1:
2: \documentclass{JHEP3}
3: \usepackage{amsmath}
4:
5: \title{String breaking in QCD: dual superconductor {\it vs.} stochastic vacuum model}
6:
7: \author{D. Antonov
8: \thanks{Permanent address:
9: ITEP, B. Cheremushkinskaya 25, RU-117 218 Moscow, Russia.}\\
10: Institute of Physics, Humboldt University of Berlin,\\
11: Newtonstr. 15, 12489 Berlin, Germany \\
12: E-mail: \email{antonov@physik.hu-berlin.de}}
13:
14:
15:
16:
17: \author{A. Di Giacomo\\
18: Dipartimento di Fisica ``E. Fermi'' dell'Universit\'a di Pisa,\\
19: INFN-Sezione di Pisa,\\
20: Largo Pontecorvo, 3, 56127 Pisa, Italy \\
21: E-mail: \email{digiaco@df.unipi.it}}
22:
23: \abstract{Effects of dispersion of the chromoelectric field of the flux tube on the
24: string-breaking distance are studied. The leading-order correction is shown to slightly diminish the result
25: following from the Schwinger formula. Instead, accounting for corrections of all orders might
26: result, at certain values of the Landau-Ginzburg parameter, in an increase of the
27: string-breaking distance up to one order of magnitude.
28: An alternative formula for this distance is obtained when produced pairs are treated as
29: holes in a confining pellicle, which spans over the contour of an external quark-antiquark pair. Generalizations of the
30: obtained results to the cases of small temperatures, as well as temperatures close to the critical one are also discussed.}
31:
32: \keywords{Nonperturbative Effects; Confinement; Phenomenological Models;
33: Lattice Gauge Field Theories}
34:
35: \preprint{HU-EP-05/01 \\
36: IFUP-TH 2005/01}
37:
38: \dedicated{}
39:
40: \begin{document}
41:
42:
43: \section{Introduction. The model and some relations between its parameters}
44: String breaking at zero and finite temperatures is known as one of the important open problems in QCD (see e.g. refs.~\cite{a,b,c,d,e,1,f}
45: for recent developments).
46: The aim of this paper is to proceed further with the calculation of the string-breaking distance, along with the
47: lines of ref.~\cite{1} (Sections~2-6), as well as using some alternative model (Section~7). In ref.~\cite{1},
48: dual Abrikosov-Nielsen-Olesen strings have been used to model flux tubes of the chromoelectric field in real QCD.
49: Such strings~\cite{15}
50: are solutions to the classical equations of motion in the 4d dual Abelian Higgs model, whose Euclidean Lagrangian reads
51: ${\cal L}=\frac14F_{\mu\nu}^2+|D_\mu\varphi|^2+\frac{\lambda}{2}(|\varphi|^2-v^2)^2$. Here
52: $F_{\mu\nu}=\partial_\mu B_\nu-\partial_\nu B_\mu$, $D_\mu\varphi=(\partial_\mu-ig_mB_\mu)\varphi$,
53: $B_\mu$ is the dual gauge field, $\varphi$ is the complex-valued dual-Higgs field, and $g_m$ is the magnetic
54: coupling constant, related to the electric one, $g$, as $g_m=2\pi/g$. The masses of the dual vector boson and
55: the dual Higgs boson are $m_V=\sqrt{2}g_m v$ and $m_H=\sqrt{2\lambda}v$, respectively.
56:
57: We will consider this model in the
58: London limit~\footnote{The ratio $\frac{m_H}{m_V}$ is called Landau-Ginzburg parameter.},
59: $L\equiv\ln\frac{m_H}{m_V}\gg 1$, where all the
60: results can be obtained analytically. The Bogomol'nyi case, $L=0$, will be considered elsewhere.
61: The electric field of a straight-line dual Abrikosov-Nielsen-Olesen string reads $E(r)=\frac{m_V^2}{g_m}K_0(m_Vr)$,
62: where $r=|{\bf x}_\perp|$ with ${\bf x}_\perp=(x_1, x_2)$ denoting
63: the direction transverse to the string, and from now on $K_\nu$'s stand for
64: MacDonald functions.
65: The field averaged over the string cross section,
66: $\left<E\right>=\frac{1}{S}\int d^2r E(r)$ obeys the relation $g\left<E\right>=4\sigma/L$. Here
67: $S=\pi m_V^{-2}$ is the area of the cross section of the string,
68: and the string tension has the form $\sigma=2\pi v^2L$. In order to establish a
69: correspondence to QCD, we will need to express all the results in terms of $\sigma$ and $g$. Two formulae useful for this
70: are $m_V^2=\frac{4\pi\sigma}{g^2L}$ and $S=\frac{g^2L}{4\sigma}$.
71:
72:
73:
74:
75:
76: \section{The string-breaking distance with the neglection of dispersion of $E(r)$}
77:
78: The rate of the pair production reads $w=\frac{2}{S}{\,}{\rm Im}{\,}\Gamma[A_i]$,
79: where the one-loop effective action of a scalar quark
80: has the form:
81:
82: $$
83: \Gamma[A_i]=N_cN_f\int\limits_{0}^{\infty}\frac{dT}{T}{\rm e}^{-m^2T}\times$$
84:
85: \begin{equation}
86: \label{1}
87: \times\int {\cal D}{\bf x}_\perp {\cal D}{\bf x}_\parallel
88: \exp\left[-\int\limits_{0}^{T}d\tau\left(\frac14\dot{\bf x}_\perp^2+\frac14\dot{\bf x}_\parallel^2-\frac{ig}{2}
89: E({\bf x}_\perp(\tau))
90: \epsilon_{ij}\dot x_i x_j\right)\right].
91: \end{equation}
92: Here, ${\bf x}_\parallel=(x_3, x_4)$, $i,j=3,4$,
93: and the field of the string is $A_i=-\frac12\epsilon_{ij}x_jE({\bf x}_\perp)$. Note that,
94: in this paper, we will study the case of scalar quarks only (except of some comment at the end of the next Section), since
95: the longitudinal part of the spin factor, $\int{\cal D}\psi_i\exp\left[-\int\limits_{0}^{T}d\tau\left(\frac12\psi_i
96: \dot\psi_i-ig\psi_i\psi_j F_{ij}(\tau)\right)\right]$,
97: where $F_{34}(\tau)=-F_{43}(\tau)=E({\bf x}_\perp(\tau))$, cannot be calculated exactly
98: as long as $F_{ij}$ is $\tau$-dependent. To evaluate the path integral~(\ref{1}), we will use the requirement that the
99: mass of the produced pair, $m$, should be much larger than $m_V$, i.e. $m\gg\frac2g\sqrt{\frac{\pi\sigma}{L}}$,
100: in order that the field of a string can be considered as a constant one. The ``largeness'' of $m$ means that characteristic
101: proper times are ``small'', $T<\frac{1}{m^2}$, that enables us to evaluate the path integral semiclassically and to
102: compute further the leading quantum correction using the
103: Feynman variational method~\cite{2}. Furthermore, we naturally assume that not only the Compton wavelength of a
104: produced pair, $m^{-1}$, is much smaller than the range of the field localization, $m_V^{-1}$, but also that the
105: characteristic pair trajectories are small compared to $m_V^{-1}$. Since, in the Euclidean space-time, pair trajectories
106: are circles of the radius $R_p=\frac{m}{g\left<E\right>}$~\footnote{This can be seen either by solving the respective
107: Euler-Lagrange equation~\cite{3}, or simply by noticing that $\exp\left(-\frac{\pi m^2}{g\left<E\right>}\right)$ in the
108: Schwinger formula should be ${\rm e}^{-\Phi}$, where $\Phi$ is the flux of the electric field through the contour
109: of a pair, $\Phi=g\left<E\right>\cdot \pi R_p^2$. Therefore, a pair is identified with a quark which moves along a
110: circle of the radius $R_p$.}, the condition of smallness of the pair trajectory, $R_p\ll m_V^{-1}$,
111: yields $m\ll 2g\sqrt{\frac{\sigma}{\pi L}}$. Both conditions,
112:
113: \begin{equation}
114: \label{ineq}
115: \frac2g\sqrt{\frac{\pi\sigma}{L}}\ll m\ll 2g\sqrt{\frac{\sigma}{\pi L}},
116: \end{equation}
117: are compatible to each other at $g\gg 1$.
118:
119: Then, due to the smallness of
120: a pair trajectory, $E({\bf x}_\perp(\tau))$ can be replaced by its value averaged along the trajectory:
121:
122: \begin{equation}
123: \label{smallT}
124: \int\limits_{0}^{T}d\tau E({\bf x}_\perp(\tau))\dot x_i x_j\simeq -\Sigma_{ij}
125: \cdot\frac{1}{T}\int\limits_{0}^{T}d\tau E({\bf x}_\perp(\tau)),
126: \end{equation}
127: where $\Sigma_{ij}\equiv\int\limits_{0}^{T}d\tau x_i \dot x_j$ is the $(i,j)$-th component of the so-called tensor area
128: of the trajectory. In the leading small-$T$ approximation, we obtain the classical result for $\int {\cal D}{\bf x}_\perp=
129: \int d^2x_\perp(0)\int\limits_{{\bf x}_\perp(0)={\bf x}_\perp(T)}^{}{\cal D}{\bf x}_\perp(\tau)$
130: in eq.~(\ref{1}):
131:
132: \begin{equation}
133: \label{class}
134: \frac{1}{4\pi T}\int d^2x_\perp\exp\left[-\frac{ig}{2}E({\bf x}_\perp)\epsilon_{ij}\Sigma_{ij}\right].
135: \end{equation}
136: Thus,
137:
138: \begin{equation}
139: \label{eff}
140: \Gamma[A_i]\simeq\frac{N_cN_f}{4\pi}\int\limits_{0}^{\infty}\frac{dT}{T^2}{\rm e}^{-m^2T}\int {\cal D}{\bf x}_\parallel
141: \exp\left(-\frac14\int\limits_{0}^{T}d\tau\dot{\bf x}_\parallel^2\right)\int d^2x_\perp\exp\left[-\frac{ig}{2}E({\bf x}_\perp)
142: \epsilon_{ij}\Sigma_{ij}\right].
143: \end{equation}
144: Neglecting for this Section the dispersion of the field $E({\bf x}_\perp)$, we have
145:
146:
147: \begin{equation}
148: \label{0}
149: \int d^2x_\perp\exp\left[-\frac{ig}{2}E({\bf x}_\perp)
150: \epsilon_{ij}\Sigma_{ij}\right]\equiv S\left<\exp\left[-\frac{ig}{2}E({\bf x}_\perp)
151: \epsilon_{ij}\Sigma_{ij}\right]\right>\simeq S\exp\left[-\frac{ig}{2}\left<E\right>
152: \epsilon_{ij}\Sigma_{ij}\right].
153: \end{equation}
154: In this approximation, we therefore
155: arrive at the Euler-Heisenberg Lagrangian in the constant field $\bar A_i\equiv-\frac12\epsilon_{ij}x_j
156: \left<E\right>$,
157:
158: \begin{equation}
159: \label{effact}
160: \Gamma[A_i]\simeq S\frac{N_cN_f}{(4\pi)^2}\int\limits_{0}^{\infty}\frac{dT}{T^2}{\rm e}^{-m^2T}
161: \frac{g\left<E\right>}{\sin(g\left<E\right>T)},
162: \end{equation}
163: and recover for $w$ the 4d Schwinger result~\footnote{Up to the factor $N_cN_f$ absent
164: in the electromagnetic case.} in the bosonic case:
165:
166: \begin{equation}
167: \label{w}
168: w=N_cN_f\frac{(g\left<E\right>)^2}{(2\pi)^3}\sum\limits_{k=1}^{\infty}\frac{(-1)^{k+1}}{k^2}
169: \exp\left(-\frac{\pi k m^2}{g\left<E\right>}\right).
170: \end{equation}
171: We can further express the inequality $T<\frac{1}{m^2}$ in terms of the parameters of our model.
172: Namely, since at least the first pole in the imaginary part of the Euler-Heisenberg Lagrangian should give its
173: contribution, $T$ may not be arbitrarily small, but the following inequality should rather hold:
174: $T>\frac{\pi}{g\left<E\right>}=\frac{\pi L}{4\sigma}$. The condition $\frac{1}{m^2}>T$ then yields
175: $m<2\sqrt{\frac{\sigma}{\pi L}}$, that also coincides with
176: the condition for $w$ not to be exponentially small. This new constraint
177: is stronger than the above-obtained one, expressed by the right inequality of~(\ref{ineq}),
178: since $g\gg 1$ is now absent.
179: The new constraint can be viewed as an upper boundary on $L$:
180:
181: \begin{equation}
182: \label{L}
183: L<\frac{4}{\pi}\frac{\sigma}{m^2}\simeq 1.27\frac{\sigma}{m^2}.
184: \end{equation}
185: Setting $\sigma=(440{\,}{\rm MeV})^2$ and a typical hadronic mass $m=200{\,}{\rm MeV}$,
186: we get an estimate $L<6.2$, that still leaves a window for $L\gg 1$.
187:
188: Approximating the whole sum~(\ref{w}) by its first term (equal to the density of produced pairs), we have
189:
190: \begin{equation}
191: \label{ww}
192: w\simeq\frac{2N_cN_f}{\pi^3}\left(\frac{\sigma}{L}\right)^2\exp\left(-\frac{\pi
193: m^2L}{4\sigma}\right).
194: \end{equation}
195: The respective string-breaking distance~\cite{1} $\bar r=\frac{1}{\sqrt{2Sw}}$ has the form
196:
197: \begin{equation}
198: \label{rr}
199: \bar r=\frac{\pi^{3/2}\sqrt{L}}{g\sqrt{N_cN_f\sigma}}\exp\left(\frac{\pi L}{8}\frac{m^2}{\sigma}\right),
200: \end{equation}
201: where, due to inequality~(\ref{L}),
202: $\exp\left(\frac{\pi L}{8}\frac{m^2}{\sigma}\right)<\sqrt{{\rm e}}\simeq 1.65$.
203:
204:
205:
206:
207: \section{Finite-temperature generalizations}
208: Setting for the mass of a produced pair the $\pi$-meson mass,
209: we can further extend the analysis of the previous Section to the case where the temperature is close to the critical one.
210: This can be done by using the formulae~\cite{1, 4} $m^2=m_\pi^2t^{1.44}$,
211: $\sigma=\sigma_0t^{0.33}$, where $t\equiv 1-\frac{T}{T_c}$ is the reduced temperature~\footnote{The fact that
212: the temperature and the proper time are denoted by the same letter ``$T$'' should not lead to reader's confusions.}, and
213: $\sigma_0\simeq (440{\,}{\rm MeV})^2$.
214: We obtain:
215:
216: \begin{equation}
217: \label{tto0}
218: w\to\frac{2N_cN_f}{\pi^3}\left(\frac{\sigma}{L}\right)^2={\cal O}\left(t^{0.66}\right),
219: \end{equation}
220: that establishes the law by which $\bar r$ grows at $t\to 0$.
221: In the same limit $t\to 0$, the condition $mr_\perp\gg 1$ with
222: the temperature-dependent $r_\perp$, $r_\perp=\frac{g}{2}\sqrt{\frac{L}{\pi\sigma}}$, leads to the following
223: boundary on $t$ from below: $\frac1L\left(\frac{1}{m_\pi}\sqrt{\frac{\sigma_0}{\alpha_s}}\right)^{1.79}\ll t$.
224: On the other hand,
225: the condition of smallness of the
226: pair trajectory, expressed by the right inequality of~(\ref{ineq}), sets a boundary on $t$ from above:
227: $t\ll\frac{1}{L^{0.91}}\left(\frac{2g}{m_\pi}\sqrt{\frac{\sigma_0}{\pi}}\right)^{1.82}$.
228: These two conditions imposed on $t$ are apparently compatible to each other at sufficiently large $L$ and/or $g$.
229: Note also that the condition~(\ref{L}) becomes
230: softer as one approaches the critical point, since $\frac{\sigma}{m^2}=\frac{\sigma_0}{m_\pi^2}t^{-1.11}\to\infty$.
231:
232: Let us now evaluate the string-breaking
233: distance at relatively low temperatures, namely at $T={\cal O}(f_\pi)$, where the following
234: formula holds~\cite{GL}:
235:
236: \begin{equation}
237: \label{mT}
238: m_\pi^2(T)\simeq m_\pi^2\left(1+\frac{T^2}{24f_\pi^2}\right).
239: \end{equation}
240: We can further approximate
241: $m_V(T)$ by $m_V(0)$, since $m_V^{-1}$ is the vacuum correlation length, whose QCD-analogue at such temperatures
242: can be considered as temperature-independent \cite{correl} \footnote{In QCD, the (magnetic) vacuum
243: correlation length becomes definitely temperature-dependent at temperatures larger than the temperature
244: of dimensional reduction (that is of the order of $2T_c$), where it is proportional to $\frac{1}{g^2(T)T}$.}.
245:
246:
247: Next, we need to know the temperature dependence of $g\left<E\right>$. Since we are exploring the region of
248: temperatures smaller than the temperature of dimensional
249: reduction, the sum over Matsubara frequencies appears:
250:
251: $$g\left<E\right>(T)=\frac{(gm_V)^2}{2\pi^2}\int d^2z\sum\limits_{n=-\infty}^{+\infty}
252: K_0\left(\sqrt{z^2+(m_V\beta n)^2}\right),$$
253: where $\beta\equiv 1/T$. To carry out the integral, it is convenient to transform
254: the sum as follows:
255:
256: $$
257: \sum\limits_{n=-\infty}^{+\infty}K_0\left(\sqrt{z^2+(m_V\beta n)^2}\right)=\frac12\sum\limits_{n=-\infty}^{+\infty}
258: \int\limits_{0}^{\infty}\frac{dt}{t}\exp\left[-\frac{1}{4t}-t\left(z^2+(m_V\beta n)^2\right)\right]=$$
259:
260: \begin{equation}
261: \label{transform}
262: =\frac{T\sqrt{\pi}}{2m_V}\sum\limits_{n=-\infty}^{+\infty}\int\limits_{0}^{\infty}\frac{dt}{t^{3/2}}
263: \exp\left[-z^2t-\frac{1+(\omega_n/m_V)^2}{4t}\right]=
264: \frac{\pi T}{m_V}\sum\limits_{n=-\infty}^{+\infty}
265: \frac{{\rm e}^{-|z|\sqrt{1+(\omega_n/m_V)^2}}}{\sqrt{1+(\omega_n/m_V)^2}},
266: \end{equation}
267: where $\omega_n\equiv2\pi Tn$.
268: The integration over $d^2z$ then immediately yields
269:
270: \begin{equation}
271: \label{gET}
272: g\left<E\right>(T)=g^2m_VT\sum\limits_{n=-\infty}^{+\infty}
273: \frac{1}{\left[1+(\omega_n/m_V)^2\right]^{3/2}}.
274: \end{equation}
275: To study the zero-temperature limit of this expression, one should perform the inverse transformation of the sum:
276:
277: $$
278: \sum\limits_{n=-\infty}^{+\infty}
279: \frac{1}{\left[1+(\omega_n/m_V)^2\right]^{3/2}}=\frac{2}{\sqrt{\pi}}
280: \int\limits_{0}^{\infty}dt\sqrt{t}
281: {\rm e}^{-t}\sum\limits_{n=-\infty}^{+\infty}{\rm e}^{-t(\omega_n/m_V)^2}=$$
282:
283: $$=\frac{m_V\beta}{\pi}\sum\limits_{n=-\infty}^{+\infty}
284: \int\limits_{0}^{\infty}dt\exp\left[-t-\frac{(m_V\beta n)^2}{4t}\right]=
285: \frac{(m_V\beta)^2}{\pi}\sum\limits_{n=-\infty}^{+\infty}|n|K_1(m_V\beta |n|).$$
286: At small $T$'s of interest, the sum here is apparently dominated by the zeroth mode.
287: Moreover, expanding the function $K_1$ in power series, one can check that no finite corrections exist
288: to the r.h.s. of the formula $|n|K_1(m_V\beta |n|)\stackrel{|n|\to 0}{\longrightarrow}T/m_V$. Therefore,
289: the leading small-$T$ (physically, at $T\ll f_\pi$) correction to $g\left<E\right>(T)$
290: stems from the terms in the sum with $|n|=1$, $2K_1(m_V\beta)$. This correction is therefore
291: exponentially small, as confirmed by the following final expression:
292:
293: \begin{equation}
294: \label{GE}
295: g\left<E\right>(T)\simeq\frac{4\sigma_0}{L}\left[1+2\sqrt{\frac{\beta}{g}}\left(\frac{\pi^3\sigma_0}{L}\right)^{1/4}
296: {\rm e}^{-\frac2g\sqrt{\frac{\pi\sigma_0}{L}}\beta}\right],
297: \end{equation}
298: where $\sigma_0\equiv2\pi v^2L$. As for the dependence $\sigma(T)$, it is derived in Appendix~B.
299:
300:
301: Equations~(\ref{mT}) and (\ref{GE}), being substituted into the formula
302: $w\simeq N_cN_f\frac{(g\left<E\right>)^2}{(2\pi)^3}{\rm e}^{-\frac{\pi m^2}{g\left<E\right>}}$,
303: determine the temperature dependence of $w$. Apparently, the correction produced by eq.~(\ref{GE})
304: is exponentially small with respect to the temperature dependence appearing by means of eq.~(\ref{mT}). Therefore,
305: $w$ decreases at $T$ increasing from zero to the temperatures of the order of $f_\pi$.
306: The corresponding increase of $\bar r$ parallels the same phenomenon we have found at $T\to T_c$.
307:
308:
309: It is finally worth making the following comment. In the case where the dispersion of $E(r)$ is
310: neglected (that we are discussing here), the spin
311: factor can be evaluated, and it is natural to address the issue of how strongly the antiperiodic boundary conditions
312: for spin-$\frac12$ quarks can
313: affect the obtained result. For such quarks, one should substitute in eq.~(\ref{effact}) $\frac{1}{\sin(g\left<E\right>T)}\to
314: -\frac{2}{\tan(g\left<E\right>T)}$, that yields, instead of eq.~(\ref{w}),
315:
316: \begin{equation}
317: \label{Wferm}
318: w=N_cN_f\frac{(g\left<E\right>)^2}{4\pi^3}\sum\limits_{k=1}^{\infty}\frac{1}{k^2}
319: \exp\left(-\frac{\pi k m^2}{g\left<E\right>}\right).
320: \end{equation}
321: It has been shown in ref.~\cite{bras} that
322: antiperiodic boundary conditions for fermions can be taken into account upon the multiplication
323: of the zero-temperature one-loop effective action by the factor
324: $\left[1+2\sum\limits_{n=1}^{\infty}(-1)^n{\rm e}^{-\frac{\beta^2n^2}{4T}}\right]$, where (only for this formula in this Section)
325: $T$ stands for the
326: proper time. In course of taking ${\rm Im}{\,}\Gamma[A_i]$, this factor transforms to another one, by which a $k$-th
327: term of the series in eq.~(\ref{Wferm}) should be multiplied:
328: $\left[1+2\sum\limits_{n=1}^{\infty}(-1)^n\exp\left(-\frac{g\left<E\right>\beta^2n^2}{4\pi k}\right)\right]$.
329: For relevant $k$'s, which are $k<\frac{g\left<E\right>}{\pi m^2}$, we obtain the following
330: relevant $n$'s: $n<2T\sqrt{\frac{\pi k}{g\left<E\right>}}=\frac{2T}{m}$, that, at $T<f_\pi$, are
331: smaller than 2. Therefore, the factor produced by the antiperiodic boundary conditions
332: for spin-$\frac12$ quarks reduces to
333: $\left[1-2\exp\left(-\frac{g\left<E\right>\beta^2}{4\pi k}\right)\right]$. Since, for the above-mentioned relevant $k$'s,
334: $\frac{g\left<E\right>\beta^2}{4\pi k}>\left(\frac{m\beta}{2}\right)^2$,
335: at $T<\frac{m_\pi}{\sqrt{2}}$ the obtained correction falls off as the tail of the
336: Gaussian distribution~\footnote{The inequality $T<\frac{m_\pi}{\sqrt{2}}$ stems from the fact that the
337: width of the Gaussian distribution ${\rm e}^{-m^2\beta^2/4}$ is $\frac{\sqrt{2}}{m}$.}.
338:
339:
340:
341:
342:
343:
344:
345:
346:
347:
348: \section{Significance of corrections due to the dispersion of $E(r)$}
349: Let us now take into account the second cumulant in eq.~(\ref{0}) [i.e. the dispersion of $E(r)$]
350: and also address the issue of convergence of the cumulant expansion. We have
351:
352: $$\left<\exp\left[-\frac{ig}{2}E({\bf x}_\perp)
353: \epsilon_{ij}\Sigma_{ij}\right]\right>\simeq \exp\left[
354: -\frac{ig}{2}\left<E\right>
355: \epsilon_{ij}\Sigma_{ij}-\frac{g^2}{8}(\epsilon_{ij}\Sigma_{ij})^2\left(\left<E^2\right>-\left<E\right>^2\right)\right]=$$
356:
357: \begin{equation}
358: \label{A}
359: =\frac{1}{2\sqrt{\pi c}}\int\limits_{-\infty}^{+\infty}df\exp\left[-\frac{f^2}{4c}+i\epsilon_{ij}\Sigma_{ij}
360: \left(f-\frac{g}{2}\left<E\right>\right)\right],
361: \end{equation}
362: where
363:
364: \begin{equation}
365: \label{c}
366: c\equiv\frac{g^2}{8}\left(\left<E^2\right>-\left<E\right>^2\right).
367: \end{equation}
368: The last formula in~(\ref{A}) means that, instead
369: of the constant field
370: $\left<E\right>$, we are now dealing with a shifted (but still space-independent)
371: one $\bar E\equiv\left<E\right>-\frac{2f}{g}$,
372: where the field $f$ should be eventually averaged over. The respective pair-production rate reads
373:
374: \begin{equation}
375: \label{neww}
376: w=\frac{N_cN_f}{2\sqrt{\pi c}}
377: \int\limits_{-\infty}^{+\infty}df{\rm e}^{-\frac{f^2}{4c}}\cdot
378: \frac{(g\bar E)^2}{(2\pi)^3}\sum\limits_{k=1}^{\infty}\frac{(-1)^{k+1}}{k^2}
379: \exp\left(-\frac{\pi k m^2}{g\bar E}\right).
380: \end{equation}
381: The parameter of the cumulant expansion, i.e. the ratio of the absolute value of the
382: second cumulant to that of the first one, is
383:
384: \begin{equation}
385: \label{par}
386: \frac{g}{2}\cdot 2\pi R_p^2\cdot \frac{\left<E^2\right>-\left<E\right>^2}{\left<E\right>}=g\left<E\right>
387: \cdot \pi R_p^2d.
388: \end{equation}
389: Here,
390:
391: \begin{equation}
392: \label{d}
393: d\equiv \frac{\left<E^2\right>-\left<E\right>^2}{\left<E\right>^2}
394: \end{equation}
395: is a measure of dispersion of the field $E(r)$. It is calculated in Appendix~A and
396: reads $d=\frac{\pi}{2}-1$. The condition of convergence of the cumulant expansion, i.e. the demand
397: that the parameter~(\ref{par}) is smaller than unity, then yields yet another upper boundary on $L$:
398:
399: \begin{equation}
400: \label{Bound}
401: L<\frac{4}{\pi d}\frac{\sigma}{m^2}\simeq 2.23\frac{\sigma}{m^2},
402: \end{equation}
403: that is, however, weaker than
404: the condition~(\ref{L}). As well as eq.~(\ref{L}), this new condition on $L$ becomes softer at $T\to T_c$, since
405: $\frac{\sigma}{m^2}=\frac{\sigma_0}{m_\pi^2}t^{-1.11}\to\infty$. In another words, at a fixed $L$,
406: obeying condition~(\ref{Bound}), cumulant expansion
407: converges better at $T\to T_c$, since its parameter vanishes in this limit as ${\cal O}(t^{1.11})$.
408:
409: To perform the average over $f$ in eq.~(\ref{neww}), note that the dispersion (i.e. the width) of
410: the Gaussian distribution, ${\rm e}^{-ax^2}$, is $\frac{1}{\sqrt{2a}}$. Therefore, characteristic $f$'s
411: obey the estimate $|f|\le\sqrt{2c}=\frac{g}{2}\sqrt{\left<E^2\right>-\left<E\right>^2}$, and
412: $\frac{2|f|/g}{\left<E\right>}\le\sqrt{d}\simeq\sqrt{0.57}
413: \simeq 0.76$. Therefore, although with the accuracy of only 76\% (cf. Section~5, where this problem will be avoided),
414: we can approximately write
415:
416: \begin{equation}
417: \label{barE}
418: \frac{1}{\bar E}\simeq\frac{1}{\left<E\right>}\left(1+\frac{2f}{g\left<E\right>}\right).
419: \end{equation}
420: When this expansion is substituted into eq.~(\ref{neww}), the $f$-integration can already be performed and yields
421: [cf. eq.~(\ref{w})]
422:
423:
424: \begin{equation}
425: \label{Wa}
426: w\simeq N_cN_f\frac{(g\left<E\right>)^2}{(2\pi)^3}\sum\limits_{k=1}^{\infty}\frac{(-1)^{k+1}}{k^2}
427: \left[1+\frac{2\pi m^2k}{g\left<E\right>}d\right]
428: \exp\left[-\frac{\pi m^2k}{g\left<E\right>}\left(1-\frac{\pi m^2k}{2g\left<E\right>}d\right)\right].
429: \end{equation}
430: Approximating again the whole sum by the first term only, we see that the obtained correction is small, provided
431: $\frac{\pi m^2}{g\left<E\right>}d\ll 1$.
432: This is precisely the condition of convergence of the cumulant expansion, eq.~(\ref{Bound}), with ``$<$''
433: replaced by ``$\ll$''. In particular,
434: the correction produced by the second cumulant becomes vanishingly smaller than
435: the leading term at $T\to T_c$. We also see that
436: the obtained correction increases $w$, thus diminishing $\bar r$. Due to condition~(\ref{L}),
437: $\bar r$ diminishes only by a factor of the order of unity.
438:
439:
440:
441:
442:
443:
444:
445: \section{Accounting for quantum effects by the Feynman variational method}
446: In this Section, we will evaluate the leading quantum correction to eq.~(\ref{Wa}). It can be obtained upon
447: the small-$T$ analysis of the path integral over ${\bf x}_\perp(\tau)$ in eq.~(\ref{1}) with the approximation~(\ref{smallT})
448: adopted. Such an integral can naturally be evaluated using the Feynman variational method~\cite{2}.
449: In 2d-case and in our notations, it looks as follows.
450: We need to evaluate the path integral
451:
452: $$
453: {\cal Z}(T)\equiv
454: \int d^2x(0)\int\limits_{{\bf x}(0)={\bf x}(T)}^{}{\cal D}{\bf x}(\tau){\rm e}^{-{\cal S}},$$
455: where ${\cal S}\equiv\int\limits_{0}^{T}d\tau
456: \left(\frac{\dot{\bf x}^2}{4}+U({\bf x})\right)$,
457: $U({\bf x})\equiv\frac{ig}{2T}E({\bf x})\epsilon_{ij}\Sigma_{ij}$. The
458: classical expression for this integral,
459: $\frac{1}{4\pi T}\int d^2x{\rm e}^{-TU({\bf x})}$, given by eq.~(\ref{class}), corresponds to $T$'s, which are
460: so small that the trajectory does not deviate from its initial point ${\bf x}(0)$. Let us further
461: introduce the coordinate describing the position of the trajectory ${\bf x}_0=\frac1T\int\limits_{0}^{T}d\tau{\bf x}(\tau)$,
462: the trial action ${\cal S}_0=\int\limits_{0}^{T}d\tau\frac{\dot{\bf x}^2}{4}+TW({\bf x}_0)$, and the respective trial partition function
463:
464: $$
465: {\cal Z}_0(T)=\int d^2x_0\int\limits_{{\rm fixed}{\,}{\bf x}_0}^{}{\cal D}{\bf x}{\rm e}^{-{\cal S}_0}=$$
466:
467: \begin{equation}
468: \label{Z0}
469: =\int d^2x_0\int\limits_{{\bf y}(0)={\bf y}(T)=0}^{}{\cal D}{\bf y}{\rm e}^{-\int\limits_{0}^{T}d\tau
470: \frac{\dot{\bf y}^2}{4}-TW({\bf x}_0)}=\frac{1}{4\pi T}\int d^2x{\rm e}^{-TW({\bf x})},
471: \end{equation}
472: where ${\bf y}(\tau)={\bf x}(\tau)-{\bf x}(0)$.
473: Here $W$ is a trial function, which
474: should be determined upon the minimization of the expression $F_0+\frac1T\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}$ (which approximates
475: the true free energy $F$ of the system from the above, $F\le F_0+\frac1T\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}$), where
476: $F=-\frac1T\ln{\cal Z}(T)$, $F_0=-\frac1T\ln{\cal Z}_0(T)$. It is further possible to demonstrate that the
477: averaged difference of actions can be written as
478:
479: $$\frac1T\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}=\left<U({\bf x}(0))\right>_{{\cal S}_0}-\left<W({\bf x}_0)\right>_{{\cal S}_0}.$$
480: For the first of the two averages here one has $\left<U({\bf x}(0))\right>_{{\cal S}_0}=\int d^2y{\cal K}({\bf y})
481: {\rm e}^{-TW({\bf y})}$, where
482:
483: $$
484: {\cal K}({\bf y})\equiv\int\limits_{{\bf x}(0)={\bf x}(T)}^{}{\cal D}{\bf x}U({\bf x}(0))\exp\left(-\int\limits_{0}^{T}d\tau
485: \frac{\dot{\bf x}^2}{4}\right)\delta({\bf y}-{\bf x}_0)=$$
486:
487: $$
488: =\int\frac{d^2q}{(2\pi)^2}\tilde U({\bf q})\int\frac{d^2k}{(2\pi)^2}{\rm e}^{-i{\bf k}{\bf y}}
489: \int\limits_{{\bf x}(0)={\bf x}(T)}^{}{\cal D}{\bf x}\exp\left[-\int\limits_{0}^{T}d\tau
490: \frac{\dot{\bf x}^2}{4}+\frac{i}{T}{\bf k}\int\limits_{0}^{T}d\tau{\bf x}+i{\bf q}{\bf x}(0)\right],$$
491: and $\tilde U({\bf q})\equiv\int d^2x{\rm e}^{-i{\bf q}{\bf x}}U({\bf x})$ is the Fourier image of $U({\bf x})$.
492: Using the formula
493:
494: $$\int\limits_{{\bf x}(0)={\bf x}(T)}^{}{\cal D}{\bf x}\exp\left[-\int\limits_{0}^{T}d\tau\left(
495: \frac{\dot{\bf x}^2}{4}+i{\bf f}{\bf x}\right)\right]=$$
496:
497: $$=\frac{\pi}{T}\delta\left(\int\limits_{0}^{T}
498: d\tau{\bf f}\right)\exp\left[\frac12\int\limits_{0}^{T}d\tau\int\limits_{0}^{T}d\tau'|\tau-\tau'|
499: {\bf f}(\tau){\bf f}(\tau')+\frac{1}{T}\sum\limits_{\alpha=1}^{2}\left(\int\limits_{0}^{T}d\tau\tau f_\alpha
500: \right)^2\right]$$
501: with ${\bf f}=\frac{{\bf k}}{T}+{\bf q}\delta(\tau)$, one obtains
502: ${\cal K}({\bf y})=\frac{3}{(2\pi T)^2}\int d^2xU({\bf x}){\rm e}^{-\frac{3}{T}({\bf x}-{\bf y})^2}$.
503: Thus, according to eq.~(\ref{Z0}),
504:
505: $$\frac1T\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}=\frac{\int d^2y{\rm e}^{-TW({\bf y})}\left[4\pi T{\cal K}({\bf y})-W({\bf y})
506: \right]}{\int d^2y{\rm e}^{-TW({\bf y})}}.$$
507: Note that the potential smeared over the Gaussian distribution,
508: $4\pi T{\cal K}({\bf y})$, goes over to the unsmeared potential $U({\bf y})$ in the limit $T\to 0$, as it should be.
509:
510: The variational equation
511: $\delta\left[F_0+\frac1T\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}\right]=0$ then yields $W({\bf y})=4\pi T{\cal K}({\bf y})$
512: as the best choice of $W({\bf y})$~\footnote{With this choice of $W({\bf y})$, $\left<{\cal S}-{\cal S}_0\right>_{{\cal S}_0}=0$.}.
513: Accordingly, the new expression, which accounts for quantum effects,
514:
515: \begin{equation}
516: \label{z0}
517: {\cal Z}_0(T)=\frac{1}{4\pi T}\int d^2y{\rm e}^{-4\pi T{\cal K}({\bf y})}=
518: \frac{1}{4\pi T}\int d^2y\exp\left[-\frac{3}{\pi}\int d^2xU({\bf x}){\rm e}^{-\frac{3}{T}({\bf x}-{\bf y})^2}\right]
519: \end{equation}
520: is a better approximation to ${\cal Z}(T)$ than its purely classical counterpart
521: $\frac{1}{4\pi T}\int d^2x{\rm e}^{-TU({\bf x})}$
522: (recoverable in the limit $T\to 0$), which was used before.
523:
524: The obtained result~(\ref{z0}) prescribes to replace $E({\bf x}_\perp)$ in eq.~(\ref{0}) by
525:
526: $${\cal E}({\bf x}_\perp)\equiv\frac{3}{\pi T}\int d^2yE({\bf y})
527: {\rm e}^{-\frac{3}{T}({\bf y}-{\bf x}_\perp)^2}.$$
528: In particular,
529: $\left<{\cal E}\right>=\left<E\right>$, i.e., at the level of the
530: first cumulant, the variational method yields the same result as the clasical approximation.
531: We further obtain
532:
533: $$\left<{\cal E}^2\right>=\frac{3}{2\pi TS}
534: \int d^2xd^2y{\rm e}^{-\frac{3}{2T}({\bf x}-{\bf y})^2}
535: E({\bf x})E({\bf y}).$$
536: This expression replaces $\left<E^2\right>$ in eq.~(\ref{A});
537: in particular, $\left<{\cal E}^2\right>\stackrel{T\to 0}{\longrightarrow}\left<E^2\right>$.
538: Thus, within the variational approach, the constant~(\ref{c}) becomes replaced by
539: $c_{\rm var}\equiv\frac{g^2}{8}\left(\left<{\cal E}^2\right>-\left<E\right>^2\right)$.
540: An apparent difference of
541: $c_{\rm var}$ from $c$ is that the former is $T$-dependent, whereas the latter is not~\footnote{In particular,
542: this means that the $f$-integral in eq.~(\ref{A}) should now be considered only inside the $T$-integral (and not vice versa).}.
543: The one-loop effective action reads
544:
545: $$\Gamma[A_i]\simeq S\frac{N_cN_f}{32\pi^{5/2}}\int\limits_{0}^{\infty}\frac{dT}{T^2}{\rm e}^{-m^2T}
546: \frac{1}{\sqrt{c_{\rm var}}}\int\limits_{-\infty}^{+\infty}df{\rm e}^{-\frac{f^2}{4c_{\rm var}}}
547: \frac{g\bar E}{\sin(g\bar E T)},$$
548: where ``$\simeq$'' means ``in the bilocal approximation to the cumulant expansion'',
549: and again $\bar E\equiv\left<E\right>-\frac{2f}{g}$.
550:
551: Let us further calculate the integral entering $c_{\rm var}$
552:
553: \begin{equation}
554: \label{intl}
555: \int d^2xd^2y{\rm e}^{-\frac{3}{2T}({\bf x}-{\bf y})^2}K_0(m_V|{\bf x}|)K_0(m_V|{\bf y}|)=
556: \frac{1}{m_V^4}\int d^2zd^2z'{\rm e}^{-\frac{3}{2Tm_V^2}({\bf z}-{\bf z}')^2}K_0(z)K_0(z'),
557: \end{equation}
558: where $z\equiv |{\bf z}|$, $z'\equiv |{\bf z}'|$.
559: This can be done approximately, by using the fact that $T<\frac{1}{m^2}\ll\frac{1}{m_V^2}$, i.e.
560: $\frac{1}{Tm_V^2}\gg 1$, that allows us to Taylor expand $K_0(z')$ around ${\bf z}$. This calculation, whose details are given in Appendix~A,
561: leads to the following result:
562:
563: \begin{equation}
564: \label{cvar}
565: c_{\rm var}=c+\frac{\sigma^3}{6g^2L^3}\frac{y}{g\bar E},
566: \end{equation}
567: where the variable $y\equiv g\bar ET$ acquires the values $\pi k$ when one is taking ${\rm Im}{\,}\Gamma[A_i]$.
568: We can further use the approximation~(\ref{barE}) with the term $\frac{2f}{g\left<E\right>}$ neglected, since
569: it would otherwise lead to the excess of accuracy, making the $f$-integral non-Gaussian. Denoting this approximation by
570: ``$\simeq$'' and substituting the above-mentioned values of $y$, we obtain the following modification of eq.~(\ref{c}):
571:
572: $$c_{\rm var}\simeq\frac{g^2}{8}\left[\left(1+\frac{\pi k}{48g^2}\right)\left<E^2\right>-\left<E\right>^2\right].$$
573: This finally results in the following change of the dispersion parameter (\ref{d}), entering eq.~(\ref{Wa}),
574: that makes this parameter $k$-dependent:
575:
576: $$d\to d_k\equiv\frac{\left(1+\frac{\pi k}{48g^2}\right)\left<E^2\right>-\left<E\right>^2}{\left<E\right>^2}.$$
577: This formula describes the quantum correction to eq.~(\ref{Wa}). We, however, see that this correction is
578: small. Indeed, due to the main exponential factor in eq.~(\ref{Wa}), we have for relevant $k$'s:
579:
580: $$
581: \frac{\pi k}{48g^2}<\frac{\pi}{48g^2}\cdot \frac{g\left<E\right>}{\pi m^2}=\frac{\sigma}{12L(gm)^2}\ll\frac{1}{48\pi},$$
582: where the last inequality stems from the condition $m\gg m_V$ [expressed by the left inequality of~(\ref{ineq})].
583:
584:
585:
586:
587:
588:
589: \section{Accounting for the space dependence of $E(r)$ without the cumulant expansion}
590: In this Section, we will evaluate $w$ in an alternative way, namely completely
591: without the use of cumulant expansion in eq.~(\ref{0}). The necessity of this calculation
592: is, firstly, because cumulant expansion has been shown to converge only provided $L$ is bounded from above
593: according to the condition~(\ref{Bound}). The new method enables one to replace this constraint
594: by some other, weaker, one, and to relax the constraint~(\ref{L}). Secondly, the approximation used
595: to arrive at eq.~(\ref{Wa}), which was holding with only 76\% accuracy, will now be not necessary anymore.
596:
597: The method of this Section is based on
598: averaging every term in the expansion of the exponent in eq.~(\ref{0})
599: over $d^2x_\perp$. In the London limit, such an average can be done analytically by making use of an explicit form of $E(r)$.
600: Namely, we have
601:
602: $$
603: \int
604: d^2x_\perp\exp\left[-\frac{ig}{2}\epsilon_{ij}\Sigma_{ij}E({\bf x}_\perp)\right]=
605: \sum\limits_{n=0}^{\infty}\frac{1}{n!}\left(-\frac{ig}{2}\epsilon_{ij}\Sigma_{ij}\right)^n
606: \left(\frac{m_V^2}{g_m}\right)^n\int d^2x_\perp(K_0(m_Vr))^n.$$
607: The last integral approximately equals $\frac{\pi}{m_V^2}2^{2-n}n!$, that
608: yields
609:
610: $$
611: \frac{4\pi}{m_V^2}\sum\limits_{n=0}^{\infty}\left(-\frac{ig}{4}\frac{m_V^2}{g_m}
612: \epsilon_{ij}\Sigma_{ij}\right)^n=\frac{4\pi/m_V^2}{1+\frac{ig^2m_V^2}{8\pi}\epsilon_{ij}\Sigma_{ij}}
613: =\frac{4\pi}{m_V^2}\int\limits_{0}^{\infty}dt{\rm
614: e}^{-t\left(1+\frac{ig^2m_V^2}{8\pi}\epsilon_{ij}\Sigma_{ij}\right)}.
615: $$
616: The series above converges at $\frac{g^2m_V^2}{8\pi}\cdot 2\pi R_p^2<1$, that
617: results in the following boundary on $L$ from above [cf. eq.~(\ref{Bound})]:
618:
619: \begin{equation}
620: \label{LLL}
621: L<\frac{16}{\pi}\frac{\sigma}{m^2}\simeq 5.09\frac{\sigma}{m^2}.
622: \end{equation}
623: We further obtain for eq.~(\ref{eff}):
624:
625: $$\Gamma[A_i]\simeq S\frac{N_cN_f}{\pi}\int\limits_{0}^{\infty}\frac{dT}{T^2}{\rm e}^{-m^2T}\int {\cal D}{\bf x}_\parallel
626: \int\limits_{0}^{\infty}dt{\rm
627: e}^{-t\left(1+\frac{ig^2m_V^2}{8\pi}\epsilon_{ij}\Sigma_{ij}\right)}=$$
628:
629: \begin{equation}
630: \label{newG}
631: =S\frac{N_cN_f}{4\pi^2}\int\limits_{0}^{\infty}\frac{dT}{T^2}{\rm e}^{-m^2T}
632: \int\limits_{0}^{\infty}dt{\rm e}^{-t}\frac{g\bar E}{\sin(g\bar E T)},
633: \end{equation}
634: where $\bar E\equiv\frac{tgm_V^2}{4\pi}$ is a new space-independent electric field. We see that
635: the constraint~(\ref{L}) is now removed. Indeed, the condition for the first pole in
636: ${\rm Im}{\,}\Gamma[A_i]$ to contribute, $\pi\le g\bar ET$, together with the inequality $T<\frac{1}{m^2}$
637: yield $\pi<\frac{g\bar E}{m^2}=\frac{t}{L}\frac{\sigma}{m^2}$.
638: This condition does not produce anymore a constraint on $L$, but
639: merely means that $t$'s obeying the inequality
640: $t>\pi L\frac{m^2}{\sigma}$ give a dominant contribution to ${\rm Im}{\,}\Gamma[A_i]$.
641:
642:
643: The pair-creation rate stemming from eq.~(\ref{newG}) takes the form
644:
645: $$
646: w=\frac{N_cN_f}{2\pi^3}\left(\frac{\sigma}{L}\right)^2\sum\limits_{k=1}^{\infty}
647: \frac{(-1)^{k+1}}{k^2}\int\limits_{0}^{\infty}dtt^2{\rm e}^{-t-\frac{\pi
648: Lm^2k}{\sigma t}}=$$
649:
650: $$=N_cN_f\frac{m^3}{\pi^{3/2}}\sqrt{\frac{\sigma}{L}}\sum\limits_{k=1}^{\infty}
651: \frac{(-1)^{k+1}}{\sqrt{k}}
652: K_3\left(2m\sqrt{
653: \frac{\pi Lk}{\sigma}}\right).$$
654: Clearly, only terms with $k\le\frac{\sigma}{4\pi Lm^2}$ are relevant in the
655: sum. At $T=0$,
656: substituting $\sigma=(440{\,}{\rm MeV})^2$ and a typical value $m=200{\,}{\rm MeV}$,
657: we obtain $k<\frac1L<1$. Therefore,
658: only the first term is relevant, that yields
659:
660: \begin{equation}
661: \label{modw}
662: w\simeq N_cN_f\frac{m^3}{\pi^{3/2}}\sqrt{\frac{\sigma}{L}}
663: K_3\left(2m\sqrt{
664: \frac{\pi L}{\sigma}}\right).
665: \end{equation}
666: According to the above-obtained constraint~(\ref{LLL}), the
667: argument of the MacDonald function in this formula is smaller than 8. For the values of $L$, at which this
668: argument is still larger than unity, i.e. $L>\frac{\sigma}{4\pi m^2}$,
669:
670: \begin{equation}
671: \label{Neww}
672: w\simeq N_cN_f\frac{m^{5/2}\sigma^{3/4}}{2\pi^{5/4}L^{3/4}}{\rm
673: e}^{-2m\sqrt{\frac{\pi L}{\sigma}}}.
674: \end{equation}
675: When $L$ obeys simultaneously inequality~(\ref{L}), the obtained expression can be compared to eq.~(\ref{ww}).
676: The parametric dependences of these two expression on $m$, $\sigma$, and $L$ are
677: apparently different from each other. The regime under
678: discussion, $2m\sqrt{\frac{\pi L}{\sigma}}> 1$, however, does not imply the exponential smallness of~(\ref{Neww}),
679: since, due to~(\ref{L}), $2m\sqrt{\frac{\pi L}{\sigma}}<4$. The string-breaking distance, corresponding to eq.~(\ref{Neww}),
680: reads
681:
682: \begin{equation}
683: \label{newdist}
684: \bar r=\frac{2\pi^{5/8}}{g\sqrt{N_cN_f}}\left(\frac{\sigma}{L}\right)^{1/8}\frac{1}{m^{5/4}}
685: {\rm e}^{m\sqrt{\frac{\pi L}{\sigma}}}.
686: \end{equation}
687: Its ratio to distance~(\ref{rr}) is given by the function
688: $\frac{2}{\pi^{1/4}x^{5/4}}{\rm e}^{x-\frac{x^2}{8}}$, where the variable $x\equiv m\sqrt{\frac{\pi L}{\sigma}}$
689: ranges between $\frac12$ and 2. This is a monotonically decreasing function, which therefore acquires its maximum
690: at $x=\frac12$ [where approximation~(\ref{Neww}) to eq.~(\ref{modw}) starts breaking down].
691: The value of the maximum, $\simeq 9.10$,
692: is, thus, an upper limit for the ratio of the two string-breaking distances. The minimal value of this
693: ratio, corresponding to $x=2$, is only $\simeq 2.84$.
694:
695: Instead, at $T\to T_c$,
696: $\frac{m}{\sqrt{\sigma}}=\frac{m_\pi}{\sqrt{\sigma_0}}t^{0.55}\to 0$, and
697: we have $w\to\frac{N_cN_f}{\pi^3}\left(\frac{\sigma}{L}\right)^2$, that differs from eq.~(\ref{tto0}) only by a factor 2.
698: The respective string-breaking distance is, therefore, larger only by a factor $\sqrt{2}$.
699:
700:
701:
702:
703:
704:
705:
706:
707:
708: \section{Considering pairs as holes in the confining pellicle}
709: In this Section, we will consider an alternative approach to the pair production, based on a combination of
710: the formulae on the metastable vacuum decay~\cite{5,6} with the stochastic vacuum model~\cite{7} (for a recent review
711: see~\cite{8}). The idea is to consider the produced pairs as holes in a 2d confining pellicle, which spans over the
712: contour of an external $q\bar q$-pair. Such a hole is a region where the pellicle is eaten up.
713: Therefore, a hole of a radius $R$ diminishes the action of the pellicle by $\sigma\cdot\pi R^2$,
714: but increases this action by $(m+\mu)\cdot 2\pi R$.
715: Here $\mu$ is some parameter of dimension [mass], which is
716: a nonperturbative part of a constant entering the
717: perimeter law of the small-sized Wilson loop of a produced pair.
718: The critical radius (i.e. such a radius, that all holes with $R<R_c$ collapse, while those
719: with $R>R_c$ expand and destroy the pellicle) stems therefore from extremization of the action
720: $S(R)\equiv(m+\mu)\cdot 2\pi R-\sigma\cdot\pi R^2$. This
721: critical radius and the action of a critical hole are $R_c=(m+\mu)/\sigma$ and $S(R_c)=\pi(m+\mu)^2/\sigma$.
722: Accordingly, in this approach,
723: the rate of the pair production is proportional to ${\rm e}^{-\pi(m+\mu)^2/\sigma}$. The proportionality
724: coefficient~\cite{6},
725: $\frac{\sigma}{2\pi}$, should furthermore be multiplied by the factor $N_f$~\footnote{The factor $N_c$ is now absent, since
726: the Wilson loop, describing a produced pair, is a colorless object. Instead, the number of Wilson loops, which
727: can potentially be created, is proportional to the number of different quark species, i.e. to $N_f$.}:
728:
729: \begin{equation}
730: \label{newW}
731: w=N_f\frac{\sigma}{2\pi}\exp\left[-\frac{\pi m^2}{\sigma}\left(
732: 1+\frac{\mu}{m}\right)^2\right].
733: \end{equation}
734: Next, to evaluate both $\mu$ and $\sigma$ within the
735: same model, it is natural to use the stochastic vacuum model. It yields the following expression for the Wilson loop
736: $\left<W(C)\right>_{\rm YM}\equiv\frac{1}{N_c}
737: \left<{\rm tr}{\,}{\cal P}\exp\left(ig\oint\limits_{C}^{}
738: dx_\mu A_\mu\right)\right>_{\rm YM}$:
739:
740: \begin{equation}
741: \label{Wloop}
742: \left<W(C)\right>_{\rm YM}\simeq
743: \frac{1}{N_c}{\,}{\rm tr}{\,}\exp\left[-\frac{1}{2!}\frac{g^2}{4}\int\limits_{\Sigma(C)}^{}d\sigma_{\mu\nu}(x)
744: \int\limits_{\Sigma(C)}^{}d\sigma_{\lambda\rho}(x')\left<F_{\mu\nu}(x)\Phi_{xx'}F_{\lambda\rho}(x')\Phi_{x'x}\right>_{\rm YM}
745: \right],
746: \end{equation}
747: where $\Sigma(C)$ is the surface encircled by the flat contour $C$. Next, $A_\mu\equiv A_\mu^aT^a$, where $T^a$'s
748: stand for the generators of the group SU($N_c$) in the fundamental representation,
749: $\left[T^a,T^b\right]=if^{abc}T^c$, ${\rm tr}~ T^aT^b=\frac12\delta^{ab}$; the average $\left<\ldots\right>_{\rm YM}$
750: is implied with respect to the Euclidean Yang-Mills action, $\frac14\int d^4x(F_{\mu\nu}^a)^2$, where
751: $F_{\mu\nu}^a=\partial_\mu A_\nu^a-\partial_\nu A_\mu^a+gf^{abc}A_\mu^b A_\nu^c$, $a=1, \ldots, N_c^2-1$;
752: $\Phi_{xx'}\equiv\frac{1}{N_c}{\,}{\cal P}{\,}\exp\left(ig\int\limits_{x'}^{x}dz_\mu A_\mu(z)\right)$ is a phase
753: factor along the straight line, which goes through $x'$ and $x$.
754: The symbol ``$\simeq$'' in eq.~(\ref{Wloop})
755: is implied in the sense of the bilocal approximation to the cumulant expansion. This approximation, supported
756: by the lattice data~\cite{8, 9, 10}, states that,
757: in the Yang-Mills theory, the
758: two-point irreducible gauge-invariant correlation function (cumulant) of $F_{\mu\nu}$'s dominates over all cumulants
759: of higher orders, which are therefore neglected. Finally,
760: the factor $1/2!$ in eq.~(\ref{Wloop})
761: is simply due to the cumulant expansion, whereas the factor
762: $1/4$ is due to the non-Abelian Stokes' theorem.
763:
764: The stochastic vacuum model suggests further the following parametrization of the two-point cumulant:
765:
766: $$\left<F_{\mu\nu}(x)\Phi_{xx'}F_{\lambda\rho}(x')\Phi_{x'x}\right>_{\rm YM}=\frac{\hat 1_{N_c\times N_c}}{N_c}{\cal N}
767: \Biggl\{(\delta_{\mu\lambda}\delta_{\nu\rho}-\delta_{\mu\rho}\delta_{\nu\lambda})D\left((x-x')^2\right)+$$
768:
769: $$
770: +\frac12\left[\partial_\mu^x\left((x-x')_\lambda\delta_{\nu\rho}-(x-x')_\rho\delta_{\nu\lambda}\right)+
771: \partial_\nu^x\left((x-x')_\rho\delta_{\mu\lambda}-(x-x')_\lambda\delta_{\mu\rho}\right)\right]\times$$
772:
773: \begin{equation}
774: \label{pa}
775: \times\left[D_1\left((x-x')^2\right)+\frac{N_cC_2}{{\cal N}\pi^2|x-x'|^4}\right]
776: \Biggr\}.
777: \end{equation}
778: Here $C_2\equiv\frac{N_c^2-1}{2N_c}$ is the quadratic Casimir operator of the fundamental representation, and
779: ${\cal N}$ is the normalization constant, which in 4d reads
780: ${\cal N}=\frac{\left<(F_{\mu\nu}^a)^2\right>_{\rm YM}}{24[D(0)+D_1(0)]}$~\footnote{Note that the one-gluon-exchange
781: contribution, represented by the $\frac{1}{|x-x'|^4}$-term, can alternatively be considered as a part of the
782: function $D_1$~\cite{8}. Here, we rather consider this perturbative contribution separately, so that $D_1(0)$ is a finite
783: quantity.}. Inserting this parametrization into
784: eq.~(\ref{Wloop}) we obtain (applying Abelian Stokes' theorem to the part containing derivatives):
785:
786: $$\left<W(C)\right>_{\rm YM}\simeq\exp\Biggl\{-\frac{g^2{\cal N}}{8N_c}\Biggl[2
787: \int\limits_{\Sigma(C)}^{}d\sigma_{\mu\nu}(x)
788: \int\limits_{\Sigma(C)}^{}d\sigma_{\mu\nu}(x')D\left((x-x')^2\right)+$$
789:
790: \begin{equation}
791: \label{G}
792: +\oint\limits_{C}^{}dx_\mu
793: \oint\limits_{C}^{}dx'_\mu\left[G\left((x-x')^2\right)+\frac{N_cC_2}{{\cal N}\pi^2}\frac{1}{(x-x')^2}\right]
794: \Biggr]\Biggr\},
795: \end{equation}
796: where $G(x^2)\equiv\int\limits_{x^2}^{\infty}dt D_1(t)$. The one-gluon-exchange contribution to the Wilson loop,
797: $\exp\left[-\frac{g^2C_2}{8\pi^2}\oint\limits_{C}^{}dx_\mu
798: \oint\limits_{C}^{}dx'_\mu\frac{1}{(x-x')^2}\right]\simeq\exp\left(-\frac{g^2C_2}{8\pi a}\right)$ is known \cite{12} to yield the
799: renormalization of mass of a produced pair. Here $a$ stands for an inverse UV cutoff; $a\ll T_g$, and $T_g\simeq 1{\,}{\rm GeV}^{-1}$
800: is the correlation length of the vacuum~\cite{7, 8, 9, 10}.
801: The asymptotics of the Wilson loop at $\sqrt{|\Sigma(C)|}\gg T_g$, where $|\Sigma(C)|$ is the area of $\Sigma(C)$,
802: is $\left<W(C)\right>_{\rm YM}\simeq{\rm e}^{-\sigma|\Sigma(C)|}$. This asymptotics is obeyed by the Wilson loop
803: of an external $q\bar q$-pair. Instead, at $\sqrt{|\Sigma(C)|}={\cal O}(T_g)$, i.e. for the Wilson loop of a produced pair,
804: $\left<W(C)\right>_{\rm YM}\simeq
805: {\rm e}^{-(m+\mu) L(C)}$, where $L(C)$ is the length of $C$~\footnote{Below, we will see that the
806: typical size of the Wilson loop of a produced pair is indeed ${\cal O}(T_g)$.}.
807: As follows from eq.~(\ref{G}), the parametrization~(\ref{pa})
808: is chosen in such a way that the function $D$ yields the area law,
809: while the term with the function $D_1$ contributes to the perimeter law. Namely,
810: the string tension reads~\cite{11}
811: $\sigma=\frac{g^2{\cal N}}{2N_c}\int d^2x D({\bf x}^2)$,
812: while for the perimeter constant $\mu$ we obtain in a way similar to the Coulomb \cite{12}
813: interaction between points lying on $C$:
814: $\mu=\frac{g^2{\cal N}}{8N_c}\int\limits_{0}^{\infty}d\xi G(\xi^2)$. A derivation of this formula is presented in Appendix~A.
815:
816: In what follows, we will adopt the exponential
817: parametrization of the functions $D$ and $D_1$~\cite{8, 9, 10}
818: $D(x^2)=D(0){\rm e}^{-|x|/T_g}$, $D_1(x^2)=D_1(0){\rm e}^{-|x|/T_g}$.
819: It yields the following values of the string tension and the perimeter constant:
820: $\sigma=\frac{\pi g^2{\cal N}}{N_c}T_g^2D(0)$,
821: $\mu=\frac{g^2{\cal N}}{2N_c}T_g^3D_1(0)$. Introducing the
822: parameter $\gamma\equiv D_1(0)/D(0)$, whose lattice value in full QCD is $0.13\pm 0.08$ (see e.g.~\cite{14}),
823: we can rewrite the obtained results as
824:
825: \begin{equation}
826: \label{sigma}
827: \sigma=\frac{\pi}{24(1+\gamma)N_c}g^2\left<(F_{\mu\nu}^a)^2\right>T_g^2,~~
828: \mu=\frac{\gamma}{48(1+\gamma)N_c}g^2\left<(F_{\mu\nu}^a)^2\right>T_g^3.
829: \end{equation}
830: For the critical radius of a hole we have $R_c=\frac{m+\mu}{\sigma}=\frac{m}{\sigma}+
831: \frac{\gamma}{2\pi}T_g$. Substituting again $\sigma=(440{\,}{\rm MeV})^2$, $m=200{\,}{\rm MeV}$, and
832: $T_g^{-1}=1{\,}{\rm GeV}$,
833: we get for the worst case of the
834: above-quoted lattice value of $\gamma$, $\gamma=0.21$, $\frac{R_c}{T_g}\simeq 1.07$~\footnote{Notice that
835: $T_g$ itself is smaller (by a factor of the order of 5)
836: than a typical size of the Wilson loop of an external $q\bar q$-pair, at which the
837: onset of string-breaking is normally expected.}.
838: This ratio is certainly ${\cal O}(1)$,
839: that justifies the use of the perimeter asymptotics for the Wilson loop of a produced pair.
840:
841: Comparing the pair production rate, eq.~(\ref{newW}), with eq.~(\ref{ww}), we see that the parameter
842: $\left(1+\frac{\mu}{m}\right)^2$ replaces the parameter $L/4$ of that equation. At
843: $\sigma=(440{\,}{\rm MeV})^2$, $m=200{\,}{\rm MeV}$, we have~\footnote{The correction $\frac{\mu}{m}$ is very small:
844: $\frac{\mu}{m}=\frac{\gamma}{2\pi}\frac{T_g\sigma}{m}\simeq 0.02$.}
845: $\left(1+\frac{\mu}{m}\right)^2\simeq 1.08$, whereas,
846: according to eq.~(\ref{L}), $\frac{L}{4}<1.55$, that is of the same order of magnitude.
847: Finally, the new string-breaking distance, at which the potential
848: $V(r)=\sigma r{\rm e}^{-\pi r^2\cdot w}$ acquires its maximum, is
849:
850: \begin{equation}
851: \label{NEWr}
852: \bar r=\frac{1}{\sqrt{2\pi w}}=\frac{1}{\sqrt{N_f\sigma}}\exp\left[\frac{\pi m^2}{\sigma}\left(
853: 1+\frac{\mu}{m}\right)^2\right].
854: \end{equation}
855: The argument of the exponent here approximately equals $0.70$, that is quite similar to $0.5$
856: we had as an upper boundary in case of eq.~(\ref{rr}).
857: However, contrary to that equation, eq.~(\ref{NEWr}) does not contain the factor $\sqrt{L}$ in the preexponent.
858:
859: Notice also that,
860: due to the fact that $\sigma={\cal O}(N_c^0)$, the string-breaking distance~(\ref{rr}) is ${\cal O}(N_c^0)$ too~\cite{1}.
861: For the same reason, as can be seen from the first of eqs.~(\ref{sigma}),
862: $T_g$ is ${\cal O}(N_c^0)$ as well.
863: Therefore, according to the second equation of~(\ref{sigma}), $\mu$ is also ${\cal O}(N_c^0)$.
864: This means that, as well as eq.~(\ref{rr}), the new string-breaking distance~(\ref{NEWr}) is ${\cal O}(N_c^0)$.
865:
866:
867:
868:
869: \section{Summary}
870: In this paper, we have considered two approaches to the problem of string breaking in QCD: one of these is based
871: on the dual superconductor model of confinement, and the other one -- on the stochastic vacuum model.
872: In the first approach, which was proposed already in ref.~\cite{1}, the pair-production mechanism is due to the
873: field of the chromoelectric flux tube (dual Abrikosov-Nielsen-Olesen string in the London limit of the
874: dual Abelian Higgs model). In the second approach,
875: pairs are considered as holes in a pellicle, which confines a test $q\bar q$-pair.
876:
877: Within the first approach,
878: we have demonstrated that the result of ref.~\cite{1}, based on the Schwinger formula, accounts only
879: for the first term of the cumulant expansion in the average~(\ref{A}). In Section~3, we have found the temperature
880: dependences of the string-breaking distance at temperatures close to the critical one and at low temperatures,
881: smaller than ${\cal O}(f_\pi)$. In both cases, the string-breaking distance increases with the increase of the
882: temperature. We have also noticed that, in the case of spin-$\frac12$ quarks, antiperiodic boundary conditions for fermions
883: produce only corrections which fall off as the tail of the Gaussian distribution,
884: as long as $T<m_\pi/\sqrt{2}$. As a by-product, we have found in Appendix~B the temperature dependence of the string tension, which
885: reproduces correctly the zero-temperature value.
886: In Section~4, we have calculated (for scalar quarks) a
887: correction, generated by the second cumulant in the expansion~(\ref{A}),
888: i.e. by the dispersion of the chromoelectric field in the direction
889: transverse to the string. This effect slightly
890: diminishes the string-breaking distance. Using the Feynman variational method,
891: we have then derived in Section~4 the leading quantum
892: correction to this effect, produced by the deviation of the trajectory of a pair from the classical one.
893: This effect further diminishes the string-breaking distance.
894:
895: The effects of dispersion of the chromoelectric field are small
896: as long as the cumulant expansion is convergent, that is
897: the case when the logarithm of the Landau-Ginzburg parameter is bounded from above
898: according to~(\ref{Bound}). This constraint is, however, always obeyed as long as at least the first pole in the
899: Schwinger formula gives its contribution, that leads to an even more severe constraint~(\ref{L}). Although both
900: constraints have been shown to relax at $T\to T_c$, they do exist at $T=0$. This fact necessitates to perform the
901: average~(\ref{A}) without the use of the cumulant expansion at all, that has been done in Section~6. As a result, the
902: upper boundary on the logarithm of the Landau-Ginzburg parameter increases by a factor 4 [cf. eq.~(\ref{LLL})].
903: Furthermore, a novel formula~(\ref{modw}) for the rate of the pair production has been derived. At $T=0$,
904: some range of the Landau-Ginzburg parameter has been found, in which the string-breaking distance is
905: larger than the one we had with the use of
906: the bilocal approximation to the cumulant expansion in a factor varying
907: between $2.84$ and $9.10$~\footnote{This result indicates that, at least at these values of the Landau-Ginzburg parameter,
908: the effect produced by cumulants higher than the quadratic one is opposite and stronger than the result produced by
909: the quadratic cumulant alone.}.
910: Instead, at $T\to T_c$, this factor is smaller, namely it equals $\sqrt{2}$.
911: Notice that such an analytic average of the exponent~(\ref{A}) without the
912: use of the cumulant expansion was only possible due to the explicit form of the Abrikosov-Nielsen-Olesen solution in the
913: London limit. Apparently, studies away from this limit will require a numerical analysis.
914:
915: Within the approach, which treats pairs as holes in the confining pellicle, the new quantity on which
916: the string-breaking distance is dependent is the constant entering the perimeter law of the Wilson loop
917: of a produced pair.
918: For typical values of the hadronic mass, string tension, and the vacuum correlation length, this dependence is,
919: however, very weak. As for the dependence of the new expression for the string-breaking distance
920: on the mass of a produced pair and on the string tension, it is the same as in the above-discussed
921: case of the approach based on the
922: Schwinger formula, where only the second cumulant is taken into account. Finally, the results
923: for the string-breaking distance, obtained within all the above-mentioned approaches, are ${\cal O}(N_c^0)$
924: in the large-$N_c$ limit.
925:
926:
927: \section*{Acknowledgments}
928: One of us (D.A.) acknowledges the Alexander~von~Humboldt foundation for the financial support.
929: He would also like to thank the staffs of the Physics Departments of the University of Pisa and of the
930: Humboldt University of Berlin for the cordial hospitality.
931:
932: \appendix
933: \section{Some technical details}
934: Let us first present evaluation of the dispersion parameter~(\ref{d}). Integrating both sides of the equality
935: $\frac{1}{2\pi}K_0(mr)=\int\frac{d^2p}{(2\pi)^2}\frac{{\rm e}^{i{\bf p}{\bf x}_\perp}}{p^2+m^2}$ over $d^2r$
936: we trivially have $\int d^2rK_0(mr)=\int\frac{d^2p}{p^2+m^2}\delta({\bf p})=\frac{1}{m^2}$. In the same way
937:
938: \begin{equation}
939: \label{a1}
940: \int d^2rK_0^2(mr)=\int d^2r\int\frac{d^2pd^2q}{(2\pi)^2}
941: \frac{{\rm e}^{i({\bf p}+{\bf q}){\bf x}_\perp}}{(p^2+m^2)(q^2+m^2)}=\frac{1}{2\pi}\int\frac{d^2p}{(p^2+m^2)^2}=
942: \frac{1}{2m^2}.
943: \end{equation}
944: The parameter~(\ref{d}) then reads
945:
946: $$
947: d=\frac{\int d^2rK_0^2(m_Vr)-\frac{m_V^2}{\pi}\left[\int d^2rK_0(m_Vr)
948: \right]^2}{\frac{m_V^2}{\pi}\left[\int d^2rK_0(m_Vr)\right]^2}=\frac{\frac{1}{2m_V^2}-
949: \frac{1}{\pi m_V^2}}{\frac{1}{\pi m_V^2}}=\frac{\pi}{2}-1.$$
950:
951: Next, we will discuss some details of derivation of $c_{\rm var}$. Taylor expanding $K_0(z')$
952: in eq.~(\ref{intl}) to the second order we have
953:
954: $$\frac{1}{m_V^4}\int d^2zd^2z'{\rm e}^{-\frac{3}{2Tm_V^2}({\bf z}-{\bf z}')^2}K_0(z)K_0(z')\simeq$$
955:
956: $$
957: \simeq\frac{2\pi T}{3m_V^2}\int d^2z K_0^2(z)-\frac{1}{m_V^4}\int d^2z\frac{z_\mu}{z}K_1(z)
958: \int d^2z'(z'-z)_\mu{\rm e}^{-\frac{3}{2Tm_V^2}({\bf z}-{\bf z}')^2}+$$
959:
960: $$+\frac{1}{2m_V^4}\int d^2z\left[
961: \frac{z_\mu z_\nu-z^2\delta_{\mu\nu}}{z^3}K_1(z)+\frac{z_\mu z_\nu}{2z^2}(K_0(z)+
962: K_2(z))\right]K_0(z)\times$$
963:
964: $$\times\int d^2z'(z'-z)_\mu(z'-z)_\nu{\rm e}^{-\frac{3}{2Tm_V^2}({\bf z}-{\bf z}')^2}.$$
965: The second term on the r.h.s. of this equation apparently vanishes, while the third ones reads
966: $\frac{\pi T^2}{9}\int d^2zK_0\left[\frac12(K_0+K_2)-\frac{1}{z}K_1\right]$. Using the definition
967: $c_{\rm var}\equiv\frac{g^2}{8}\left(\left<{\cal E}^2\right>-\left<E\right>^2\right)$, we then arrive at
968: the following intermediate result:
969:
970: $$
971: c_{\rm var}=c+\frac{\pi\sigma^3T}{3g^2L^3}\int\limits_{0}^{\infty}dz K_0\left[
972: z(K_0+K_2)-2K_1\right].$$
973: Here, the addendum $c$, eq.~(\ref{c}), is apparently produced by the term with no derivatives, while the other addendum is produced
974: by the second-derivative term of the Taylor expansion. Finally, the integral over $z$ can be calculated
975: exactly. It reads
976:
977: $$\frac{1}{2\pi}\int d^2z K_0^2+\int\limits_{0}^{\infty}dzK_0\left[z\left(K_0+\frac2z K_1\right)-2K_1\right]=
978: \frac{1}{\pi}\int d^2z K_0^2=\frac{1}{2\pi},$$
979: where eq.~(\ref{a1}) on the last step has been used. This yields eq.~(\ref{cvar}).
980:
981:
982: We will finally present a proof of the formula
983:
984: \begin{equation}
985: \label{length}
986: \oint\limits_{C}^{}dx_\mu\oint\limits_{C}^{}dx_\mu'
987: G\left((x-x')^2\right)\simeq L\cdot\int\limits_{0}^{\infty}d\xi G(\xi^2),
988: \end{equation}
989: where $L=\int\limits_{0}^{1}ds
990: |\dot x_\mu(s)|$ is the length of the contour $C$. Since the function $G(x^2)$ is rapidly decreasing
991: [e.g. $G(x^2)=2D_1(0)T_g(|x|+T_g){\rm e}^{-|x|/T_g}$ for the adopted Ansatz $D_1(x^2)=D_1(0){\rm e}^{-|x|/T_g}$],
992: we can Taylor expand $x_\mu'\equiv x_\mu(s+t)$ as $x_\mu(s+t)\simeq x_\mu(s)+t\dot x_\mu(s)$. In the expression
993: under study,
994: $\int\limits_{0}^{1}ds \int\limits_{0}^{1}dt\dot x_\mu(s)\dot x_\mu(s+t) G(t^2\dot x_\mu^2(s))$,
995: we therefore have $\dot x_\mu(s)\dot x_\mu(s+t)\simeq \dot x_\mu^2(s)+t\dot x_\mu(s)\ddot x_\mu(s)=
996: \dot x_\mu^2(s)$, where, at the last step, the proper-length parametrization $\dot x_\mu^2={\rm const}$ has been fixed.
997: Introducing instead of $t$ the new integration variable $\xi=t|\dot x_\mu|$, we have
998: for the expression under study
999: $\int\limits_{0}^{1}ds|\dot x_\mu(s)|\int\limits_{0}^{|\dot x_\mu|}d\xi G(\xi^2)$.
1000: Finally, due to the rapid decrease of $G(\xi^2)$, the upper integration limit
1001: in the last integral can be replaced by infinity, that completes the proof of eq.~(\ref{length}).
1002:
1003:
1004:
1005:
1006: \section{Temperature dependence of the string tension}
1007: To find the dependence $\sigma(T)$, notice that the string tension can be derived from the
1008: string representation of the
1009: partition function~\cite{strrepr}. An essential result of
1010: this representation is the following string effective action:
1011:
1012: $$2(\pi v)^2\int d\sigma_{\mu\nu}(x)\int d\sigma_{\mu\nu}(x')D_{m_V}(x-x'),$$
1013: where $D_m(x)\equiv mK_1(m|x|)/(4\pi^2|x|)$
1014: is the Yukawa propagator. The zero-temperature string tension then stems from this action according to the general (for this type
1015: of non-local string actions) formula~\cite{11}: $\sigma_0=(\pi v)^2\cdot\frac{4}{m_V^2}
1016: \int\limits_{m_V/m_H}^{}d^2z\frac{m_V^2K_1(|z|)}{4\pi^2|z|}$. This indeed equals the above-used value $2\pi v^2L$
1017: (following from the Landau-Ginzburg equations). At finite temperatures (smaller than the temperature of dimensional reduction),
1018: we rather have the following
1019: expression in terms of Matsubara frequencies:
1020:
1021: $$
1022: \sigma(T)=v^2\int\limits_{m_V/m_H}^{}d^2z\sum\limits_{n=-\infty}^{+\infty}
1023: \frac{K_1\left(\sqrt{z^2+(m_V\beta n)^2}\right)}{\sqrt{z^2+(m_V\beta n)^2}}.$$
1024: To perform the integration, it is useful to transform
1025: the sum as follows:
1026:
1027: $$\sum\limits_{n=-\infty}^{+\infty}
1028: \frac{K_1\left(\sqrt{z^2+(m_V\beta n)^2}\right)}{\sqrt{z^2+(m_V\beta n)^2}}=\sum\limits_{n=-\infty}^{+\infty}
1029: \int\limits_{0}^{\infty}dt\exp\left[-\frac{1}{4t}-t\left(z^2+(m_V\beta n)^2\right)\right]=$$
1030:
1031: $$=\frac{T\sqrt{\pi}}{m_V}\sum\limits_{n=-\infty}^{+\infty}\int\limits_{0}^{\infty}\frac{dt}{\sqrt{t}}
1032: \exp\left[-z^2t-\frac{1+(\omega_n/m_V)^2}{4t}\right]=
1033: \frac{\pi T}{m_V|z|}\sum\limits_{n=-\infty}^{+\infty}{\rm e}^{-|z|\sqrt{1+(\omega_n/m_V)^2}}.$$
1034: Integration of this expression over $d^2z$ is now trivial and yields
1035:
1036: $$
1037: \sigma(T)=\frac{2\pi^2Tv^2}{m_V}\sum\limits_{n=-\infty}^{+\infty}\frac{{\rm e}^{-\frac{m_V}{m_H}\sqrt{1+(\omega_n/m_V)^2}}}
1038: {\sqrt{1+(\omega_n/m_V)^2}}.$$
1039: Performing now with the sum a transformation of the form~(\ref{transform}) in the opposite direction, we obtain
1040:
1041: $$
1042: \sigma(T)=2\pi v^2\sum\limits_{n=-\infty}^{+\infty}K_0\left(\frac{m_V}{m_H}\sqrt{1+(m_H\beta n)^2}\right).$$
1043: In particular, at $T\to 0$, one may approximate the sum by the zeroth term, that recovers the value $\sigma_0=2\pi v^2 L$.
1044:
1045:
1046:
1047:
1048: %\newpage
1049:
1050:
1051: \begin{thebibliography}{99}
1052:
1053: \bibitem{a}
1054: I.~T.~Drummond, \plb{442}{1998}{279}
1055: [arXiv:hep-lat/9808014].
1056: %%CITATION = HEP-LAT 9808014;%%
1057:
1058: \bibitem{b}
1059: I.~T.~Drummond and R.~R.~Horgan, \plb{447}{1999}{298}
1060: [arXiv:hep-lat/9811016].
1061: %%CITATION = HEP-LAT 9811016;%%
1062:
1063: \bibitem{c}
1064: A.~Duncan, E.~Eichten and H.~Thacker, \prd{63}{2001}{111501}
1065: [arXiv:hep-lat/0011076].
1066: %%CITATION = HEP-LAT 0011076;%%
1067:
1068: \bibitem{d}
1069: C.~W.~Bernard {\it et al.}, \prd{64}{2001}{074509}
1070: [arXiv:hep-lat/0103012].
1071: %%CITATION = HEP-LAT 0103012;%%
1072:
1073: \bibitem{e}
1074: C.~DeTar, O.~Kaczmarek, F.~Karsch and E.~Laermann, \prd{59}{1999}{031501}
1075: [arXiv:hep-lat/9808028].
1076: %%CITATION = HEP-LAT 9808028;%%
1077:
1078: \bibitem{1}
1079: D.~Antonov, L.~Del~Debbio, A.~Di~Giacomo, \jhep{08}{2003}{011}
1080: [arXiv:hep-lat/0302015].
1081: %%CITATION = HEP-LAT 0302015;%%
1082:
1083:
1084: \bibitem{f}
1085: O.~Kaczmarek, F.~Karsch, P.~Petreczky and F.~Zantow,
1086: {\it Nucl. Phys. Proc. Suppl.} {\bf B 129} (2004) 560
1087: [arXiv:hep-lat/0309121].
1088: %%CITATION = HEP-LAT 0309121;%%
1089:
1090: \bibitem{15}
1091: A.~A.~Abrikosov, {\it Sov. Phys. JETP} {\bf 5} (1957) 1174;
1092: H.~B.~Nielsen and P.~Olesen, \npb{61}{1973}{45}.
1093: %%CITATION = SPHJA,5,1174;%%
1094: %%CITATION = NUPHA,B61,45;%%
1095:
1096: \bibitem{2}
1097: R.~P.~Feynman, {\it Statistical mechanics. A set of lectures} (Addison-Wesley, Reading, MA, 1972).
1098:
1099: \bibitem{3}
1100: I.~K.~Affleck and N.~S.~Manton,
1101: \npb{194}{1982}{38};
1102: I.~K.~Affleck, O.~Alvarez, and N.~S.~Manton, \npb{197}{1982}{509}.
1103: %%CITATION = NUPHA,B197,509;%%
1104: %%CITATION = NUPHA,B194,38;%%
1105:
1106: \bibitem{4}
1107: R.~D.~Pisarski and F.~Wilczek, \prd{29}{1984}{338};
1108: K.~Rajagopal and F.~Wilczek, \npb{399}{1993}{395} [arXiv:hep-ph/9210253].
1109: %%CITATION = HEP-PH 9210253;%%
1110: %%CITATION = PHRVA,D29,338;%%
1111:
1112:
1113: \bibitem{GL}
1114: J.~Gasser and H.~Leutwyler, \plb{184}{1987}{83}.
1115: %%CITATION = PHLTA,B184,83;%%
1116:
1117:
1118:
1119: \bibitem{correl}
1120: A.~Di Giacomo, M.~D'Elia, H.~Panagopoulos, and E.~Meggiolaro,
1121: {\it Gauge invariant field strength correlators in QCD}, in
1122: ``Vancouver 1998, High energy physics, vol. 2'', p.p. 1809-1813 [arXiv:hep-lat/9808056];
1123: M.~D'Elia, A.~Di Giacomo and E.~Meggiolaro,
1124: \prd{67}{2003}{114504} [arXiv:hep-lat/0205018].
1125: %%CITATION = HEP-LAT 0205018;%%
1126: %%CITATION = HEP-LAT 9808056;%%
1127:
1128:
1129: \bibitem{bras}
1130: H.~Boschi-Filho, C.~P.~Natividade and C.~Farina, \prd{45}{1992}{586}.
1131: %%CITATION = PHRVA,D45,586;%%
1132:
1133:
1134: \bibitem{5}
1135: T.~D.~Lee and G.~C.~Wick, \prd{9}{1974}{2291};
1136: M.~B.~Voloshin, I.~Yu.~Kobzarev, and L.~B.~Okun, {\it Sov. J. Nucl. Phys.} {\bf 20} (1975) 644;
1137: S.~Coleman, \prd{15}{1977}{2929}; C.~G.~Callan~Jr. and S.~Coleman, \prd{16}{1977}{1762}.
1138: %%CITATION = PHRVA,D9,2291;%%
1139: %%CITATION = SJNCA,20,644;%%
1140: %%CITATION = PHRVA,D15,2929;%%
1141: %%CITATION = PHRVA,D16,1762;%%
1142:
1143: \bibitem{6}
1144: M.~B.~Voloshin, {\it Sov. J. Nucl. Phys.} {\bf 42} (1985) 644.
1145: %%CITATION = YAFIA,42,1017;%%
1146:
1147: \bibitem{7}
1148: H.~G.~Dosch, \plb{190}{1987}{177};
1149: Yu.~A.~Simonov, \npb{307}{1988}{512};
1150: H.~G.~Dosch and Yu.~A.~Simonov, \plb{205}{1988}{339}.
1151: %%CITATION = PHLTA,B190,177;%%
1152: %%CITATION = PHLTA,B205,339;%%
1153: %%CITATION = NUPHA,B307,512;%%
1154:
1155: \bibitem{8}
1156: A.~Di~Giacomo, H.~G.~Dosch, V.~I.~Shevchenko, and Yu.~A.~Simonov,
1157: \prep{372}{2002}{319} [arXiv:hep-ph/0007223].
1158: %%CITATION = HEP-PH 0007223;%%
1159:
1160:
1161: \bibitem{9}
1162: A.~Di~Giacomo and H.~Panagopoulos, \plb{285}{1992}{133}.
1163: %%CITATION = PHLTA,B285,133;%%
1164:
1165: \bibitem{10}
1166: A.~Di~Giacomo, {\it Nonperturbative QCD}, in ``Lisbon 1999, QCD: Perturbative or nonperturbative'',
1167: p.p. 1-29 [arXiv:hep-lat/9912016];
1168: {\it Topics in nonperturbative QCD}, in
1169: {\it Czech. J. Phys.} {\bf 51} (2001) B9
1170: [arXiv:hep-lat/0012013]; {\it QCD vacuum and confinement},
1171: in ``Campos do Jordao 2002, New states of matter in hadronic interactions'', p.p. 168-190
1172: [arXiv:hep-lat/0204001].
1173: %%CITATION = HEP-LAT 9912016;%%
1174: %%CITATION = HEP-LAT 0012013;%%
1175: %%CITATION = HEP-LAT 0204001;%%
1176:
1177: \bibitem{12}
1178: A.~M.~Polyakov, \npb{164}{1980}{171}.
1179: %%CITATION = NUPHA,B164,171;%%
1180:
1181: \bibitem{14}
1182: E.~Meggiolaro, \plb{451}{1999}{414}
1183: [arXiv:hep-ph/9807567].
1184: %%CITATION = HEP-PH 9807567;%%
1185:
1186: \bibitem{strrepr}
1187: See e.g. M.~Sato and S.~Yahikozawa, \npb{436}{1995}{100}
1188: [arXiv:hep-th/9406208].
1189: %%CITATION = HEP-TH 9406208;%%
1190:
1191: \bibitem{11}
1192: D.~V.~Antonov, D.~Ebert, and Yu.~A.~Simonov, \mpla{11}{1996}{1905}
1193: [arXiv:hep-th/9605086].
1194: %%CITATION = HEP-TH 9605086;%%
1195:
1196:
1197:
1198: \end{thebibliography}
1199: \end{document}
1200:
1201:
1202:
1203:
1204:
1205:
1206:
1207:
1208:
1209:
1210:
1211:
1212:
1213:
1214:
1215:
1216:
1217:
1218:
1219:
1220:
1221:
1222:
1223:
1224:
1225:
1226:
1227:
1228:
1229:
1230:
1231:
1232:
1233:
1234:
1235:
1236: