3}}{2}V_{CL}^{\prime\prime}(r)=K,\quad\frac{1}{2r}\left(r^{2}V_{CL}^{\prime}(r)\right)^{\prime}=\sigma,
r32Vc′′(r)=Kcr32V3′′(r),12r(r2Vc(r))=σc+12r(r2V3(r)).\displaystyle-\frac{r^{3}}{2}V_{c}^{\prime\prime}(r)=K_{c}-\frac{r^{3}}{2}V_{3}^{\prime\prime}(r),\quad\frac{1}{2r}\left(r^{2}V_{c}^{\prime}(r)\right)^{\prime}=\sigma_{c}+\frac{1}{2r}\left(r^{2}V_{3}^{\prime}(r)\right)^{\prime}.

Since Eq.(5.2) with n=2n=2 gives rc3VCL′′(rc)=rc3Vc′′(rc)r_{c}^{3}V_{CL}^{\prime\prime}(r_{c})=r_{c}^{3}V_{c}^{\prime\prime}(r_{c}), this condition becomes

K=Kef(rc),Kef(rc)=Kcrc32V3′′(rc).K=K_{ef}(r_{c}),\quad K_{ef}(r_{c})=K_{c}-\frac{r_{c}^{3}}{2}V_{3}^{\prime\prime}(r_{c}). (5.3)

In the same way, the condition (rc2VCL(rc))=(rc2VC(rc))\displaystyle\left(r_{c}^{2}V_{CL}^{\prime}(r_{c})\right)^{\prime}=\left(r_{c}^{2}V_{C}^{\prime}(r_{c})\right)^{\prime} leads to

σ=σef(rc),σef(rc)=σc+12rc(rc2V3(rc)),\sigma=\sigma_{ef}(r_{c}),\quad\sigma_{ef}(r_{c})=\sigma_{c}+\frac{1}{2r_{c}}\left(r_{c}^{2}V_{3}^{\prime}(r_{c})\right)^{\prime}, (5.4)

and (rc3VCL′′(rc))=(rc3Vc′′(rc))\displaystyle\left(r_{c}^{3}V_{CL}^{\prime\prime}(r_{c})\right)^{\prime}=\left(r_{c}^{3}V_{c}^{\prime\prime}(r_{c})\right)^{\prime} becomes

(rc3V3′′(rc))=0.\left(r_{c}^{3}V_{3}^{\prime\prime}(r_{c})\right)^{\prime}=0. (5.5)

Of course, the true potential VT(r)V_{T}(r) is unknown. So, instead of the first step proposed in Eq.(5.1), we choose appropriate values of KK and σ\sigma. Thus we determine Vc(r)V_{c}(r) as follows. Choosing the values of the scale parameter ΛQCD\Lambda_{\mathrm{QCD}} and the off-diagonal gluon mass MM, Eq.(3.4) determines the cut-off Λc\Lambda_{c}. Then, substituting the values of Λc\Lambda_{c}, KK and σ\sigma into Eqs.(5.3)-(5.5), the quantities QQ, mm and rcr_{c} are determined numerically. As an example, we choose the values

ΛQCD=0.2GeV,M=1.2GeV,K=0.3,σ=0.18GeV2.\Lambda_{\mathrm{QCD}}=0.2\ \mathrm{GeV},\quad M=1.2\ \mathrm{GeV},\quad K=0.3,\quad\sigma=0.18\ \mathrm{GeV}^{2}. (5.6)

We note that the off-diagonal gluon mass M1.2M\simeq 1.2 GeV in the region of r0.2r\gtrsim 0.2 fm was obtained by using SU(2) lattice QCD in the maximal Abelian gauge as . The values K0.3K\lesssim 0.3 and σ0.2\sigma\lesssim 0.2 GeV2 come from lattice simulations bali .

Now using the values of ΛQCD\Lambda_{\mathrm{QCD}} and MM in Eq.(5.6), Eq.(3.4) gives Λc2.03\Lambda_{c}\simeq 2.03 GeV. Next, we substitute this Λc\Lambda_{c} and (K,σ)(K,\ \sigma) in Eq.(5.6) into Eqs.(5.3)-(5.5), and solve these equations. The details are explained in Appendix B. The results are

rc1.145GeV1=0.226fm,a=m2/Λc20.263,r_{c}\simeq 1.145\ \mathrm{GeV}^{-1}=0.226\ \mathrm{fm},\quad a=m^{2}/\Lambda_{c}^{2}\simeq 0.263,

and these values lead to

m=1.04GeV,Kc=Q24π=0.285,σc=Q2m28πln(Λc2+m2m2)=0.242GeV2.m=1.04\ \mathrm{GeV},\quad K_{c}=\frac{Q^{2}}{4\pi}=0.285,\quad\sigma_{c}=\frac{Q^{2}m^{2}}{8\pi}\ln\left(\frac{\Lambda_{c}^{2}+m^{2}}{m^{2}}\right)=0.242\ \mathrm{GeV}^{2}. (5.7)

Thus we obtain

Vc(r)=k=13Vk(r),V1(r)=0.285r,V2(r)=0.242r,V3(r)=0.747S(r,0.263),V_{c}(r)=\sum_{k=1}^{3}V_{k}(r),\quad V_{1}(r)=-\frac{0.285}{r},\quad V_{2}(r)=0.242\cdot r,\quad V_{3}(r)=0.747\cdot S(r,0.263), (5.8)

where

S(r,a)=0Λcr𝑑xsinxxarx2+a(Λcr)2.S(r,a)=\int_{0}^{\Lambda_{c}r}dx\frac{\sin x}{x}\frac{ar}{x^{2}+a(\Lambda_{c}r)^{2}}. (5.9)

In Fig.3, the potentials Vk(r),(k=1,2,3)V_{k}(r),(k=1,2,3) and Vc(r)V_{c}(r) are plotted. Since V1(r)+V3(r)V_{1}(r)+V_{3}(r) is a substitute for the Yukawa potential, V1(r)+V3(r)0V_{1}(r)+V_{3}(r)\approx 0 for r0.35fmr\gtrsim 0.35\ \mathrm{fm} is reasonable. Using the values of (K,σ)(K,\ \sigma) in Eq.(5.6), VCL(r)V_{CL}(r) becomes

VCL(r)=0.3r+0.18r.V_{CL}(r)=-\frac{0.3}{r}+0.18\cdot r. (5.10)

Eqs.(5.8) and (5.10) are plotted in Fig. 4. Since Vc(r)V_{c}(r) is fitted to VCL(r)V_{CL}(r) at rc0.226fmr_{c}\simeq 0.226\ \mathrm{fm}, they fit very well for 0.1fmr0.4fm0.1\ \mathrm{fm}\lesssim r\lesssim 0.4\ \mathrm{fm}. However, when rr becomes large, as V1(r)+V3(r)0V_{1}(r)+V_{3}(r)\approx 0 and σc>σ\sigma_{c}>\sigma, Vc(r)>VCL(r)V_{c}(r)>V_{CL}(r) holds for r0.4fmr\gtrsim 0.4\ \mathrm{fm}. In the same way, Kc<KK_{c}<K leads to Vc(r)>VCL(r)V_{c}(r)>V_{CL}(r) for r<0.09fmr<0.09\ \mathrm{fm}.

Refer to caption
Figure 3: The potentials V1,V2,V3V_{1},V_{2},V_{3} and Vc=k=13VkV_{c}=\sum_{k=1}^{3}V_{k} in Eq.(5.8). The unit of rr is fm, and the unit of the potentials is GeV.
Refer to caption
Figure 4: The potentials VcV_{c} in Eq.(5.8) and VCLV_{CL} with (K=0.3,σ=0.18GeV2)(K=0.3,\sigma=0.18\ \mathrm{GeV}^{2}). Additive Constants are chosen to become Vc(rc)=VCL(rc)=0V_{c}(r_{c})=V_{CL}(r_{c})=0, and the unit of rr is fm.

6 The scales R0R_{0} and R~0\tilde{R}_{0}

In Sect. 5, the scale rc0.226fmr_{c}\simeq 0.226\ \mathrm{fm} appears. On the other hand, considering the force V(r)-V^{\prime}(r), the intermediate scale R0R_{0}, which satisfies

r2V(r)|r=R0=1.65,R00.5fm,r^{2}V^{\prime}(r)|_{r=R_{0}}=1.65,\quad R_{0}\simeq 0.5\ \mathrm{fm}, (6.1)

was proposed so . In successful potential models, this relation holds fairly well. For example, the VCL(r)V_{CL}(r) with (K=0.52,σ=0.183GeV2)(K=0.52,\ \sigma=0.183\ \mathrm{GeV}^{2}) cor2 gives R0=0.49fmR_{0}=0.49\ \mathrm{fm}. If we substitute Vc(r)V_{c}(r) in Eq.(5.8) into Eq.(6.1), we obtain

Kef(R0)+R02σef(R0)=0.285+0.242R02+0.747R02H(R0,0.263)=1.65,K_{ef}(R_{0})+R_{0}^{2}\sigma_{ef}(R_{0})=0.285+0.242R_{0}^{2}+0.747R_{0}^{2}\cdot H(R_{0},0.263)=1.65, (6.2)

where

H(r,a)=0Λcr𝑑x(cosxsinxx)ax2+a(Λcr)2H(r,a)=\int_{0}^{\Lambda_{c}r}dx\left(\cos x-\frac{\sin x}{x}\right)\frac{a}{x^{2}+a(\Lambda_{c}r)^{2}} (6.3)

comes from V3(r)V_{3}^{\prime}(r). Eq.(6.2) gives the solution R0=0.51fmR_{0}=0.51\ \mathrm{fm}. We note, VCL(r)V_{CL}(r) with (K=0.3,σ=0.18GeV2)(K=0.3,\ \sigma=0.18\ \mathrm{GeV}^{2}) in Sect. 5 gives the larger value R0=0.54fmR_{0}=0.54\ \mathrm{fm}.

Refer to caption
Figure 5: The effective Coulomb coupling Kef(r)=Kcr32V3′′(r)K_{ef}(r)=K_{c}-\frac{r^{3}}{2}V_{3}^{\prime\prime}(r) and the effective string tension Se(r)=σef(r)=σc+12r(r2V3(r))S_{e}(r)=\sigma_{ef}(r)=\sigma_{c}+\frac{1}{2r}\left(r^{2}V_{3}^{\prime}(r)\right)^{\prime}. The unit of rr is fm.

To see the meanings of these scales, the effective Coulomb coupling Kef(r)=Kcr32V3′′(r)K_{ef}(r)=K_{c}-\frac{r^{3}}{2}V_{3}^{\prime\prime}(r) in Eq.(5.3) and the effective string tension σef(r)=σc+12r(r2V3(r))\sigma_{ef}(r)=\sigma_{c}+\frac{1}{2r}\left(r^{2}V_{3}^{\prime}(r)\right)^{\prime} in Eq.(5.4) are plotted in Fig. 5. We find Kef(rc)=0.3K_{ef}(r_{c})=0.3 is the maximal value and σef(rc)=0.18GeV2\sigma_{ef}(r_{c})=0.18\ \mathrm{GeV}^{2} is the minimal value. Namely rcr_{c} is the position where Kef(r)K_{ef}(r) is maximum and σef(r)\sigma_{ef}(r) is minimum. We also notice that Kef(r)0K_{ef}(r)\approx 0 at r0.58fmr\approx 0.58\ \mathrm{fm}, and 0.23GeV2<σef(r)<0.25GeV20.23\ \mathrm{GeV}^{2}<\sigma_{ef}(r)<0.25\ \mathrm{GeV}^{2} for r0.5fmr\gtrsim 0.5\ \mathrm{fm}.

In Fig. 6, r2Vc(r)=Kef(r)+r2σef(r)r^{2}V_{c}^{\prime}(r)=K_{ef}(r)+r^{2}\sigma_{ef}(r) and r2σef(r)r^{2}\sigma_{ef}(r) are plotted. We find that r2Vc(r)r^{2}V_{c}^{\prime}(r) satisfies r2Vc(r)Kef(r)r^{2}V_{c}^{\prime}(r)\approx K_{ef}(r) for r<0.2fmr<0.2\ \mathrm{fm}, and r2Vc(r)r2σef(r)r^{2}V_{c}^{\prime}(r)\approx r^{2}\sigma_{ef}(r) for r0.5fmr\gtrsim 0.5\ \mathrm{fm}. Namely, the force-related quantity r2Vc(r)r^{2}V_{c}^{\prime}(r) is almost saturated with the string part r2σef(r)r^{2}\sigma_{ef}(r) above R00.5fmR_{0}\simeq 0.5\ \mathrm{fm}. Especially, as Kef(r)0K_{ef}(r)\approx 0 at r0.58fmr\approx 0.58\ \mathrm{fm}, we find

r2Vc(r)|r=R~0r2σef(r)|r=R~0=2.13,R~00.58fm.r^{2}V_{c}^{\prime}(r)|_{r=\tilde{R}_{0}}\simeq r^{2}\sigma_{ef}(r)|_{r=\tilde{R}_{0}}=2.13,\quad\tilde{R}_{0}\simeq 0.58\ \mathrm{fm}.
Refer to caption
Figure 6: Fc(r)=r2Vc(r)=Kef(r)+r2σef(r)F_{c}(r)=r^{2}V_{c}^{\prime}(r)=K_{ef}(r)+r^{2}\sigma_{ef}(r) and FS(r)=r2σef(r)FS(r)=r^{2}\sigma_{ef}(r). For comparison, FCL(r)=r2VCL(r)=K+r2σFCL(r)=r^{2}V_{CL}^{\prime}(r)=K+r^{2}\sigma with (K=0.3,σ=0.212GeV2)(K=0.3,\sigma=0.212\ \mathrm{GeV}^{2}) is plotted. The unit of rr is fm.

Before closing this section, based on the above analysis, we present the potential VCL(r)V_{CL}(r) that fits to Vc(r)V_{c}(r) better in the region of r>0.6fmr>0.6\ \mathrm{fm}. When rr is small, the Coulomb part Kef(r)K_{ef}(r) dominates. So keeping the condition Eq.(5.3) intact, we set K=0.3K=0.3. When rr becomes large, the string part dominates. To determine the value of σ\sigma, it is reasonable to set the condition

r2V(r)|r=R~0=2.13,R~00.58fm.r^{2}V^{\prime}(r)|_{r=\tilde{R}_{0}}=2.13,\quad\tilde{R}_{0}\simeq 0.58\ \mathrm{fm}. (6.4)

We find VCLV_{CL} with (K=0.3,σ=2.12GeV2)(K=0.3,\sigma=2.12\ \mathrm{GeV}^{2}) satisfies Eq.(6.4). We note this VCLV_{CL} satisfies Eq.(6.1) as well.

In Fig. 7, the potential Vc(r)V_{c}(r) and

VCL(r)=0.3r+2.12rV_{CL}(r)=-\frac{0.3}{r}+2.12\cdot r (6.5)

are plotted . As we explained in Sect. 5, the behavior Vc(r)>VCL(r)V_{c}(r)>V_{CL}(r) comes about for large rr. However Eq.(6.5) fits fairly well for r<1.2fmr<1.2\ \mathrm{fm}.

Refer to caption
Figure 7: The potentials VcV_{c} in Eq.(5.8) and VCLV_{CL} with (K=0.3,σ=0.212GeV2)(K=0.3,\sigma=0.212\ \mathrm{GeV}^{2}). The unit of rr is fm.

7 Summary and comments

In this paper, we considered the SU(2) gauge theory, and studied the QQ¯Q\bar{Q} potential Eq.(2.15). This potential is derived under the gauge field condensation. In Refs. suz ; mts ; sst ; sst2 , the dual Ginzburg-Landau model, which describes the monopole condensation, leads to the potential.

In our approach hs ; hs1 , the ghost condensation v0v\neq 0 appears, and it induces the VEV Gμμ(0)(x,x)G_{\mu\mu}^{(0)}(x,x), that is the lowest term of Aμ+Aμ\langle A_{\mu}^{+}A_{\mu}^{-}\rangle. If we divide the diagonal gluon as Aμ3=aμ3+bμA_{\mu}^{3}=a_{\mu}^{3}+b_{\mu}, the classical part bμb_{\mu} acquires the mass m=g2Gμμ(0)(x,x)gΛQCD/32πm=g\sqrt{2G_{\mu\mu}^{(0)}(x,x)}\simeq g\Lambda_{\mathrm{QCD}}/\sqrt{32\pi}, whereas the quantum part aμ3a_{\mu}^{3} is massless. The off-diagonal gluons Aμ±A_{\mu}^{\pm} acquire the mass MM through Eq.(3.3). The low energy effective Lagrangian is Eq.(2.9). As the classical solution bμb_{\mu}, we choose the electric monopole solution of the dual gauge field μ\mathcal{B}_{\mu} hs1 . Then the propagator of μ\mathcal{B}_{\mu} leads to the QQ¯Q\bar{Q} potential Eq.(2.15).

In calculating Eqs.(3.3) and (4.3), ultraviolet cut-off is necessary. Above the scale ΛQCD\Lambda_{\mathrm{QCD}}, although vv vanishes, the masses MM and mm can exist. We assumed that these masses disappear above the cut-off Λc\Lambda_{c}. Then the condition in Eq.(3.4) and the confining potential Vc(r)V_{c}(r) in Eq.(4.10) are obtained. This potential has the linear potential V2(r)V_{2}(r) and, instead of the Yukawa potential, the Coulomb potential V1(r)V_{1}(r) and the additional term V3(r)V_{3}(r).

Although we derived Vc(r)V_{c}(r), there are unknown parameters. To determine them, we chose the values presented in Eq.(5.6). Then, from Eq.(3.4), the cut-off Λc=2.03\Lambda_{c}=2.03 GeV was obtained. Next, assuming that the Cornell potential VCL(r)V_{CL}(r) with (K=0.3,σ=0.18GeV2)(K=0.3,\sigma=0.18\ \mathrm{GeV}^{2}) describes a true potential well at some point rcr_{c}, we required Vc(r)VCL(r)V_{c}(r)\simeq V_{CL}(r) near r=rcr=r_{c}. To realize this requirement, the conditions in Eq.(5.2) are imposed. By solving these conditions, rc0.226fmr_{c}\simeq 0.226\ \mathrm{fm}, the values of mm and Q2Q^{2} in Eq.(5.7), and Vc(r)V_{c}(r) in Eq.(5.8) were obtained.

There are two implicit scales rcr_{c} and R0R_{0} (or R~0\tilde{R}_{0}) in Vc(r)V_{c}(r). To understand them, the effective Coulomb coupling Kef(r)K_{ef}(r) in Eq.(5.3) and the effective string tension σef(r)\sigma_{ef}(r) in Eq.(5.4) were studied. Since V3(r)V_{3}(r) contributes to them, they depend on rr. At r=rcr=r_{c}, Kef(r)K_{ef}(r) becomes maximum and σef(r)\sigma_{ef}(r) becomes minimum. For r>R00.5fmr>R_{0}\simeq 0.5\ \mathrm{fm}, 0.23GeV2<σef(r)<0.25GeV20.23\ \mathrm{GeV}^{2}<\sigma_{ef}(r)<0.25\ \mathrm{GeV}^{2} holds.

If we consider the quantity r2Vc(r)=Kef(r)+r2σef(r)r^{2}V_{c}^{\prime}(r)=K_{ef}(r)+r^{2}\sigma_{ef}(r), we find

r2Vc(r){Kef(r),(r<0.2fm)r2σef(r),(r>0.5fm).r^{2}V_{c}^{\prime}(r)\approx\left\{\begin{array}[]{ll}K_{ef}(r),&\quad(r<0.2\ \mathrm{fm})\\ r^{2}\sigma_{ef}(r),&\quad(r>0.5\ \mathrm{fm}).\end{array}\right.

Namely the main force between QQ and Q¯\bar{Q} is the effective Coulomb force Kef/r2-K_{ef}/r^{2} for r<r0r<r_{0}, and the effective string force σef(r)-\sigma_{ef}(r) for r>R0r>R_{0}.

Although Vc(r)V_{c}(r) was determined to fit to VCL(r)V_{CL}(r) with (K=0.3,σ=0.18GeV2)(K=0.3,\sigma=0.18\ \mathrm{GeV}^{2}) at r=rcr=r_{c}, it becomes larger than VCL(r)V_{CL}(r) for r>0.4fmr>0.4\ \mathrm{fm}. The Cornell potential is often used to fit lattice simulation data. Can we find VCL(r)V_{CL}(r) that fits to Vc(r)V_{c}(r) better? To answer this question, we used the above scales. At rcr_{c}, Eq.(5.3) was applied to determine KK. To determine σ\sigma, we used Eq.(6.4) at R~0\tilde{R}_{0}. Then we obtained VCL(r)V_{CL}(r) with (K=0.3,σ=0.212GeV2)(K=0.3,\sigma=0.212\ \mathrm{GeV}^{2}). This potential satisfies Eq.(6.1) at R0R_{0} as well, and fits fairly well in the region of r<1.2fmr<1.2\ \mathrm{fm}.

We make three comments.

(1). In quark confinement, Abelian dominance ei is expected. The lattice simulation in the maximal Abelian gauge shows that the linear part of the QQ¯Q\bar{Q} potential comes from the Abelian part ss . In the present case, Abelian dominance is realized by the massive classical U(1) field μ\mathcal{B}_{\mu}. This field brings about the potential Vc(r)V_{c}(r).

(2) Although Vc(r)V_{c}(r) comes from μ\mathcal{B}_{\mu}, to determine its behavior, information on the fields aμAa_{\mu}^{A} for ΛQCD<μ<Λc\Lambda_{\mathrm{QCD}}<\mu<\Lambda_{c} is necessary. The low energy Lagrangian (2.9) will be modified in this region. However, as the first approximation, we used Eq. (2.9) and assumed the cut-off Λc\Lambda_{c} for mm.

(3). The values of rcr_{c}, R~0\tilde{R}_{0}, mm and Q2Q^{2} in Eq.(5.7) depend on the choice of Eq.(5.6). However the existence of these scales, and the behavior of Vc(r)V_{c}(r), e.g., limrV3(r)=0\lim_{r\to\infty}V_{3}(r)=0, σc>σ\sigma_{c}>\sigma, Kc<KK_{c}<K, and Vc(r)>VCL(r)V_{c}(r)>V_{CL}(r) for large rr, are unchanged.

Appendix A Propagator Gμν(0)(y,x)G_{\mu\nu}^{(0)}(y,x) for AμaA_{\mu}^{a}

Referring to Eqs.(2.2) and (2.9), we consider the Lagrangian with the massive gauge fields Aμ±=(Aμ1±iAμ2)/2A_{\mu}^{\pm}=(A_{\mu}^{1}\pm iA_{\mu}^{2})/\sqrt{2}:

a=12{14(Aa)μν2+M22(Aμa)2α12(Ba)2+Ba(μAμa+φa)+(φa)22α2}.\sum_{a=1}^{2}\left\{\frac{1}{4}(\partial\wedge A^{a})_{\mu\nu}^{2}+\frac{M^{2}}{2}(A_{\mu}^{a})^{2}-\frac{\alpha_{1}}{2}(B^{a})^{2}+B^{a}(\partial_{\mu}A_{\mu}^{a}+\varphi^{a})+\frac{(\varphi^{a})^{2}}{2\alpha_{2}}\right\}.

The fields AμA_{\mu}, BB and φ\varphi mix. The inverse propagators of these fields are

AνBφAμ( (p2+M2)PμνT+M2PμνLipμ0) Bipνα11φ011α2,\bordermatrix{&A_{\nu}&B&\varphi\cr A_{\mu}&(p^{2}+M^{2})P^{T}_{\mu\nu}+M^{2}P^{L}_{\mu\nu}&-ip_{\mu}&0\cr B&ip_{\nu}&-\alpha_{1}&1\cr\varphi&0&1&\frac{1}{\alpha_{2}}}, (A.1)

and the corresponding propagators are

AνBφAμ( 1p2+M2PμνT+(α1+α2)ΞPμνLipμΞipμα2Ξ) BipνΞM2Ξα2M2Ξφipνα2Ξα2M2Ξα2(p2+α1M2)Ξ,\bordermatrix{&A_{\nu}&B&\varphi\cr A_{\mu}&\frac{1}{p^{2}+M^{2}}P^{T}_{\mu\nu}+\frac{(\alpha_{1}+\alpha_{2})}{\Xi}P^{L}_{\mu\nu}&\frac{-ip_{\mu}}{\Xi}&\frac{ip_{\mu}\alpha_{2}}{\Xi}\cr B&\frac{ip_{\nu}}{\Xi}&\frac{-M^{2}}{\Xi}&\frac{\alpha_{2}M^{2}}{\Xi}\cr\varphi&\frac{-ip_{\nu}\alpha_{2}}{\Xi}&\frac{\alpha_{2}M^{2}}{\Xi}&\frac{\alpha_{2}(p^{2}+\alpha_{1}M^{2})}{\Xi}}, (A.2)

where Ξ=p2+(α1+α2)M2\Xi=p^{2}+(\alpha_{1}+\alpha_{2})M^{2}, and

PμνT=δμνpμpνp2,PμνL=pμpνp2.P^{T}_{\mu\nu}=\delta_{\mu\nu}-\frac{p_{\mu}p_{\nu}}{p^{2}},\quad P^{L}_{\mu\nu}=\frac{p_{\mu}p_{\nu}}{p^{2}}.

Under the BRS transformation δB\delta_{B}, as δBc¯=iB\delta_{B}\bar{c}=iB and δBB=0\delta_{B}B=0,

BB=iδB(c¯B)=0\langle BB\rangle=-i\langle\delta_{B}(\bar{c}B)\rangle=0

holds, if the BRS symmetry is not broken spontaneously. When M0M\neq 0, to make

BB=M2p2+(α1+α2)M2\langle BB\rangle=\frac{-M^{2}}{p^{2}+(\alpha_{1}+\alpha_{2})M^{2}}

vanish, we choose α2\alpha_{2}\to\infty. Then Eq.(A.2) becomes

AνBφAμ( 1p2+M2PμνT+1M2PμνL0ipμM2) B001φipνM21p2+α1M2)M2..\bordermatrix{&A_{\nu}&B&\varphi\cr A_{\mu}&\frac{1}{p^{2}+M^{2}}P^{T}_{\mu\nu}+\frac{1}{M^{2}}P^{L}_{\mu\nu}&0&\frac{ip_{\mu}}{M^{2}}\cr B&0&0&1\cr\varphi&\frac{-ip_{\nu}}{M^{2}}&1&\frac{p^{2}+\alpha_{1}M^{2})}{M^{2}}}.. (A.3)

Eq.(A.3) shows that AμA_{\mu} mixes with φ\varphi. At the one-loop order, the ghost loop contributes to AμAν\langle A_{\mu}A_{\nu}\rangle, Aμφ\langle A_{\mu}\varphi\rangle and φφ\langle\varphi\varphi\rangle as well. However tachyonic behavior only appears in AμAν\langle A_{\mu}A_{\nu}\rangle. In Sect. 3, as the mixing like Aμφ\langle A_{\mu}\varphi\rangle does not contribute, the propagator

Gμν(0)(p)=1p2+M2PμνT+1M2PμνLG_{\mu\nu}^{(0)}(p)=\frac{1}{p^{2}+M^{2}}P^{T}_{\mu\nu}+\frac{1}{M^{2}}P^{L}_{\mu\nu}

in Eq.(A.3) is used.

Appendix B Solution of Eqs.(5.3)-(5.5)

From Eq.(4.9), we find the derivatives

V3(r)\displaystyle V_{3}^{\prime}(r) =Q2m22π20Λc𝑑q(cosqrrsinqrqr2)1q2+m2,\displaystyle=Q^{2}\frac{m^{2}}{2\pi^{2}}\int_{0}^{\Lambda_{c}}dq\left(\frac{\cos qr}{r}-\frac{\sin qr}{qr^{2}}\right)\frac{1}{q^{2}+m^{2}},
V3′′(r)\displaystyle V_{3}^{\prime\prime}(r) =Q2m22π20Λc𝑑q(2sinqrqr32cosqrr2qsinqrr)1q2+m2,\displaystyle=Q^{2}\frac{m^{2}}{2\pi^{2}}\int_{0}^{\Lambda_{c}}dq\left(2\frac{\sin qr}{qr^{3}}-2\frac{\cos qr}{r^{2}}-\frac{q\sin qr}{r}\right)\frac{1}{q^{2}+m^{2}},
(r3V3′′(r))\displaystyle\left(r^{3}V_{3}^{\prime\prime}(r)\right)^{\prime} =Q2m22π20Λc𝑑q(q2r2cosqr)1q2+m2.\displaystyle=Q^{2}\frac{m^{2}}{2\pi^{2}}\int_{0}^{\Lambda_{c}}dq\left(-q^{2}r^{2}\cos qr\right)\frac{1}{q^{2}+m^{2}}.

Then, introducing the variables x=qrx=qr and a=m2/Λc2a=m^{2}/\Lambda_{c}^{2}, Eqs.(5.3)-(5.5) become

K=Q24π[1+Λc2πrc2{2H(rc,a)+L(rc,a)}],\displaystyle K=\frac{Q^{2}}{4\pi}\left[1+\frac{\Lambda_{c}^{2}}{\pi}r_{c}^{2}\left\{2H(r_{c},a)+L(r_{c},a)\right\}\right], (B.1)
σ=Q24π{Λc22aln(1+1a)Λc2πL(rc,a)},\displaystyle\sigma=\frac{Q^{2}}{4\pi}\left\{\frac{\Lambda_{c}^{2}}{2}a\ln\left(1+\frac{1}{a}\right)-\frac{\Lambda_{c}^{2}}{\pi}L(r_{c},a)\right\}, (B.2)
G(rc,a)=0Λcrc𝑑xx2cosxx2+a(Λcrc)2=0,\displaystyle G(r_{c},a)=\int_{0}^{\Lambda_{c}r_{c}}dx\frac{x^{2}\cos x}{x^{2}+a(\Lambda_{c}r_{c})^{2}}=0, (B.3)

where H(r,a)H(r,a) is defined in Eq.(6.3), and

L(r,a)=0Λcr𝑑xxsinxax2+a(Λcr)2.L(r,a)=\int_{0}^{\Lambda_{c}r}dxx\sin x\frac{a}{x^{2}+a(\Lambda_{c}r)^{2}}.

By eliminating Q2Q^{2}, Eqs.(B.1) and (B.2) leads to

F(rc,a)=σ+2σΛc2πrc2H(rc,a)+Λc2π(σrc2+K)L(rc,a)KΛc22aln(1+1a)=0.F(r_{c},a)=\sigma+2\sigma\frac{\Lambda_{c}^{2}}{\pi}r_{c}^{2}H(r_{c},a)+\frac{\Lambda_{c}^{2}}{\pi}\left(\sigma r_{c}^{2}+K\right)L(r_{c},a)-K\frac{\Lambda_{c}^{2}}{2}a\ln\left(1+\frac{1}{a}\right)=0. (B.4)
Refer to caption
Figure 8: The behavior of the function G(r,a)G(r,a). G2=G(r,0.2)G2=G(r,0.2), G3=G(r,0.3)G3=G(r,0.3) and Gs=G(r,0.263)Gs=G(r,0.263) are plotted. The unit of rr is GeV-1. Gs=0Gs=0 holds at r1.145GeV1r\simeq 1.145\ \mathrm{GeV}^{-1}.
Refer to caption
Figure 9: The behavior of the function F(r,a)F(r,a). F2=F(r,0.2)F2=F(r,0.2), F3=F(r,0.3)F3=F(r,0.3) and Fs=F(r,0.263)Fs=F(r,0.263) are plotted. The unit of rr is GeV-1. Fs=0Fs=0 holds at r1.145GeV1r\simeq 1.145\ \mathrm{GeV}^{-1}.

Now we substitute K=0.3,σ=0.18GeV2K=0.3,\sigma=0.18\ \mathrm{GeV}^{2} and Λc=2.03GeV\Lambda_{c}=2.03\ \mathrm{GeV}. To solve Eq.(B.3) and Eq.(B.4) numerically, choosing a=0.2,0.263a=0.2,0.263 and 0.30.3, G(r,a)G(r,a) and F(r,a)F(r,a) are plotted in Fig. B1 and Fig. B2, respectively. We find that G(rc,a)=0G(r_{c},a)=0 and F(rc,a)=0F(r_{c},a)=0 give the solutions a0.263a\simeq 0.263 and rc1.145GeV1r_{c}\simeq 1.145\mathrm{GeV}^{-1}.

Using these values, Eqs.(B.1) and (B.2) give

Kc=Q24π0.285,σc=Q24πΛc22aln(1+1a)0.242GeV2.K_{c}=\frac{Q^{2}}{4\pi}\simeq 0.285,\quad\sigma_{c}=\frac{Q^{2}}{4\pi}\frac{\Lambda_{c}^{2}}{2}a\ln\left(1+\frac{1}{a}\right)\simeq 0.242\ \mathrm{GeV}^{2}.

In the same way, using S(r,a)S(r,a) in Eq.(5.9), Eq.(4.9) becomes

V3(r)=2Q24πΛc2πS(r,0.263)=0.747S(r,0.263).V_{3}(r)=2\frac{Q^{2}}{4\pi}\frac{\Lambda_{c}^{2}}{\pi}S(r,0.263)=0.747\cdot S(r,0.263). (B.5)

References

  • (1) G. Bali, Phys. Rept. 343, 1 (2001).
  • (2) E. Eichten et al., Phys. Rev. Lett. 34, 369 (1975).
  • (3) E. Eichten et al., Phys. Rev. D 21, 203 (1980).
  • (4) G. Ripka, arXiv:hep-ph/0310102.
  • (5) T. Suzuki, Prog. Theor. Phys. 80, 929 (1988).
  • (6) S. Maedan and T. Suzuki, Prog. Theor. Phys. 81, 229 (1989).
  • (7) S. Sasaki, H. Suganuma and H. Toki, Prog. Theor. Phys. 94, 373 (1995).
  • (8) H. Suganuma, S. Sasaki and H. Toki, Nucl. Phys. B 435, 207 (1995).
  • (9) H. Sawayanagi, Prog. Theor. Exp.  Phys. 2019, 033B03 (2019).
  • (10) H. Sawayanagi, Prog. Theor. Exp.  Phys. 2017, 113B02 (2017).
  • (11) R. Sommer, Nucl. Phys. B 411, 839 (1994).
  • (12) G. Curci and R. Ferrari, Phys. Lett. B63, 91 (1976).
  • (13) H. Sawayanagi, Phys. Rev. D 67, 045002 (2003).
  • (14) H. Sawayanagi, Prog. Theor. Phys. 117, 305 (2007).
  • (15) D. Zwanziger, Phys. Rev. D 3, 880 (1971).
  • (16) K. Amemiya and H. Suganuma, Phys. Rev. D 60, 114509 (1999).
  • (17) Z. F. Ezawa and A. Iwazaki, Phys. Rev. D 25, 2681 (1982).
  • (18) N. Sakumichi and H. Suganuma, Phys. Rev. D 90, 111501(R) (2014).