1: %\usepackage{feynmf}
2: %\usepackage{amssymbol}
3: %\renewcommand{\baselinestretch}{2}\normalsize
4: %\input{tcilatex}
5: %\input{tcilatex}
6:
7:
8: \documentclass[10pt,preprint]{article}
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: \usepackage{amssymb}
11: \usepackage{amsmath}
12: \usepackage{graphics}
13: \usepackage{graphicx}
14: \usepackage{epsfig}
15: %\usepackage{pdffig}
16: %\usepackage{psfig}
17:
18: \setcounter{MaxMatrixCols}{10}
19: %TCIDATA{OutputFilter=LATEX.DLL}
20: %TCIDATA{Version=5.00.0.2606}
21: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
22: %TCIDATA{BibliographyScheme=Manual}
23: %TCIDATA{LastRevised=Saturday, September 09, 2006 22:24:09}
24: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
25: %TCIDATA{Language=American English}
26:
27: \renewcommand{\vec}[1]{{\bf #1}}
28: \setlength{\baselineskip}{20mm}
29: \textwidth 15.0 true cm
30: \textheight 22.0 true cm
31: \headheight 0 cm
32: \headsep 0 cm
33: \topmargin 0.4 true in
34: \oddsidemargin 0.25 true in
35: \def\bra{\langle}
36: \def\ket{\rangle}
37: \newcommand{\eqb}{\begin{equation}}
38: \newcommand{\eqe}{\end{equation}}
39: \newcommand{\dmb}{\begin{displaymath}}
40: \newcommand{\dme}{\end{displaymath}}
41: \newcommand{\pd}{\partial}
42: \newcommand{\ep}{\varepsilon}
43: \newcommand{\eab}{\begin{eqnarray}}
44: \newcommand{\eae}{\end{eqnarray}}
45: \newcommand{\ra}{\right\rangle}
46: \newcommand{\la}{\left\langle}
47: \newcommand{\e}{\mbox{e}}
48: \newcommand{\be}{\begin{equation}}
49: \newcommand{\ee}{\end{equation}}
50: \newcommand{\sgn}{\text{sgn}\,}
51: \newcommand{\munu}{{\mu\nu}}
52: \newcommand{\ad}{{\dot{\alpha}}}
53: \newcommand{\bd}{{\dot{\beta}}}
54: \newcommand{\La}{\Lambda}
55: \setlength{\unitlength}{1mm}
56:
57: %\input{tcilatex}
58:
59: \begin{document}
60:
61:
62: \begin{titlepage}
63: \begin{flushright}
64: .
65: \end{flushright}
66: \vspace{0.2cm}
67: \begin{center}
68: \Large{Spatial Wilson loop in continuum, deconfining SU(2) Yang-Mills thermodynamics}
69: \vspace{0.5cm}\\
70: \large{Josef Ludescher$^*$, Jochen Keller$^*$,
71: Francesco Giacosa$^\dagger$, and\\
72: Ralf Hofmann$^{*}$}
73: \end{center}
74: \vspace{0.5cm}
75: \begin{center}
76: {\em $\mbox{}^*$ Institut f\"ur Theoretische Physik\\
77: Universit\"at Heidelberg\\
78: Philosophenweg 16\\
79: 69120 Heidelberg, Germany}
80: \end{center}
81: \vspace{0.1cm}
82: \begin{center}
83: {\em $\mbox{}^\dagger$ Institut f\"ur Theoretische Physik\\
84: Universit\"at Frankfurt\\
85: Johann Wolfgang Goethe - Universit\"at\\
86: Max von Laue--Str. 1\\
87: 60438 Frankfurt, Germany}
88: \end{center}
89: \vspace{2.8cm}
90: \begin{abstract}
91:
92: The uniquess of the effective actions describing 4D SU(2) and
93: SU(3) continuum, infinite-volume Yang-Mills thermodynamics in their deconfining
94: and preconfining phases is made explicit. Subsequently,
95: the spatial string tension
96: is computed in the approach proposed by
97: Korthals-Altes. This SU(2) calculation is
98: based on a particular, effective two-loop correction to
99: the pressure needed for the extraction of the hypothetic number density
100: of isolated and screened magnetic monopoles or antimonopoles in the
101: deconfining phase.
102: By exponentiating the exchange of the tree-level massless but one-loop
103: dressed mode within a quadratic
104: spatial contour of side-length $L$ in the effective theory we demonstrate that for $L\to\infty$
105: the Wilson loop exhibits {\sl perimeter} law. This is in contrast to a rigorous
106: {\sl lattice} result subject to the Wilson action and for this action
107: valid at sufficiently high temperature. In the framework of the effective theory
108: there is, however, a regime for small
109: (spatially unresolved) $L$ were the exponent of the spatial Wilson loop
110: possesses curvature as a function of $L$.
111:
112: \end{abstract}
113:
114: \end{titlepage}
115:
116: \section{Introduction\label{intro}}
117:
118: In the past the pursuit of a nonperturbative
119: understanding of 4D Yang-Mills theory at high
120: temperature was limited to lattice formulations. A prominent
121: target of lattice investigations is the spatial Wilson loop evaluated on a
122: rectangular and planar contour $C$. Based on an analogy to the
123: strong-coupling expansion, whose validity is suggested at sufficiently
124: high temperature by the part of the Wilson action containing
125: purely spacelike plaquettes, it was shown in \cite{Borgs1985}
126: that at a given, finite, spatial lattice spacing the spatial Wilson loop always
127: exhibits area law implying the existence of a spatial string tension
128: $\sigma_s$. For the gauge group SU(2), which we will mainly be concerned
129: with here, this
130: result was verified subsequently by direct
131: simulation \cite{Polonyi1987,Bali1993,spatialWL}.
132:
133: As already pointed out in \cite{Polonyi1987} the argument made
134: in \cite{Borgs1985} neither has much to say about the cutoff dependence of
135: $\sigma_s$ in the ultraviolet nor about the behavior of $\sigma_s$
136: in the infinite-volume limit since it
137: inherently relies on a finite spatial discretization allowing, at
138: leading order in the strong-`coupling'-expansion, to
139: tile the minimal area spanned by $C$ in terms of spacelike
140: plaquettes. This then implies the area law in the lattice
141: formulation.
142:
143: In contrast to lattice formulations,
144: the nonperturbative approach to SU(2) and SU(3) Yang-Mills
145: thermodynamics developed in
146: \cite{Hofmann2005,HofmannGiacosa2006,HofmannLE2006,Hofmann2007} uniquely derives an
147: effective theory for the deconfining phase starting from a genuine
148: spacetime continuum, space being infinitely extended, and from the
149: fundamental (euclidean) Yang-Mills action $S_{\tiny\mbox{YM}}$. The effective theory
150: emerges as a consequence of pursuing the following chain of steps:
151: 1) Consider the BPS saturated (topologically nontrivial) sector of (periodic) euclidean field
152: configurations (time variable imaginary). Independent of their topological charge
153: these (anti)selfdual configurations possess vanishing energy-stress.
154: This implies that 2) whatever effective field is obtained as a consequence
155: of performing a spatial coarse graining over them it is {\sl nonpropagating}.
156: But a nonpropagating effective field acts as a static background. 3) Such a background configuration would break the spatial isotropy and/or homogeneity of the thermal system (one-point functions) unless it is a scalar
157: field. That is, no Lorentz tensor of rank greater than zero may emerge as an effective field from the sector
158: of fundamental BPS saturated field configurations. If this effective scalar would be neutral
159: under the gauge group then it would not participate in the thermodynamics due
160: to its decoupling from the $Q=0$ sector of propagating field configurations and its vanishing effective energy-stress. 4) Since this scalar field does
161: not propagate its classical equation of motion is, in a given gauge,
162: solely determined by the euclidean time dependence of its phase. 5) It turns out \cite{HerbstHofmann2004,Hofmann2007} that only a single nonlocal
163: definition of this time dependence, involving an adjointly transforming two-point function of the fundamental field strength, is possible. 6) This definition only admits the contribution of configurations with topological charge modulus $|Q|=1$ (Harrington-Shepard (anti)calorons, trivial holonomy) because of too many dimensionful parameters spanning the moduli spaces\footnote{Due to the temperature independence
164: of the weight $\e^{8\pi^2|Q|/g^2}$ ($g$ the fundamental, classical and thus temperature independent coupling {\sl constant} the definition of the time dependence of the
165: field's phase must not invoke any explicit temperature dependence. This only leaves
166: room for the spatial coarse-graining over a two-point function of the fundamental field strength.} of higher-charge calorons and the
167: instability of (anti)calorons of nontrivial holonomy \cite{Diakonov,GPY1981}.
168: 7) Consistency of the thus derived second-order equation and the effective BPS saturation (first-order equation) of the adjoint scalar field uniquely fixes its gauge invariant
169: potential with the Yang-Mills scale $\Lambda$ entering as a (nonperturbative) integration
170: constant \cite{HofmannGiacosa2006, Hofmann2007}. 8) Since no field other than the inert
171: adjoint scalar may emerge from the sector(s) of BPS saturated fundamental
172: field configurations the coupling to the sector of propagating gauge fields
173: ($Q=0$), which by perturbative renormalizability \cite{tHooftVeltman}
174: by itself cannot generate higher dimensional operators upon the invoked
175: spatial coarse-graining, occurs minimally via the square of the adjoint
176: covariant derivative, see also Sec.\,\ref{Uni}. 9) The thus derived effective
177: action yields an accurate a priori estimate for the thermal ground state
178: by solving the Yang-Mills equation of the effective $Q=0$ sector
179: in the background of the adjoint scalar field. 10) Radiative corrections to the free quasiparticle situation described by the effective action can be integrated out in a rapidly converging loop expansion subject to constraints imposed by the existence of an effective thermal ground state \cite{HofmannLE2006}. Interestingly, only planar bubble
180: diagrams contribute. As we will see, the maximal resolution of the
181: effective theory, which is given by the scalar field's modulus, almost
182: everywhere in temperature is too small to allow for the existence and dynamics of single magnetic monopoles and antimonopoles, generated by (anti)caloron deformation away from trivial holonomy and away from BPS saturation, to be resolved.
183:
184: We consider it useful to repeat the above steps including physical interpretations.
185: A sufficiently local spatial coarse-graining \cite{HerbstHofmann2004}
186: over stable and absolute minima of the
187: action in the sector with topological charge modulus $|Q|=1$
188: (Harrington-Shepard (anti)calorons \cite{HS1977}, nonpropagating
189: configurations) generates a spatially homogeneous and inert adjoint
190: scalar field $\phi$ whose modulus sets the maximally available
191: resolution in the effective theory\footnote{A nontrivial-holonomy
192: (anti)caloron \cite{NahmVanBaalLeeLu} is not
193: stable. Moreover, it is suggested to consider a slow dependence in real time of
194: the holonomy parameter $u$ since a constancy of $u$ in space and time would completely
195: suppress the weight of the nontrivial-holonomy caloron in the thermodynamic limit
196: \cite{GPY1981}. This time dependence, however, necessarily lifts the associated field
197: configuration above the BPS bound. On the level of the effective theory,
198: where the thermal ground state is described by the field $\phi$ together
199: with a pure-gauge configuration of the coarse-grained $Q=0$ sector, this
200: microscopic lift above the BPS bound is seen in terms a positive ground-state
201: energy density which is the negative of its pressure
202: \cite{Hofmann2005,Hofmann2007}.}. Recall, that only this
203: sector is allowed to contribute to the kernel containing $\phi$'s
204: phase \cite{Hofmann2005,Hofmann2007,HerbstHofmann2004}.
205: The inertness of the field
206: $\phi$ and the spatial homogeneity of its (gauge-invariant) modulus
207: derives from the fact that it is obtained as a spatial average over
208: nonpropagating, stable BPS saturated (zero energy-stress) field
209: configurations\footnote{Inertness of $\phi$ follows from the fact that a spatial
210: average over nonpropagating fields can not generate propagating modes in
211: the effective field $\phi$. Thus an inhomogeneity of $|\phi|$ would
212: explicitely break the homogeneity of one-point functions which is in
213: contradiction to infinite-volume continuum thermodynamics.}. This can
214: be checked explicitely by computing the curvature of $\phi$'s potential and by
215: comparing it with the squares of maximal resolution $|\phi|$ in the effective
216: theory and $T$ \cite{Hofmann2005}. So the effect of performing the spatial
217: coarse graining over Harrington-Shepard (anti)calorons is to make explicit the
218: spatial homogeneity and isotropy of the thermal system as contributed to
219: by the only topologically nontrivial gauge-field
220: configurations which are admissible in computing a useful a priori estimate of the thermal ground state. The average effect of domainizations of
221: the field $\phi$, which necessarily would lead to the
222: emergence\footnote{Microscopically, the emergence of these charges is
223: understood by the dissociation of large-holonomy, that is,
224: strongly-deformed-away-from-the-Harrington-Shepard-situation
225: (anti)calorons.} of spatially localized and isolated magnetic charges \cite{Kibble}, is accounted for by the quantum field theoretic method of loop
226: expansion \cite{HofmannLE2006,SHG2006,KH2007,KHG2007}.
227: Again, to each order of this expansion, spatial isotropy
228: and homogeneity are granted features. The sector with $Q=0$ and its
229: coupling to the inert field $\phi$ after spatial coarse-graining
230: is uniquely determined by perturbative
231: renormalizability \cite{tHooftVeltman}, gauge invariance, and the
232: spacetime symmetries of the underlying theory, for an explicit
233: presentation see Sec.\,\ref{Uni}.
234:
235: Apart from 1-PI reducible diagrams for the polarization tensor,
236: which as compared to the tree-level situation can be resummed into mildly modified
237: dispersion laws, the expansion
238: into irreducible loop diagrams is expected to terminate at a finite loop order \cite{HofmannLE2006}. Numerically, there is a large hierarchy
239: between leading, next-to-leading, and next-to-next-to-leading
240: corrections to the pressure \cite{SHG2006,KH2007}. As far as propagating,
241: effective gauge modes are concerned, this is the reason
242: why essentially all physics is contained in the
243: tree-level quasi-particle spectrum and the resummed,
244: one-loop polarization \cite{SHG2006,KH2007,LH2008}.
245:
246: Interactions between
247: two topologically distinct
248: sectors are hard to grasp when working with fundamental fields. As it
249: turns out, in a situation determined by a global temperature scale $T$
250: (infinite-volume thermodynamics) the explicit
251: consideration of fundamental fields and their interactions
252: would necessitate external probes whose resolution needed to exceed the
253: value $|\phi|$ emerging in terms of the Yang-Mills scale $\Lambda$ and $T$
254: when (sufficiently locally) integrating out the sector with $|Q|=1$.
255: That is, an attempt to saturate the partition function in terms of
256: fundamental field
257: configurations, resolved with a higher momentum transfer than $|\phi|$,
258: necessarily introduces unsurmountable complications in assuring
259: that the according approximations sustain
260: homogeneous and isotropic thermalization.
261:
262: The main purpose of the present work is to
263: investigate the physics of screened magnetic
264: charges, being held responsible for the emergence of an area
265: law for the spatial Wilson loop at high temperatures in lattice gauge
266: theory. We perform the analysis both in a hypothetic microscopic fashion and by directly appealing to
267: the effective theory. As a warm-up to the effective theory for the deconfining
268: phase we start by showing the uniqueness of the effective action in
269: Sec.\,\ref{Uni}. In Sec.\,\ref{KA} we review the argument given by
270: Korthals-Altes for the existence of a spatial string tension as
271: orginated by isolated, screened and statistically independent
272: magnetic charges. This argument assumes (anti)monopoles to behave like
273: classical (resolvable) particles which, due to sufficiently strong
274: screening, separately
275: contribute their magnetic flux through the minimal surface spanned by
276: the spatial contour. In Sec.\,\ref{MP} we use a two-loop correction
277: to the pressure, calculated in the effective theory and surviving the
278: high-temperature limit, to extract the average mass of the
279: screened and stable monopole-antimonopole system. Subsequently, their
280: hypothetic\footnote{By hypothetic we mean a physical number
281: density seen by applying a sufficiently large spatial
282: resolution $a^{-1}$ to typical configurations thermalized
283: according to an associated perfect lattice action. The latter is
284: constructed such that
285: the full partition function of the theory is invariant under a change
286: $a\to a^\prime$. Due to an efficient magnetic-charge neutralization
287: it will turn out that this number density appears to be
288: zero when applying the low resolution $|\phi|$
289: of our effective theory. The practical advantage of the
290: low resolution $|\phi|$ is that it renders the perfect action to be
291: extremely simple and thus useful: The assumptions of \ref{KH} about screened
292: magnetic monopoles and antimonopoles contributing their magnetic flux essentially uncancelled
293: through the minimal contour spanning the spatial loop can be quantitatively
294: checked by comparing screening length with mean spatial separation.} number density
295: $n$ is determined. As an aside,
296: we demonstrate that, despite the hypothetic screening length being about three times
297: larger than the intermonopole distance in an {\sl ensemble} of, isolated, stable and screened
298: (anti)monopoles there is no large effect on their mass when comparing
299: with the BPS
300: mass of an {\sl isolated} monopole-antimonopole pair immersed into a sea of
301: instable dipoles. The hypothetic number density $n$ in turn would imply a spatial string tension if the potentials of these
302: stable magnetic (anti)monopoles would not overlap such as to cancel their
303: magnetic flux. Recall that the
304: derivation of an area law, as performed
305: in \cite{Korthals-Altes}, relies on a counting argument for
306: Poisson distributed (anti)monopoles. This argument assumes
307: the independence of individual (anti)monopoles contributing to the total
308: magnetic flux through a given spatial surface. This independence, however, turns out to be strongly violated
309: because stable magnetic objects essentially cancel their fluxes locally through their
310: overlaping magnetic potentials. The immediate implication then is
311: that the spatial Wilson
312: loop cannot exhibit area-law behavior at high temperatures. We confirm this result
313: in Sec.\,\ref{SWLE} where
314: the spatial Wilson loop is computed involving the lowest
315: nontrivial radiative corrections for the dispersion law of the massless mode
316: in the effective theory: A
317: resummation of the one-loop polarization of the massless mode both
318: on- and off-shell. In
319: computing the spatial Wilson loop in real time
320: we take into account the quantum and the thermal parts of
321: the modified propagator for the massless mode, and we discuss the
322: potential contribution of the massive modes. We also
323: consider the magnetic screening mass, defined as the zero-momentum
324: limit of the square root of the screening function when setting the
325: temporal component of the momentum equal to zero
326: (Secs.\,\ref{screeningmass} and \ref{vacuumpart}). In Sec.\,\ref{sum} we summarize our
327: work and conclude.
328:
329: \section{Only minimal coupling of $Q=0$ with $|Q|=1$ sector in the effective action for the deconfining phase\label{Uni}}
330:
331: Here we would like to argue that the effective action of the deconfining phase
332: %*************
333: \eqb
334: \label{effdec}
335: S_{\tiny\mbox{dec}}=\mbox{tr}\,\int d^4x\,\left\{\frac12\,
336: G_{\mu\nu}G_{\mu\nu}+D_\mu\phi D_\mu\phi+\frac{\Lambda^6}{\phi^2}\right\}\,.
337: \eqe
338: %*************
339: is complete. Namely, no operators representing local vertices of the field $a_\mu$ with three or
340: more external legs together with (powers of) the field $\phi$ are possible while the nonexistence of operators of mass
341: dimension larger than four {\sl solely} involving the gauge field $a_\mu$ is excluded
342: by perturbative renormalizability \cite{tHooftVeltman}. Notice that nonlocal contributions
343: to the effective action always can be expanded in powers of covariant derivatives.
344:
345: Examples for excluded operators, still allowed by the spacetime symmetries and gauge invariance, would be
346: %*******
347: \eqb
348: \label{excops}
349: \mbox{tr}\,\frac{1}{M^{2(3n-2)}}\left(G_{\mu\nu}[D_\mu\phi,D_\nu\phi]\right)^n\,,\ \ (n\ge 1)\,,
350: \eqe
351: %********
352: $M$ representing some mass scale, while the operator $\mbox{tr}\,D_\mu\phi D_\mu\phi$ appearing in the action of
353: Eq.\,(\ref{effdec}) is allowed. Fig.\,\ref{Fig-0A} depicts the situation of the allowed operator
354: $\mbox{tr}\,D_\mu\phi D_\mu\phi$ and the $n=1$ operator of (\ref{excops}), which suffices to state our argument
355: diagrammatically.
356: %***********************
357: \begin{figure}
358: \begin{center}
359: \leavevmode
360: %\epsfxsize=9.cm
361: \leavevmode
362: %\epsffile[80 25 534 344]{}
363: \vspace{4.9cm}
364: \special{psfile=Fig-0A.ps angle=0 voffset=-140
365: hoffset=-210 hscale=65 vscale=60}
366: \end{center}
367: \caption{\protect{\label{Fig-0A}} Allowed vertex (a) and examples of lowest mass dimension for excluded
368: vertices in the effective theory (b). Solid lines represent the topologically trivial gauge field $a_\mu$ while a dashed
369: line terminating in a cross corresponds to a local insertion of the operator $\phi$. }
370: \end{figure}
371: %************************
372: Field $\phi$ is energy and
373: pressure free. (Its existence is owed to $|Q|=1$ (anti)selfdual fundamental
374: configurations of trivial holonomy.) Assuming a local vertex of $\phi$
375: with more than three external legs (in consistency with gauge
376: invariance and spacetime symmetries) coming from the gauge fields of the
377: coarse-grained trivial-topology sector $a_\mu$ necessarily would imply momentum
378: transfer to the field $\phi$, see (b) of Fig.\,\ref{Fig-0A} for a lowest-dimension example. That is, the probability
379: of not transferring momentum to $\phi$ for nonvanishing momenta associated with the external legs
380: of field $a_\mu$ is zero since this situation would correspond to a hypersurface in the space of all possible momentum transfers. But a transfer of energy-momentum to $\phi$ would change its own
381: energy-momentum to nonvanishing values. This, however,
382: contradicts $\phi$'s BPS property. For the vertex of $\phi$ with
383: two external legs of the field $a_\mu$ (mass term in unitary gauge), as it is represented by
384: the term tr\,$(D_\mu\phi)^2$ in the effective
385: action of Eq.\,(\ref{effdec}) and depicted in (a) of Fig.\,\ref{Fig-0A}, no momentum is transferred. (Massiveness of the
386: off-Cartan topologically trivial gauge field $a_\mu$ emerges only after an
387: infinite resummation of such a vertex.) The fact that the ground state as
388: such has a finite energy density in the effective theory is owed to a pure
389: gauge configuration. This configuration on the effective level sums up interactions
390: (energy-momentum transfers {\sl larger} than $|\phi|$, that is at distances smaller than $|\phi|^{-1}$) of topologically
391: trivial fundamental gauge fields with (anti)calorons which are not visible
392: at resolution lower than $|\phi|$ but lift the ground-state energy density from zero to a finite
393: value (temporary shift of holonomy plus radiative corrections).
394:
395: To conclude, the existence of vertices involving $\phi$ and more than
396: two legs of the coarse-grained topologically trivial gauge field $a_\mu$ in the effective theory for the deconfining phase
397: is excluded by the BPS property of $\phi$: The very existence of $\phi$,
398: which is guaranteed to emerge from the $|Q|=1$ sector of the theory upon
399: spatial coarse-graining \cite{HerbstHofmann2004,Hofmann2005,Hofmann2007},
400: would be contradicted by it absorbing energy-momentum {\sl after} coarse-graining.
401:
402: As for the preconfining phase, renormalizability is trivial since we
403: coarse-grain over the noninteracting $Q=0$ sector of a U(1) (or U(1)$^2$
404: in the case SU(3)) gauge
405: theory. Thus the same arguments as for the deconfining phase when
406: adjusted to this simpler, abelian situation do fix the action of the
407: associated Higgs model \cite{Hofmann2005} uniquely in the preconfining phase.
408:
409:
410: \section{Korthals-Altes' derivation of fundamental spatial string
411: tension\label{KA}}
412:
413: We only consider the case of deconfining SU(2) Yang-Mills thermodynamics
414: with Yang-Mills scale $\Lambda$. In
415: the effective theory for this phase the tree-level situation is described by an
416: inert adjoint and spatially homogeneous scalar field of modulus
417: $|\phi|=\sqrt{\frac{\Lambda^3}{2\pi T^3}}$,
418: a pure-gauge configuration of the coarse-grained $Q=0$ sector,
419: and propagating gauge modes (one direction of the
420: SU(2) algebra massless, two directions of mass $2e|\phi|$). This
421: effective theory neatly decomposes physics into
422: a thermal ground state (equation of state equal to that of a cosmological constant
423: but with a temperature dependent energy density) and thermally fluctuating
424: (quasi)particles \footnote{Due to the existence of the maximal resolution
425: scale $|\phi|$ the effect of having coarse-grained gauge modes propagate off
426: their mass shell without explicit interaction is, on the one-loop level, completely negligible for
427: thermodynamical quantities such as the pressure \cite{Hofmann2005}.}.
428: Notice that on tree-level this `a-priori-estimate' for the thermal ground state only
429: captures the physics of small-holonomy (anti)calorons which are associated
430: with short-lived monopole-antimonopole pairs. This leads to an
431: explicit manifestation of spatial homogeneity and isotropy in terms of $\phi$
432: and a pure gauge $a^{gs}_\mu$ as it is demanded
433: by thermodynamics. The effects attributed to the
434: rare occurrence of isolated, long-lived, and screened
435: magnetic charges in the plasma are captured by certain radiative corrections. This
436: can be pictured as an implicit average over domainized $\phi$-field
437: configurations \cite{Kibble} of typical domain size smaller than $|\phi|^{-1}$
438: facilitated by the quantum field theoretic
439: method of loop expansion in effective variables. Notice that spatial
440: homogeneity and isotropy are explicitely honored at each loop order.
441:
442: Before we actually establish the connection between radiative corrections to
443: the pressure as computed in the effective theory and the hypothetic
444: physics of screened magnetic
445: monopoles we would like to review the arguments given by Korthals-Altes
446: \cite{KH}
447: on how a spatial string tension emerges once the existence and
448: independence of
449: isolated and screened (magnetic screening length $l_s$) magnetic charges is assumed in the
450: plasma. Korthals-Altes considers a quadratic, spatial contour of
451: side-length $L$ and a number density of monopoles $n_M$. Since monopoles are
452: created pairwise by the dissociation of large-holonomy calorons
453: \footnote{Notice that the occurrence of a nontrivial caloron holonomy in the sense
454: of a spacetime independent parameter, as it occurs in (anti)selfdual Yang-Mills
455: fields, is unphysical because of the vanishing impact of these configurations
456: onto the partition function \cite{GPY1981}. In case of small
457: holonomy it is still
458: useful though to think of the holonomy parameter as possessing a
459: finite-support real-time dependence of width that is comparable to the
460: life-time of the monopole-antimonopole pair. Upon a continuation of this real-time
461: dependence back to the euclidean a departure from
462: (anti)selfduality takes place.} one has $n_M=n_A\equiv n$ where $n_A$ is the
463: number density of antimonopoles. Furthermore, he assumes (and justifies this
464: assumption by lattice data obtained with the Wilson action)
465: that monopoles and
466: antimonopoles fluctuate in an independent way and that they are rare. This is
467: certainly true if the screening length $l_s$ is much smaller than the mean
468: interparticle distance $d=\left(\frac{1}{n}\right)^{1/3}$. We will show
469: in Sec.\,\ref{dens}, however, that this situation is actually {\sl not}
470: realized. Still, the mass of the screened
471: monopole-antimonopole system
472: is close to its BPS bound.
473:
474: By virtue of the length scale $l_s$ Korthals-Altes considers
475: a slab of thickness $2l_s$ containing magnetic quasiparticles. These quasiparticles contribute a mean
476: magnetic flux through the minimal surface spanned by the afore-mentioned
477: contour. Only
478: those quasiparticle that are contained within the slab are held responsible
479: for the flux. A more refined treatment introducing no such
480: constraint but taking into account
481: a Yukawa-like potential for the static magnetic field sourced by each of these
482: objects
483: yields similar results.
484:
485: As a brief interlude, we name an important property of these
486: monopole-antimonopole pairs. Since we know now that pairs of magnetic monopoles and
487: antimonopoles are liberated through the dissociation of large-holonomy
488: (anti)calorons we also know -- by studying the BPS (anti)monopole constituents in
489: such (anti)calorons \cite{NahmVanBaalLeeLu} --
490: that the combined mass $m\equiv m_M+m_A$ of the monopole and its
491: antimonopole (a holonomy-independent quantity) is roughly given as
492: \cite{NahmVanBaalLeeLu}
493: %***********
494: \eqb
495: \label{monAmonMass}
496: m\sim\frac{8\pi^2 T}{e}=\sqrt{8}\pi\,T\sim 8.89\,T\,.
497: \eqe
498: %***********
499: Here the high-temperature plateau-value $e=\sqrt{8}\,\pi$
500: for the {\sl effective} gauge coupling $e$ was used
501: \cite{Hofmann2005,Hofmann2007,GiacosaHofmann2007}.
502: Eq.\,(\ref{monAmonMass}) is valid for an isolated, noninteracting (test)
503: monopole-antimonopole system with their magnetic charges being
504: screened by the surrounding, short-lived magnetic dipoles belonging to
505: small-holonomy (anti)calorons. This is the situation described by the
506: one-loop expressions for thermodynamical quantities in the effective theory.
507: By virtue of Eq.\,(\ref{monAmonMass}) magnetic monopoles and
508: antimonopoles appear as nonrelativistic objects in the plasma.
509: They are Boltzmann suppressed as
510: %************
511: \eqb
512: \label{Boltzmann}
513: P_{M+A}\sim \exp(-\sqrt{8}\pi)\sim 1.4\times 10^{-4}\,,
514: \eqe
515: %************
516: and thus occur rarely as compared to the instable pairs
517: that are associated with small (anti)caloron holonomies.
518:
519: Now Korthals-Altes exponentiates the
520: normalized flux $\Phi_{l=1}$ of a single (anti)monopole through the minimal
521: surface spanned by the contour in the limit $L\to\infty$ as
522: %***********
523: \eqb
524: \label{V(L)}
525: \left.V(L)\right|_{l=1}\equiv \exp\left(2\pi i \frac{\Phi_{l=1}}{Q}\right)\,,
526: \eqe
527: %***********
528: where $Q$ is the magnetic charge.
529: Depending on the sign of $Q$ and on the location w.r.t. the minimal
530: surface one has $\Phi_{l=1}=\pm \frac12$ and thus
531: %***********
532: \eqb
533: \label{V(L)aus}
534: \left.V(L)\right|_{l=1}\equiv -1\,.
535: \eqe
536: %***********
537: Korthals-Altes now assumes
538: the probability $P(l)$ for the occurrence of $l$
539: charges within the slab to be given by the Poisson distribution
540: %**********
541: \eqb
542: \label{Poisson}
543: P(l)=\frac{\bar{l}^l}{l!}\,\exp(-\bar{l})\,,
544: \eqe
545: %***********
546: where $\bar{l}$ is the mean value of the number of magnetic charges
547: contained inside the slab. One then obtains
548: %**********
549: \eqb
550: \label{meanflux}
551: \bar{V}(L)=\sum_{l=0}^\infty P(l) (-1)^l=\exp(-2\bar{l})\,.
552: \eqe
553: %***********
554: Since $\bar{l}=4\,A(L)l_s n(T)$, where $A(L)$ is the minimal surface and
555: $n(T)$ is the (temperature-dependent) density of monopoles (or antimonopoles
556: or monopole-antimonopole pairs), we observe that the Wilson loop shows area
557: law with tension $\sigma_s$ given as
558: %**********
559: \eqb
560: \label{sigma_s}
561: \sigma_s=8\,l_s n(T)\,.
562: \eqe
563: %***********
564: Notice that
565: the hypothetic magnetic screening length $l_s$ is given in terms of $n$
566: and $T$ as \cite{KH}
567: %**********
568: \eqb
569: \label{magscl}
570: l_s=\frac{1}{g}\sqrt{\frac{T}{n}}\,,
571: \eqe
572: %***********
573: where $g$ is the magnetic coupling: $g\equiv\frac{4\pi}{e}$.
574:
575:
576: \section{Monopole physics from two-loop correction to the pressure \label{MP}}
577:
578: In this section we would like to investigate the high-temperature
579: implications of the radiative corrections to the pressure, calculated in
580: the effective theory, for the physics of unresolved, stable and screened
581: magnetic monopoles and antimonopoles as they emerge from the
582: dissociation of large-holonomy (anti)calorons. Notice
583: that the density of these objects is inferred purely
584: energetically without making reference to the existence or nonexistence
585: of their net magnetic fluxes.
586:
587:
588: \subsection{Mass of interacting monopole-antimonopole system}
589:
590: On distances larger than the minimal spatial length \cite{Hofmann2005}
591: %************
592: \eqb
593: \label{phiinTc}
594: |\phi|^{-1}=\frac{(13.87)^{3/2}}{2\pi T_c}\,\sqrt{\frac{T}{T_c}}
595: \eqe
596: %***********
597: and not taking into account radiatively liberated stable and isolated
598: monopole-antimonopole pairs, the depletion of a single magnetic test charge is
599: described by the effective, electric coupling $e=\sqrt{8}\pi$ for $T\gg
600: T_c$. Notice that the (constant) value of $e$ signals that for $T\gg T_c$ the
601: system forgets about the presence of the Yang-Mills scale
602: $\Lambda$ as far as its propagating degrees of freedom are concerned: The
603: thermodynamics of these modes then solely is determined by
604: topology and temperature. This fact may be conceived as a
605: nonperturbative manifestation of asymptotic freedom.
606:
607: It is important to stress that the effective, electric coupling rapidly
608: approaching the constant $e=\sqrt{8}\pi$ at increasing temperature is a
609: consequence of an evolution equation assuring that the interaction of the thermal ground state with propagating quasiparticle excitations honours Legendre transformations
610: \cite{Hofmann2005, Hofmann2007}. The constancy $e=\sqrt{8}\pi$ is approached
611: power-like fast in $T$. So the error of using the constant
612: $e=\sqrt{8}\pi$, which is a lower bound on the behavior close to the phase transition, at moderate temperatures dies off in a power-like way. In \cite{SHG2006}
613: we have used the exact evolution of $e$ and, indeed, have observed, within a few percent variation, the asymptotic constancy of the two-loop correction to the pressure divided by $T^4$ starting from $T=3\,T_c$. Asymptotic constancy of $e$ is not to be confused with the behavior of the {\sl fundamental} gauge coupling which can be defined via the trace anomaly of the energy-momentum tensor \cite{GiacosaHofmann2007}. The latter approaches zero logarithmically
614: slowly with increasing $T$.
615:
616: The holonomy-independent sum $m$ of the masses of the BPS monopole and
617: BPS antimonopole (BPS mass), being the constituents of a nontrivial-holonomy (anti)caloron, is given
618: as \cite{Hofmann2005,NahmVanBaalLeeLu}
619: %************
620: \eqb
621: \label{sumofmasses}
622: m_{>|\phi|^{-1}}=\frac{8\pi^2 T}{e}=\sqrt{8}\pi\,T\,.
623: \eqe
624: %***********
625: Notice that Eq.\,(\ref{sumofmasses}) describes the situation of a pair of BPS
626: saturated monopole and antimonopole placed as test charges into a surrounding where
627: small-holonomy (anti)calorons generate short-lived magnetic
628: dipoles effectively leading to a finite renormalization of the
629: magnetic charge of the test particles. Here no departure from the BPS
630: limit is implied, and the U(1) gauge
631: field of the test (anti)monopole still is infinite-range. A linear
632: superposition of these potentials then leads to a dipole form which for
633: large distances $R$ decays like $1/R^3$. This
634: corresponds to the approximation of a massless and
635: two massive {\sl free} thermal quasiparticles in the
636: effective theory (tree-level).
637:
638: On the level of radiative corrections, however, we are
639: concerned with the physics of isolated, stable magnetic charges whose average
640: distance $n^{-1/3}$ at high temperature will be given by $n^{-1/3}=c/T$
641: where $c$ is a positive, real constant (determined in
642: Sec.\,\ref{dens}). Thus
643: %************
644: \eqb
645: \label{ratiolength}
646: |\phi|\,n^{-1/3}=2\pi c\,\left(\frac{T_c}{13.87\, T}\right)^{3/2}\,,
647: \eqe
648: %************
649: which is smaller than unity for sufficiently high $T$. In fact, an
650: estimate implies that for $T\ge 1.91\,T_c$ isolated and screened
651: (anti)monopoles are not resolved in the effective theory.
652: This estimate uses the high-temperature value
653: $e=\sqrt{8}\,\pi$ for the effective gauge coupling. A more careful
654: investigation for $T$ shortly above $T_c$ shows that isolated and screened
655: (anti)monopoles actually are never fully resolved in the effective theory for the
656: deconfining phase. However, their effect on the propagation of the tree-level
657: massless gauge mode is sizable at temperatures a few times $T_c$ \cite{SHG2006,LH2008}.
658:
659: To make contact with degrees of freedom, whose
660: collective long-distance effects are detectable (antiscreening and screening of effective
661: gauge-field propagation) but which never appear
662: explicitely in the effective theory, we may consider the situation of thermally fluctuating,
663: free quasiparticles (one-loop truncation of loop expansion of the
664: pressure)
665: as the asymptotic starting
666: point. Small interactions between the quasiparticles, as they are described by
667: radiative corrections (for $T\gg T_c$ only the two-loop diagram
668: involving a four-vertex connecting on shell a massless with
669: either of the two massive modes survives) introduce interesting
670: physics but do not change the energy density of the
671: one-loop situation. The important implication then is that radiative
672: corrections to the one-loop situation, calculable in the effective theory,
673: must be {\sl cancelled} by the physics of unresolved degrees of freedom on the
674: level of thermodynamic quantities. By
675: the result of \cite{Diakonov}, which shows that (anti)calorons dissociate upon
676: strong quantum deformation into isolated magnetic monopole--antimonopole pairs, we are led
677: to conclude that these degrees of freedom are indeed isolated, stable, and screened magnetic
678: monopoles and antimonopoles.
679:
680: Thus we need to compute the mass $m$ of monopole-antimonopole systems for distances of the order of
681: $n^{-1/3}$. Since the spatial coarse-graining, which determines the field
682: $\phi$, saturates exponentially fast on distances of a few $\beta\equiv 1/T$
683: \cite{Hofmann2007,HerbstHofmann2004} we expect that
684: %************
685: \eqb
686: \label{ratiomasses}
687: \frac{m}{m_{>|\phi|^{-1}}}=O(1)
688: \eqe
689: %************
690: which, indeed, is the case. Let us show this.
691:
692: As mentioned above, at large temperatures the two-loop diagram for the pressure
693: involving a four-vertex connecting a massless with
694: either of the two massive modes is the only surviving
695: radiative correction \cite{SHG2006,HerbstHofmannRohrer2004}. For $T\gg T_c$
696: one has \cite{SHG2006}
697: %************
698: \eqb
699: \label{Prescorr}
700: \frac{\Delta P_{\tiny\mbox{2-loop}}}{P_{\tiny\mbox{1-loop}}}=-4.39\times
701: 10^{-4}\,,
702: \eqe
703: %************
704: where $\Delta P_{\tiny\mbox{2-loop}}\equiv
705: P_{\tiny\mbox{2-loop}}-P_{\tiny\mbox{1-loop}}$.
706: Since for $T\gg T_c$ we have \cite{Hofmann2005}
707: %************
708: \eqb
709: \label{Pres1loop}
710: P_{\tiny\mbox{1-loop}}=\frac{4}{45}\pi^2\,T^4
711: \eqe
712: %************
713: the correction $\Delta P_{\tiny\mbox{2-loop}}$ is also proportional to
714: $T^4$. Now $\Delta P_{\tiny\mbox{2-loop}}$ and the associated energy density
715: $\Delta\rho_{\tiny\mbox{2-loop}}$, which both are negative, must obey
716: the following relation\footnote{Legendre transformations are linear and thus
717: hold for each order in the loop expansion separately when integrating out
718: residual quantum fluctuations.}
719: %************
720: \eqb
721: \label{enecorr}
722: \Delta\rho_{\tiny\mbox{2-loop}}=T\frac{d\Delta
723: P_{\tiny\mbox{2-loop}}}{dT}-\Delta P_{\tiny\mbox{2-loop}}\,.
724: \eqe
725: %************
726: Eq.\,(\ref{enecorr}) and the fact that $\Delta
727: P_{\tiny\mbox{2-loop}}\propto T^4$ imply an equation of state
728: %************
729: \eqb
730: \label{eos}
731: \Delta\rho_{\tiny\mbox{2-loop}}=3\,\Delta P_{\tiny\mbox{2-loop}}\,,
732: \eqe
733: %************
734: and Eq.\,(\ref{Prescorr}) tells us that the thermal
735: energy density $\Delta\rho_{M+A}$, attributed to the
736: presence of unresolved, screened, isolated pairs of (no
737: longer BPS saturated) magnetic monopoles and
738: antimonopoles is given as\footnote{Since the process of (anti)caloron dissociation, creating pairs of isolated and screened magnetic
739: (anti)monopoles subject to an exact, overall charge neutrality, is irreversible the density of these objects is sharply fixed
740: at a given temperature. Thus there is no chemical potential associated with monopole-antimonopole pairs.}
741: %**********
742: \eab
743: \label{monpairdensenrg}
744: \frac{\Delta\rho_{M+A}}{T^4}&\equiv&\frac{1}{2\pi^2} \int_{\mu}^\infty
745: dy\,\sqrt{y^2-\mu^2}\,y^2\,n_B(y)\stackrel{!}=-\frac{\Delta\rho_{\tiny\mbox{2-loop}}}{T^4}\nonumber\\
746: &=&4.39\times 10^{-4}\times \frac{4}{15}\pi^2\,,
747: \eae
748: %*********
749: where $\mu\equiv m/T$ and $n_B(y)\equiv 1/(\exp(y)-1)$. The only
750: positive solution $\mu$ of Eq.\,(\ref{monpairdensenrg}) is
751: numerically given as $\mu=10.1224$. Thus we have
752: %************
753: \eqb
754: \label{ratiomassesnum}
755: \frac{m}{m_{>|\phi|^{-1}}}=1.139\,.
756: \eqe
757: %************
758: The effect of all other, screened, isolated, and stable
759: (anti)monopoles, generated by the dissociation of large-holonomy
760: (anti)calorons, hence is a lift of the mass of the system by about
761: 14\% above the BPS bound even though the hypothetic magnetic
762: screening length $l_s$ at high $T$ is larger than the mean
763: interparticle distance $\bar{d}$, see Sec.\,\ref{dens}.
764:
765: \subsection{Monopole density, (anti)monopole distance, and screening length\label{dens}}
766:
767: Knowing that $\mu=10.1224$ we now can
768: compute the hypothetic monopole density $n$ as
769: %******
770: \eqb
771: \label{density}
772: n=\frac{T^3}{2\pi^2}\int_\mu^\infty
773: dy\,y\sqrt{y^2-\mu^2}\,n_B(y)=9.799\times 10^{-5}\,T^3\,.
774: \eqe
775: %********
776: For reasons of symmetry there is a {\sl universal} mean monopole-antimonopole
777: distance $\bar{d}$. That is, the distance between a monopole and its
778: antimonopole, both stemming from the dissociation of the same
779: (anti)caloron, is, on average, the same as the distance between a monopole
780: adjacent to an antimonopole who did not originate from the same (anti)caloron.
781: From Eq.\,(\ref{density}) we have
782: %******
783: \eqb
784: \label{distance}
785: \bar{d}=n^{-1/3}=21.691\,\beta\,.
786: \eqe
787: %******
788: Thus the constant $c$ in Eq.\,(\ref{ratiolength}) is given as
789: $c=21.691$, and the right-hand side of (\ref{ratiolength}) is smaller
790: than unity for $T>2\,T_c$. That is, isolated and screened
791: (anti)monopoles are {\sl not resolved} at high temperatures in the
792: effective theory. Because the magnetic flux of a given pair essentially
793: cancels due to the large overlap of magnetic potentials no area law is
794: to take place for the spatial Wilson
795: loop, see Eq.\,(\ref{ratioscdis}).
796: This is in accord with our investigation performed in Sec.\,\ref{SWLE}.
797:
798: By virtue of Eqs.\,(\ref{magscl}) and (\ref{density}) the hypothetic magnetic
799: screening length $l_s$ is given as
800: %**********
801: \eqb
802: \label{magscl}
803: l_s=\frac{1}{g}\sqrt{\frac{T}{n}}=\frac{e}{4\pi}\,(21.691)^{3/2}\,\beta=71.43\,\beta\,.
804: \eqe
805: %***********
806: Thus we have
807: %**********
808: \eqb
809: \label{ratioscdis}
810: \frac{l_s}{\bar{d}}=3.293\,.
811: \eqe
812: %*********
813: Eq.\,(\ref{ratioscdis}) tells us that monopoles and
814: antimonopoles are not far separated on the scale of their screening
815: length. Notice that both of these length scales are much
816: smaller than $|\phi|^{-1}$. However, the impact of all other stable and screened
817: monopoles and antimonopoles on a given stable
818: pair is small due to efficient cancellations of mutual
819: attraction or repulsion. This is expressed by the small lift of the BPS
820: mass $m_{>|\phi|^{-1}}$, compare with Eq.\,(\ref{ratiomassesnum}).
821:
822:
823: \section{Spatial Wilson loop in effective variables\label{SWLE}}
824:
825: The effective theory for deconfining SU(2) Yang-Mills theory and the
826: computation of radiative corrections to the pressure
827: is described at length in \cite{Hofmann2005,Hofmann2007,SHG2006}. We only
828: name those results explicitely that are directly needed and for the
829: remainder refer the reader to these
830: papers.
831:
832: \subsection{Magnetic screening mass\label{screeningmass}}
833:
834: The SU(2) gauge symmetry of the fundamental action is
835: dynamically broken to U(1) by coarse-grained (anti)calorons. Notice that the order parameter
836: $\phi$ of this gauge-symmetry breaking is determined in a highly
837: nonlocal way \cite{Hofmann2007,HerbstHofmann2004} thus appealing to the strong magnetic-magnetic
838: correlations mediated by the $|Q|=1$ (anti)caloron configuration of trivial
839: holonomy \cite{HS1977}.
840:
841: In unitary-Coulomb gauge there are in the effective theory on
842: tree-level one massless vector excitation (hencefore referred to as $\gamma$)
843: and two thermal vector quasiparticle excitations (hencefore referred to as
844: $V^\pm$). The transversal $\gamma$ polarization
845: tensor $\Pi^{\mu\nu}$ is decomposed as
846: %*********
847: \eqb
848: \label{Pidec}
849: \Pi^{\mu\nu}=G(p_0,\vec{p})\,P^{\mu\nu}_T+F(p_0,\vec{p})\,P^{\mu\nu}_L
850: \eqe
851: %*********
852: where
853: %*********
854: \eqb
855: \label{PL}
856: P^{\mu\nu}_L\equiv \frac{p^\mu p^\nu}{p^2}-g^{\mu\nu}-P^{\mu\nu}_T\,,
857: \eqe
858: %*********
859: and
860: %*********
861: \eab
862: \label{transproj}
863: P^{00}_T&=&P^{0i}_T=P^{i0}_T=0\,,\nonumber\\
864: P^{00}_T&=&\delta^{ij}-p^ip^j/\vec{p}^2\,.
865: \eae
866: %*********
867: The functions $G(p_0,\vec{p})$ and $F(p_0,\vec{p})$ determine the propagation of
868: the interacting $\gamma$ mode. For $\mu=\nu=0$ Eq.\,(\ref{Pidec}) yields upon rotation
869: to real-time
870: %********
871: \eqb
872: \label{P00F}
873: F(p_0,\vec{p})=\left(1+\frac{p_0^2}{p^2}\right)^{-1}\,\Pi^{00}\,.
874: \eqe
875: %********
876: The function $F(p_0,\vec{p})$ measures the screening of electric
877: fields which we are not interested in when discussing the {\sl spatial} Wilson
878: loop. For propagation of the $\gamma$ mode into the 3-direction we have
879: %********
880: \eqb
881: \label{gxxgyy}
882: \Pi_{11}=\Pi_{22}=G(p_0,\vec{p})\,.
883: \eqe
884: %********
885: From the magnetic screening function $G(p_0,\vec{p})$ we obtain a definition for the magnetic screening $m_s$ as
886: %********
887: \eqb
888: \label{ms}
889: m_s\equiv \lim_{\vec{p}\to 0}\sqrt{\mbox{Re}\,G(p_0=0,\vec{p})}\,.
890: \eqe
891: %********
892: In Eqs.\,(\ref{gxxgyy}) and (\ref{ms}) we have suppressed the dependence on temperature
893: of the magnetic screening function $G$. The radiatively generated mass scale $m_s$ measures the exponential
894: decay rate of an externally applied, homogeneous magnetic field
895: $B_i=\epsilon_{ijk} G^3_{jk}$ penetrating the
896: Yang-Mills plasma. In the effective theory,
897: the screening function $G$ only has support
898: for the four-momentum $p$ satisfying the constraint
899: %****
900: \eqb
901: \label{constr}
902: |p^2-G(p^0,\vec{p})|\le |\phi|^2\,,
903: \eqe
904: %****
905: where $|\phi|$ is the (temperature-dependent) modulus of the
906: inert, adjoint scalar field $\phi$ emerging as a consequence of
907: spatial coarse-graining over the stable BPS sector with topological
908: charge modulus $|Q|=1$. Eq.\,(\ref{constr}) is the condition that a propagating effective tree-level massless gauge mode in physical Coulomb-unitary gauge must not be further off its radiatively induced mass shell
909: then the scale $|\phi|$ of maximal resolution.
910:
911: In Fig.\,\ref{Fig-1} the two diagrams a priori
912: contributing to $\Pi^{\mu\nu}$ are depicted.
913: %***********************
914: \begin{figure}
915: \begin{center}
916: \leavevmode
917: %\epsfxsize=9.cm
918: \leavevmode
919: %\epsffile[80 25 534 344]{}
920: \vspace{4.9cm}
921: \special{psfile=Fig-1.ps angle=0 voffset=-160
922: hoffset=-195 hscale=90 vscale=60}
923: \end{center}
924: \caption{\protect{\label{Fig-1}} The diagrams for $\gamma$'s
925: polarization tensor $\Pi_{\mu\nu}$.}
926: \end{figure}
927: %************************
928: Using standard Feynamn rules, we have for diagram A
929: \begin{equation}
930: \label{Anproc}
931: \begin{split}
932: \Pi^{\mu\nu}_{A}(p)\,=\,&
933: \frac{1}{2i}\int\frac{d^4k}{(2\pi)^4} e^2
934: \epsilon_{ace}[g^{\mu\rho}(-p-k)^\lambda+g^{\rho\lambda}(k-p+k)^\mu+g^{\lambda\mu}(p-k+p)^\rho]\times\\
935: & \epsilon_{dbf}[g^{\sigma\nu}(-k-p)^\kappa+g^{\nu\kappa}(p+p-k)^\sigma+g^{\kappa\sigma}(-p+k+k)^\nu]\times\\
936: &(-\delta_{cd})\left(g_{\rho\sigma}-\frac{k_\rho k_\sigma}{m^2}\right)
937: \left[\frac{i}{k^2-m^2}+2\pi\delta(k^2-m^2)\,n_B(|k_0|/T) \right]\times\\
938: &(-\delta_{ef})\left(g_{\lambda\kappa}-\frac{(p-k)_\lambda(p-k)_\kappa}{(p-k)^2}\right)\times\\
939: &\left[\frac{i}{(p-k)^2-m^2}+2\pi\delta((p-k)^2-m^2)\,n_B(|p_0-k_0|/T) \right]\,.
940: \end{split}
941: \end{equation}
942: From the one-loop evolution \cite{Hofmann2005} we know that $e\ge
943: \sqrt{8\pi}$ \cite{Hofmann2007}. Due to constraint $|k^2-m^2|\le
944: |\phi|^2$ \cite{Hofmann2005}, where $m=2e\,|\phi|$, the vacuum part in the $V^\pm$ propagator is
945: forbidden. For an explicit presentation of the $\gamma$ and $V^\pm$
946: real-time propagators see Eqs.\,(\ref{Vpm}) and (\ref{gamma}).
947:
948: Diagram B
949: reads
950: \begin{equation}
951: \label{vactad}
952: \begin{split}
953: \Pi^{\mu\nu}_{B}(p)\,=\,&\frac{1}{i} \int \frac{d^4k}{(2\pi)^4}
954: (-\delta_{ab}) \left( g_{\rho\sigma}-\frac{k_\rho k_\sigma}{m^2} \right)
955: \left[\frac{i}{k^2-m^2}+2\pi\delta(k^2-m^2)n_B(|k_0|/T) \right]\times\\
956: & \quad (-ie^2)[
957: \epsilon_{abe}\epsilon_{cde}(g^{\mu\rho}g^{\nu\sigma}-g^{\mu\sigma}g^{\nu\rho})
958: +\epsilon_{ace}\epsilon_{bde}(g^{\mu\nu}g^{\rho\sigma}-g^{\mu\sigma}g^{\nu\rho})+\\
959: &\quad\epsilon_{ade}\epsilon_{bce}(g^{\mu\nu}g^{\rho\sigma}-g^{\mu\rho}g^{\nu\sigma})]\,.
960: \end{split}
961: \end{equation}
962: Again, the part in Eq.\,(\ref{vactad}) arising from the vacuum contribution in
963: Eq.\,(\ref{Vpm}) vanishes. For the four-vertex in diagram B the
964: following constraint holds \cite{Hofmann2005}: $|(p+k)^2|\le
965: |\phi|^2$.
966:
967: It is easily seen that only diagram B contributes to
968: $m_s$. Generalizing $m_s\to m_s(p_0=0,|\vec{p}|)$ within the finite
969: support in $\vec{p}$ given by the condition (\ref{constr}) now
970: specializing to $|\vec{p}^2+G(p^0=0,\vec{p})|\le |\phi|^2$, we obtain a behavior as
971: depicted in Fig.\,\ref{G} in dependence of
972: $X=|\vec{p}|/T$.
973: %******************
974: \begin{figure}
975: \begin{center}
976: %\leavevmode
977: %%\epsfxsize=9.cm
978: %\leavevmode
979: %%\epsffile[80 25 534 344]{}
980: \vspace{5.8cm}
981: \special{psfile=Fig-G.ps angle=0 voffset=-5
982: hoffset=40 hscale=50 vscale=44}
983: \end{center}
984: \caption{\protect{\label{G}}The function $\mbox{Re}\,G/T^2$ in dependence of
985: $X=|\vec{p}|/T$ setting $p_0=0$ for $T=2\,T_c$ (black) and $T=3\,T_c$ (gray).}
986: \end{figure}
987: %******************
988: From Fig.\,\ref{G} and comparing with Eq.\,(\ref{ms}) one sees
989: that $m_s=0$ which is in agreement with our
990: microscopic analysis of Sec.\,\ref{MP}. Recall that due to the
991: unresolvability of stable and screened monopole-antimopole
992: pairs no net monopole or antimonopole density\footnote{In contrast to
993: Sec.\,\ref{dens} we are here concerned with the monopole density
994: as seen in our {\sl effective} theory at
995: resolution $|\phi|$ and not with the
996: hypothetic monopole density detected by a higher resolution.}
997: is detected and thus, according
998: to Eq.\,(\ref{magscl}), $l_s\equiv 1/m_s=\infty$.
999:
1000:
1001: \subsection{Generalities on spatial Wilson loop in effective theory\label{GSWE}}
1002:
1003: To compute a Wilson loop in effective
1004: variables associated with a given, finite resolution $\mu$ is in general something different than the definition in fundamental variables at infinite resolution demands.
1005: However, at finite temperature in the deconfining phase of a
1006: Yang-Mills theory there is only one resolution scale where the effective action minimally differs from the fundamental action: at $\mu=|\phi|$ technically topologically trivial and nontrivial field configurations
1007: are separated modulo mass generation and constraints on the hardness of quantum fluctuations of the former imposed by the latter. Since an ensemble average over
1008: isolated and screened magnetic monopoles is generated by the radiative corrections
1009: in the effective theory it is suggestive that the spatial Wilson loop evaluated on propagating gauge mode conveying the magnetic flux of the effective theory, with a resummation of the lowest-order radiative effects (polarization tensor) indeed, measures the flux sourced by the ensemble of
1010: magnetic monopoles determining the polarization tensor. This argument is purely intuitive,
1011: and no proof of this suggested property is yet available.
1012:
1013: Here we quote some general results on the calculation
1014: of exponentiated one-effective-gauge-boson exchanges within the spatial
1015: quadratic contour $C$ of side-length $L$.
1016:
1017: In order to calculate the spatial string tension, we use an expansion
1018: into loops for the $N$-point functions in the effective theory \cite{Bassetto}. Because of the rapid
1019: numerical convergence of the effective loop expansion
1020: \cite{SHG2006,HofmannLE2006,KH2007} we are content here
1021: with a resummation of the one-loop polarization tensor for the $\gamma$
1022: mode into its effective propagator and the tree-level propagator for the
1023: $V^\pm$ modes (2-point functions). The exchange of these
1024: modes is subsequently exponentiated in order to take into
1025: account higher $N$-point functions in a trivial way, see Fig.\,\ref{Wilson}.
1026:
1027: Thus we can write for the logarithm of $W[C]$
1028: \begin{equation}
1029: \ln W[C]=-\frac{1}{2}C_F \oint\!dx_\mu dy_\nu\,D_{\mu\nu}(x-y)\,,
1030: \end{equation}
1031: where $C_F$ denotes the Dynkin index in the fundamental representation,
1032: defined as the normalization factor of the generators of the algebra (i.e. $C_F\equiv\frac{1}{2}$), and
1033: \begin{equation}
1034: D_{\mu\nu}=\sum\limits_{a=1}^3 D_{\mu\nu}^{(a)}
1035: \end{equation}
1036: is the sum of the tree-level propagators ($V^\pm$ corresponds to $a=1,2$) and the one-loop
1037: resummed propagator ($\gamma$ corresponds to $a=3$).
1038:
1039: The tree-level propagators for $a=1,2$ in
1040: position space are given as the Fourier transformations of their momentum space counterparts
1041: \begin{equation}
1042: \label{Vpm}
1043: D_{\mu\nu}^{1,2}(\beta,x-y)
1044: =-\int\!\frac{d^4p}{(2\pi)^4}\,\e^{-ip(x-y)}\left(g_{\mu\nu}-\frac{p_\mu p_\nu}{m^2}
1045: \right)\left[\frac{i}{p^2-m^2+i\varepsilon}
1046: +2\pi\delta(p^2-m^2)n_B(\beta|p_0|)\right]\,.
1047: \end{equation}
1048: The one-loop resummation of the magnetic part\footnote{We are not
1049: interested in electric screening effects since we study the {\sl
1050: spatial} Wilson loop.} of the polarization
1051: tensor (screening function $G$) into the dressed
1052: $\gamma$ propagator, see
1053: Eq.\,\ref{Pidec}, yields the following result
1054: \begin{equation}
1055: \label{gamma}
1056: D_{\mu\nu}^{3}(\beta,x-y)
1057: =\int\!\frac{d^4p}{(2\pi)^4}\,\e^{-ip(x-y)}
1058: \left[P^T_{\mu\nu}\left(\frac{i}{p^2-G(p^0,\vec{p})+i\varepsilon}
1059: +2\pi\delta(p^2-G(p^0,\vec{p}))n_B(\beta|p_0|)\right)-i\frac{u_\mu u_\nu}{\vec{p}^2}\right]\,,
1060: \end{equation}
1061: where the transversal projection operator is given as
1062: \begin{eqnarray}
1063: P^T_{00}(p)&=&P^T_{0i}(p)=P^T_{i0}(p)=0\nonumber\\
1064: P^T_{ij}(p)&=&\delta_{ij}-\frac{p_i p_j}{\vec{p}^2}.
1065: \end{eqnarray}
1066: We calculate the contour integral in the 1-2-plane, that is $x_0=y_0=x_3=y_3=0$,
1067: and consider at first an arbitrary propagator $D_{\mu\nu}(p)$.
1068: Later we will insert the explicit expressions for $D_{\mu\nu}^{1,2}$
1069: ($V^\pm$ gauge modes) and for $D_{\mu\nu}^{3}$ ($\gamma$ mode).
1070: We obtain \cite{KellerDA}
1071: \begin{eqnarray}
1072: \label{konturintegral}
1073: \ln W[C]&=&-\frac{1}{4}\int \frac{d^4p}{(2\pi)^4}\oint\oint dx_\mu dy_\nu
1074: D_{\mu\nu}(p)\,\e^{-ip(x-y)}\left.\right|_{x_0=y_0=x_3=y_3=0}\nonumber\\
1075: &=&-4\int \frac{d^4p}{(2\pi)^4}\,\sin^2\left(\frac{p_1L}{2}\right)\sin^2\left(\frac{p_2L}{2}\right)
1076: \left(\frac{D_{11}}{p_1^2}-\frac{D_{12}}{p_1p_2}
1077: -\frac{D_{21}}{p_1p_2}+\frac{D_{22}}{p_2^2}\right)\,.
1078: \end{eqnarray}
1079: %******************
1080: \begin{figure}
1081: \begin{center}
1082: %\leavevmode
1083: %%\epsfxsize=9.cm
1084: %\leavevmode
1085: %%\epsffile[80 25 534 344]{}
1086: \vspace{5.3cm}
1087: \special{psfile=Fig-Wilson.ps angle=0 voffset=-5
1088: hoffset=55 hscale=50 vscale=40}
1089: \end{center}
1090: \caption{\protect{\label{Wilson}Illustration of the diagrammatic
1091: approach.
1092: Every summand in this formula represents a whole
1093: class of $N$-gauge-boson exchange diagrams, which is encoded in terms of a
1094: (suppressed)
1095: combinatoric factor.}}
1096: \end{figure}
1097: %******************
1098:
1099: \subsection{Thermal part due to massive modes\label{massive}}
1100:
1101: Here we present the result for the contribution of the (thermal part of)
1102: the $V^\pm$ propagators to the logarithm of the spatial Wilson
1103: loop. Inserting the thermal part of the $V^\pm$ propagator in
1104: Eq.\,(\ref{Vpm}) into Eq.\,(\ref{konturintegral}, we have
1105: %*********
1106: \eqb
1107: \label{VpmWL}
1108: \ln W[C]_{V^\pm}=\frac{1}{\pi^3}\int\!d^3p\,
1109: \frac{\sin^2\left(\frac{p_1L}{2}\right)\sin^2\left(\frac{p_2L}{2}\right)}
1110: {\sqrt{p_1^2+p_2^2+p_3^2+m^2}}
1111: n_B(\beta\sqrt{p_1^2+p_2^2+p_3^2+m^2})\left(\frac{1}{p_1^2}+\frac{1}{p_2^2}\right)\,.
1112: \eqe
1113: %********
1114: Let us now rescale the momenta $p_i$
1115: and the squared $V^\pm$ mass $m^2$ in Eq.\,(\ref{VpmWL})
1116: to dimensionless variables as follows
1117: \begin{equation}
1118: \hat{p_i}=p_i\cdot L\,,
1119: \end{equation}
1120: and using $e=\sqrt{8}\pi$,
1121: \begin{equation}
1122: \hat{m}^2=\frac{m^2}{T^2}=\frac{(2e)^2}{T^2}\frac{\Lambda}{2\pi T}=\frac{128\pi^4}{\lambda^3}\,,
1123: \end{equation}
1124: where $L$ is the side-length of the spatial quadratic contour $C$.
1125: To eventually perform the limit $L\rightarrow\infty$, we introduce the
1126: dimensionless parameter $\tau$ as
1127: \begin{equation}
1128: \label{tau}
1129: \tau=T\cdot L\,.
1130: \end{equation}
1131: Eq.\,(\ref{VpmWL}) is then recast as
1132: %*********
1133: \eqb
1134: \label{VpmWLdl}
1135: \ln W[C]_{V^\pm}=\frac{1}{\pi^3}\int\!d\hat{p}_1 d\hat{p}_2 d\hat{p}_3\,
1136: \frac{\sin^2\left(\frac{\hat{p}_1}{2}\right)\sin^2\left(\frac{\hat{p}_2}{2}\right)}
1137: {\sqrt{\hat{p}_1^2+\hat{p}_2^2+\hat{p}_3^2+\frac{128\pi^4}{\lambda^3}\tau^2}}\,
1138: n_B\left(\frac{\sqrt{\hat{p}_1^2+\hat{p}_2^2+\hat{p}_3^2
1139: +\frac{128\pi^4}{\lambda^3}\tau^2}}{\tau}\right)
1140: \left(\frac{1}{\hat{p}_1^2}+\frac{1}{\hat{p}_2^2}\right)\,.
1141: \eqe
1142: %*********
1143: From Eq.\,(\ref{VpmWLdl}) it is obvious that in the limit
1144: $\tau\to\infty$ the contribution to $\frac{\ln W[C]}{\tau^2}$ of the $V^\pm$ modes is
1145: nil, and we no longer need to discuss their potential impact on the spatial string tension.
1146:
1147: \subsection{Thermal part due to massless mode\label{thermalpart}}
1148:
1149: Let us now discuss the contribution of the thermal part
1150: of the $\gamma$ mode to $\ln W[C]$. In \cite{LH2008} we have computed the
1151: screening function $G$ selfconsistently on the radiatively induced mass
1152: shell $p^2-G(p^0,\vec{p})=0$. On this mass-shell, only diagram $B$ in
1153: Fig.\,\ref{Fig-1} contributes. In Fig.\,\ref{Fig-3} the dependence
1154: of $\log\left|\frac{G}{T^2}\right|$ on the dimensionless spatial
1155: momentum modulus $X\equiv\frac{|\vec{p}|}{T}$ (left-panel) and on
1156: dimensionless frequency $Y=\sqrt{X^2+\frac{G}{T^2}}$ (right-panel)
1157: is depicted for various
1158: temperatures.
1159: %***********************
1160: \begin{figure}
1161: \begin{center}
1162: \leavevmode
1163: %\epsfxsize=9.cm
1164: \leavevmode
1165: %\epsffile[80 25 534 344]{}
1166: \vspace{7.2cm}
1167: \special{psfile=Fig-3bi.ps angle=0 voffset=-350
1168: hoffset=-170 hscale=50 vscale=65}
1169: \end{center}
1170: \caption{\protect{\label{Fig-3}} Plots of $\log\left|\frac{G}{T^2}\right|$ in the full calculation
1171: (solid grey curves) and for the approximation $p^2=0$ (dashed grey
1172: curves). The cusps in $\log\left|\frac{G}{T^2}\right|$ correspond to
1173: zeros separating the regime of screening ($G>0$) from the regime of
1174: antiscreening ($G<0$). The left panel depicts
1175: $\log\left|\frac{G}{T^2}\right|$ as a function of $X$.
1176: The right panel shows $\log\left|\frac{G}{T^2}\right|$ as a function of $Y\equiv\sqrt{X^2+\frac{G}{T^2}}$.
1177: Here the dashed black curve is the function $2\log Y$. In order of
1178: increasing lightness the curves correspond to $T=2\,T_c$, $T=3\,T_c$, and $T=4\,T_c$.}
1179: \end{figure}
1180: %***********************
1181: Inserting the thermal part of the magnetically dressed $\gamma$
1182: propagator of Eq.\,(\ref{gamma}) into Eq.\,(\ref{konturintegral}), we obtain
1183: %***********************
1184: \eqb
1185: \label{Wilsongamma}
1186: \ln W[C]_\gamma^{\tiny\mbox{th}}=-\frac{1}{2\pi^3}\int\!d\hat{p}_1 d\hat{p}_2 d\hat{p}_3\,
1187: \frac{\sin^2\left(\frac{\hat{p}_1}{2}\right)\sin^2\left(\frac{\hat{p}_2}{2}\right)}
1188: {\sqrt{\hat{p}_1^2+\hat{p}_2^2+\hat{p}_3^2+\hat{G}\tau^2}}\,n_B\left(\frac{\sqrt{\hat{p}_1^2+\hat{p}_2^2+\hat{p}_3^2+\hat{G}\tau^2}}{\tau}\right)
1189: \left(\frac{1}{\hat{p}_1^2}+\frac{1}{\hat{p}_2^2}\right)\,,
1190: \eqe
1191: %************
1192: where $\hat{G}\equiv\frac{G}{T^2}$. In Fig.\,\ref{Fig-4} plots of
1193: $-\ln W[C]_\gamma^{\tiny\mbox{th}}$ for $T=2\,T_c, 3\,T_c,$ and $4\,T_c$
1194: are shown as functions of $\tau$.
1195: %***********************
1196: \begin{figure}
1197: \begin{center}
1198: \leavevmode
1199: %\epsfxsize=9.cm
1200: \leavevmode
1201: %\epsffile[80 25 534 344]{}
1202: \vspace{6.8cm}
1203: \special{psfile=Fig-4i.ps angle=0 voffset=-200
1204: hoffset=-230 hscale=60 vscale=50}
1205: \end{center}
1206: \caption{\protect{\label{Fig-4}} Plots of
1207: $-\ln W[C]_\gamma^{\tiny\mbox{th}}$ as a function of $\tau=T\cdot L$. The dotted
1208: lines correspond to the value of $L$ coinciding with the minimal length
1209: scale $|\phi|^{-1}$ in the effective theory at a given temperature.
1210: The left line is for $T=2\,T_c$, the right line for $T=3\,T_c$, and the
1211: line for $T=4\,T_c$ would be at $\tau^*[4\,T_c]=65.77$ and thus is not
1212: contained in the figure.}
1213: \end{figure}
1214: %***********************
1215: Clearly, for each temperature and for large $\tau$ we observe a {\sl perimeter}
1216: law in the approximation used\footnote{Again, we believe that this
1217: approximation captures all the essential physics due to the rapid
1218: convergence in the number of external legs in $N$-point
1219: functions. Notice that in a one-loop diagram making up a radiative correction to $\gamma$'s
1220: $N$-point function there are $N-1$ independent external four momenta
1221: $p_i$ all of which are subject to the constraint
1222: $|p_i^2|\le|\phi|^2$. With increasing $N$ this should rapidly suppress the contribution of
1223: one-loop corrections to $\gamma$'s $N$-point function. Moreover, the
1224: expansion into the number $M$ of loops at a given $N$ does converge
1225: rapidly \cite{HofmannLE2006,SHG2006,KH2007}.}. For $\tau$ considerably
1226: below $\tau^*$, which is associated with the maximal resolution
1227: $|\phi|$ (we set $L=|\phi|^{-1}$ in Eq.\,(\ref{tau})), we do, however, observe curvature in $-\ln W[C]_\gamma^{\tiny\mbox{th}}$.
1228:
1229: \subsection{Quantum part due to massless mode\label{vacuumpart}}
1230:
1231: We now turn to the contribution to $\ln W[C]_\gamma$ of the quantum part of $\gamma$'s
1232: dressed propagator. Inserting the quantum part of the magnetically dressed $\gamma$
1233: propagator of Eq.\,(\ref{gamma}) into Eq.\,(\ref{konturintegral}), we obtain
1234: %***********************
1235: \eqb
1236: \label{Wilsongamma}
1237: \ln W[C]_\gamma^{\tiny\mbox{vac}}=-\frac{i}{4\pi^4}\int\!d^4\hat{p}\,
1238: \sin^2\left(\frac{\hat{p}_1}{2}\right)\sin^2\left(\frac{\hat{p}_2}{2}\right)
1239: \frac{1}{\hat{p}^2-\hat{G}\tau^2+i\epsilon}
1240: \left(\frac{1}{\hat{p}_1^2}+\frac{1}{\hat{p}_2^2}\right)\,,
1241: \eqe
1242: %************
1243: where $\hat{G}\equiv\frac{G}{T^2}$ and $\hat{p}_0\equiv p_0\,L$.
1244: Notice that in this case the screening function $G$ receives
1245: contributions from both diagrams A and B in Fig.\,\ref{Fig-1}
1246: because the only constraint on $\gamma$'s momentum is
1247: $|p^2-\mbox{Re}\,G(p^0,\vec{p})|\le|\phi|^2$.
1248: In Fig.\,\ref{Fig-5} the imaginary part of $G$ (due to diagram A) is
1249: plotted as a function of $X_0\equiv\frac{p_0}{T}$ and of
1250: $X\equiv\frac{\vec{p}}{T}$ at $T=2\,T_c$. Clearly, $\mbox{Im}\,G$ is nonvanishing only
1251: within a small region centered at the origin of the $X_0-X$ plane.
1252: %***********************
1253: \begin{figure}
1254: \begin{center}
1255: \leavevmode
1256: %\epsfxsize=9.cm
1257: \leavevmode
1258: %\epsffile[80 25 534 344]{}
1259: \vspace{6.8cm}
1260: \special{psfile=Fig-5P.ps angle=0 voffset=-200
1261: hoffset=-220 hscale=55 vscale=45}
1262: \end{center}
1263: \caption{\protect{\label{Fig-5}}$\mbox{Im}\,G/T^2$, as orginating from
1264: diagram A in Fig.\,\ref{Fig-1}, as a function of
1265: $X_0\equiv\frac{p_0}{T}$ and of
1266: $X\equiv\frac{\vec{p}}{T}$ at $T=2\,T_c$}
1267: \end{figure}
1268: %***********************
1269: Fig.\,\ref{Fig-5} suggests that the modulus of the factor
1270: $\frac{1}{\hat{p}^2-\hat{G}\tau^2+i\epsilon}$ in
1271: Eq.\,(\ref{Wilsongamma}) can be estimated by the situation of
1272: unadulterated $\gamma$ propagation. That is, we approximately may evaluate the
1273: integral in Eq.\,(\ref{Wilsongamma}) by setting $G=0$. To do this, one
1274: may imagine the condition $|p^2|\le|\phi|^2$ to be implemented in such a
1275: way that a strongly-decaying analytic factor (for $|p^2|>|\phi|^2$),
1276: is introduced to make the use of the theorem of residues
1277: applicable. One then obtains
1278: %***********************
1279: \eqb
1280: \label{Wilsongammafree}
1281: \ln W[C]_\gamma^{\tiny\mbox{vac}}=\frac{1}{2\pi^3}\int\!d^3\hat{p}\,
1282: \sin^2\left(\frac{\hat{p}_1}{2}\right)\sin^2\left(\frac{\hat{p}_2}{2}\right)
1283: \frac{1}{\sqrt{\hat{p}_1^2+\hat{p}_2^2+\hat{p}_3^2}}
1284: \left(\frac{1}{\hat{p}_1^2}+\frac{1}{\hat{p}_2^2}\right)\,.
1285: \eqe
1286: %************
1287: The integral in Eq.\,(\ref{Wilsongammafree}) is UV divergent.
1288: Introducing a UV cutoff, one sees that the part of the spectrum, where $G>0$,
1289: contributes a real monotonic-decreasing-in-$\tau$ function whereas for $G<0$ and sufficiently
1290: large $\tau$ the contribution is imaginary with modulus that is also a
1291: monotonic-decreasing-in-$\tau$ function. Furthemore the UV
1292: divergence itself does not depend on $\tau$. Thus we may refrain from considering this quantum contribution
1293: any further.
1294:
1295: \subsection{Summary of results obtained in effective
1296: variables\label{sev}}
1297:
1298: To summarize, the only nontrivial contribution to $\ln W[C]_\gamma$
1299: arises from the thermal part of the resummed $\gamma$-propagator as
1300: investigated in Sec.\,\ref{thermalpart}. There we observe that
1301: for hypothetic values of $L$ smaller than the minimal length
1302: $|\phi|^{-1}$ in the effective theory an area law emerges. This is
1303: qualitatively in line with lattice investigations using the Wilson
1304: action at finite lattice spacing, see Sec.\,\ref{sum}, in the sense that
1305: an artificial resolution
1306: scale is introduced to probe the system at small spatial
1307: distances.
1308:
1309:
1310: \section{Conclusions\label{sum}}
1311:
1312: For the deconfining phase of SU(2) Yang-Mills thermodynamics we have argued
1313: that there is a unique effective action emerging when applying a combination of
1314: spatial-coarse graining and separation of BPS saturated field configurations from trivial-toplogy
1315: fluctuations. This programme essentially appeals to spatial homogeneity and isotropy of thermodynamical systems, the perturbative renormalizability of the theory \cite{tHooftVeltman}, and the fact that fundamental as well as effective
1316: BPS configurations possess no energy-stress and thus do not propagate.
1317:
1318: Subsequently, we have investigated the physics of screened magnetic (anti)monopoles,
1319: as they emerge by the dissociation of large-holonomy (anti)calorons,
1320: both in a hypothetic setting assuming a larger resolution than the maximal resolution of the effective theory for the deconfining phase $|\phi|$ \cite{Hofmann2007} and in terms of the spatial Wilson loop
1321: computed in effective variables. While the former investigation is safe since radiative corrections \cite{HofmannLE2006,SHG2006} are under accurate control\footnote{There are kinematic constraints on the propagation of effective fields as imposed by the thermal ground state: Maximal offshellness and maximal momentum transfer in a local vertex. Using the Euler-L'Huillier characteristics
1322: for spherical polyhedra (including a nontrivial genus) one shows that increasing the number of vertices in irreducible loop diagrams the number of constraints on a priori noncompact integration variables rapidly exceeds
1323: the number of variables. This suggests a termination of the expansion in irreducible
1324: loop diagrams at a finite loop order. Example calculation for the pressure up to three
1325: loops \cite{SHG2006,KH2007} provide ample evidence for this conjecture.} in the effective theory, the latter approach assumes that the physical content of the definition of the Wilson loop in fundamental field variables, which is modulo certain approximations (see below)
1326: employed in direct lattice computations, is the same as the definition in terms of thermal-ground-state influenced propagating, effective field variables. Although it is suggestive that this is true, we so far have no proof for this correspondence. So this question remains open. At the same time, however, the computation of the mean density of screened and stable monopole-antimonopole pairs from an energy-density deficit introduced by a certain effective
1327: two-loop correction to the pressure \cite{SHG2006} as compared to the free quasiparticle situation and the subsequent calculation of the magnetic screening length implies that according to the interpretation of
1328: Ref.\,\cite{KH} no area law can arise as a consequence of the magnetic flux of a given pair decaying too rapidly.
1329: This is also what the computation of the effective Wilson loop indicates.
1330:
1331: Therefore, our results using these two alternative methods match in the sense that there is
1332: no area-law for the effective spatial Wilson loop at a fixed temperature and in
1333: the limit of large contour size although an area law emerges in the direct calculation of the Wilson loop in the effective theory when the contour size falls below the resolution $|\phi|$ which, of course, is an unphysical situation.
1334:
1335: In lattice simulations the typical spatial resolution -- the inverse
1336: spatial lattice spacing -- even at temperatures a few times $T_c$
1337: is considerably larger then the natural\footnote{By `natural' we mean
1338: that at maximal resolution $|\phi|$ the effective action for
1339: deconfining Yang-Mills thermodynamics is of the simple
1340: form of Eq.\,(\ref{effdec}).} resolution $|\phi|$ of continuum and
1341: infinite-volume Yang-Mills thermodynamics
1342: \cite{Polonyi1987,Bali1993}. Lattice simulations are usually
1343: performed with the Wilson action at {\sl finite} values of the lattice
1344: spacing $a$. However, due to the contribution of topologically nontrivial field
1345: configurations to the partition function a renormalization-group
1346: evolved perfect lattice action certainly is more
1347: complicated than the Wilson action obtained from the continuum Yang-Mills
1348: action by naive discretization. Using the Wilson action, the scaling
1349: regime for the fundamental coupling hardly makes any reference to bulk
1350: properties of the highly nonperturbative ground-state physics
1351: (trace anomaly \cite{GiacosaHofmann2007}). So what we have indirectly argued for in
1352: this work is that the physics of screened and isolated
1353: magnetic (anti)monopoles is very sensitive to a mild resolution
1354: dependence of the {\sl partition function} as it is artificially introduced by
1355: the Wilson action. In
1356: principle, this action should be modified by nonperturbative effects
1357: yielding the perfect lattice action. The latter, however, is
1358: extremely hard to generate at finite temperatures.
1359:
1360: Given this observation and the principle problem of comparison between fundamental Wilson loop
1361: and its effective counterpart it is not surprising that an area law for the
1362: spatial Wilson loop is measured in lattice simulations (in accord with
1363: a 3D strong-`coupling'-expansion argument \cite{Borgs1985}) subject to the
1364: Wilson action. Notice that the introduction of a finite spatial
1365: lattice spacing still working with the Wilson action actually acts
1366: physically (which it should not) in separating monopoles
1367: from antimonopoles. As a result, a net magnetic flux is measured through
1368: the spatial contour in lattice simulations although it is not clear in what sense the
1369: nonabelian flux through the Wilson loop defined in fundamental variables is related to the abelian flux that we investigate by the $\gamma$ mode's contribution to the effective Wilson loop. We stress that lattice
1370: simulations using the Wilson action are interesting and important because they strongly point to certain
1371: aspects of the highly nonperturbative ground-state physics. However, we suspect that they are not sufficiently adapted to describe the subtle effects attributed to large-holonomy
1372: (anti)caloron dissociation as they take place in infinite-volume
1373: continuum Yang-Mills thermodynamics. To turn this into a completely
1374: rigorous statement for the Wilson loop in effective variables the above-mentioned correspondence
1375: would have to be proved, and a precise estimate for the contribution to the integration over the
1376: spatial contour of $N$-point functions with
1377: higher internal loop number would have to be obtained. We leave this to future
1378: investigation.
1379:
1380:
1381: \section*{Acknowledgments}
1382: We would like to thank Markus Schwarz for very useful conversations and helpful comments on the
1383: manuscript. A Referee's helpful queries are gratefully acknowledged.
1384:
1385:
1386: \begin{thebibliography}{10}
1387:
1388: \bibitem{Borgs1985}
1389: C. Borgs, Nucl. Phys. B {\bf 261}, 455 (1985).
1390:
1391: \bibitem{Polonyi1987}
1392: E. Manousakis and J. Polonyi, Phys. Rev. Lett. {\bf 58}, 847 (1987).
1393:
1394: \bibitem{Bali1993}
1395: G. S. Bali et al., Phys. Rev. Lett. {\bf 71}, 3059 (1993).
1396:
1397: \bibitem{spatialWL}
1398: M. Laine and Y. Schr\"oder, JHEP {\bf 0503} 067 (2005) and references
1399: therein.
1400:
1401: \bibitem{Hofmann2005}
1402: R.~Hofmann, Int.\ J.\ Mod.\ Phys.\ A{\bf 20} (2005) 4123, Erratum-ibid.\ A {\bf 21} (2006) 6515.\\
1403: R.~Hofmann, Mod. Phys. Lett. A{\bf 21}, 999 (2006), Erratum-ibid. A {\bf
1404: 21}, 3049 (2006).
1405:
1406: \bibitem{HofmannGiacosa2006}
1407: F. Giacosa and R. Hofmann, Prog. Theor. Phys. {\bf 118}, 759 (2007)
1408: [hep-th/0609172].
1409:
1410: \bibitem{HofmannLE2006}
1411: R. Hofmann, hep-th/0609033.
1412:
1413: \bibitem{Hofmann2007}
1414: R. Hofmann, arXiv:0710.0962 [hep-th].
1415:
1416: \bibitem{HerbstHofmann2004}
1417: U. Herbst and R. Hofmann, hep-th/0411214.
1418:
1419: \bibitem{HS1977}
1420: B. J. Harrington and H. K. Shepard, Phys. Rev. D {\bf 17}, 2122 (1978).
1421:
1422: \bibitem{NahmVanBaalLeeLu}
1423: W. Nahm, Phys. Lett. B {\bf 90}, 413 (1980).\\
1424: W. Nahm, {\sl Selfdual Monopoloes and Calorons.} Lect. Notes in Physics. 201, eds. G. Denaro, e.a. (1984)
1425: p. 189, Springer Berlin-Heidelberg-New York.\\
1426: K.-M. Lee and C.-H. Lu, Phys. Rev. D {\bf 58}, 025011 (1998).\\
1427: T. C. Kraan and P. van Baal, Phys. Lett. B {\bf 428}, 268 (1998).\\
1428: T. C. Kraan and P. van Baal, Phys. Lett. B {\bf 435}, 389 (1998).
1429:
1430: \bibitem{GPY1981}
1431: D. J. Gross, R. D. Pisarski, and L. G. Yaffe, Rev. Mod. Phys. {\bf 53},
1432: 43 (1981).
1433:
1434: \bibitem{Kibble}
1435: T. W. B. Kibble, {\it J. Phys. A}, {\bf 9} 1387 (1976).
1436:
1437: \bibitem{SHG2006}
1438: M.~Schwarz, R.~Hofmann and F.~Giacosa, Int. J. Mod. Phys. A {\bf 22},
1439: 1213 (2007).
1440:
1441: \bibitem{KH2007}
1442: D. Kaviani and R. Hofmann, Mod. Phys. Lett. A {\bf 22}, 2343 (2007).
1443:
1444: \bibitem{KHG2007}
1445: J. Keller, R. Hofmann, and F. Giacosa, Int. J. Mod. Phys. A {\bf 23 }, 5181 (2008).
1446:
1447: \bibitem{tHooftVeltman}
1448: G. 't Hooft and M. J. G. Veltman, Nucl. Phys. B {\bf 44}, 189 (1972).\\
1449: G. 't Hooft, Nucl. Phys. B {\bf 33}, 173 (1971).\\
1450: G. 't Hooft, Nucl. Phys. B {\bf 62}, 444 (1973).\\
1451: G. 't Hooft and M. J. G. Veltman, Nucl. Phys. B {\bf 50}, 318 (1972).\\
1452: B. W. Lee and Jean Zinn-Justin, Phys. Rev. D {\bf 5}, 3121 (1972).\\
1453: B. W. Lee and Jean Zinn-Justin, Phys. Rev. D {\bf 5}, 3137 (1972).\\
1454: B. W. Lee and Jean Zinn-Justin, Phys. Rev. D {\bf 5}, 3155 (1972).
1455:
1456: \bibitem{LH2008}
1457: J. Ludescher and R. Hofmann, Ann. Phys.{\bf 18}, 271 (2009) [arXiv:0806.0972 [hep-th]].
1458:
1459: \bibitem{Korthals-Altes}
1460: A. M. Polyakov, Phys. Lett. B {\bf 59}, 82 (1975).\\
1461: C. P. Korthals Altes, hep-ph/0408301.\\
1462: C. P. Korthals Altes, Acta Phys. Polon. B {\bf 34}, 5825 (2003).\\
1463: P. Giovannangeli and C. P. Korthals Altes, Nucl. Phys. B {\bf 608}, 203 (2001).\\
1464: C. Korthals-Altes, A. Kovner, and M. A. Stephanov, Phys. Lett. B {\bf
1465: 469}, 205 (1999).
1466:
1467: \bibitem{KH}
1468: C. Korthals-Altes, hep-ph/0406138v2.
1469:
1470: \bibitem{GiacosaHofmann2007}
1471: F. Giacosa and R. Hofmann, Phys. Rev. D {\bf 76}, 085022 (2007).
1472: [hep-th/0703127]
1473:
1474: \bibitem{Diakonov}
1475: D. Diakonov, N. Gromov, V. Petrov, S. Slizovskiy, Phys. Rev. D {\bf 70},
1476: 036003 (2004). [hep-th/0404042]
1477:
1478: \bibitem{HerbstHofmannRohrer2004}
1479: U. Herbst, R. Hofmann, J. Rohrer, Acta Phys. Pol. B{\bf 36}, 881 (2005).
1480:
1481: \bibitem{Bassetto}
1482: A. Bassetto, G. Nardelli, and R. Soldati, {\sl Yang-Mills Theories in
1483: Algebraic Non-Covariant Gauges: Canonical Quantization and
1484: Renormalization}, World Scientific Publishing Company (1991), Singapore.
1485:
1486: \bibitem{KellerDA}
1487: J. Keller, Diploma thesis Universit\"at Heidelberg, arXiv:0801.3961.
1488:
1489: \end{thebibliography}
1490:
1491: \end{document}
1492: