1: \documentclass[12pt,preprint]{article}
2: \usepackage{epsfig}
3: \usepackage{psfig}
4: \renewcommand{\baselinestretch}{2}\normalsize
5: \setlength{\baselineskip}{20mm}
6: \textwidth 15.0 true cm
7: \textheight 22.0 true cm
8: \headheight 0 cm
9: \headsep 0 cm
10: \topmargin 0.4 true in
11: \oddsidemargin 0.25 true in
12: \newcommand{\eqb}{\begin{equation}}
13: \newcommand{\eqe}{\end{equation}}
14: \newcommand{\dmb}{\begin{displaymath}}
15: \newcommand{\dme}{\end{displaymath}}
16: \newcommand{\pd}{\partial}
17: \newcommand{\ep}{\varepsilon}
18: \newcommand{\eab}{\begin{eqnarray}}
19: \newcommand{\eae}{\end{eqnarray}}
20: \newcommand{\ra}{\right\rangle}
21: \newcommand{\la}{\left\langle}
22: \newcommand{\e}{\mbox{e}}
23: \newcommand{\be}{\begin{equation}}
24: \newcommand{\ee}{\end{equation}}
25: \newcommand{\sgn}{\text{sgn}\,}
26: \newcommand{\munu}{{\mu\nu}}
27: \newcommand{\ad}{{\dot{\alpha}}}
28: \newcommand{\bd}{{\dot{\beta}}}
29: \newcommand{\La}{\Lambda}
30:
31: \begin{document}
32: \begin{titlepage}
33: \begin{flushright}
34: 2004-09 \\
35: HD-THEP-04-19\\
36: \end{flushright}
37: \vspace{0.6cm}
38:
39:
40: \begin{center}
41: \Large{{\bf Nonperturbative approach to
42: Yang-Mills thermodynamics}}
43:
44: \vspace{1cm}
45:
46: Ralf Hofmann
47:
48: \end{center}
49: \vspace{0.3cm}
50:
51: \begin{center}
52: {\em
53: Institut f\"ur Theoretische Physik\\
54: Universit\"at Heidelberg\\
55: Philosophenweg 16, 69120 Heidelberg, Germany}\vspace{0.5cm}
56:
57:
58: \end{center}
59: \vspace{0.5cm}
60:
61:
62: \begin{abstract}
63:
64: An analytical, macroscopic approach to SU(N) Yang-Mills thermodynamics is
65: developed. This approach self-consistently assumes that at a
66: temperature much larger than the Yang-Mills scale $\Lambda_{\tiny\mbox{YM,N}}$ (embedded and noninteracting) SU(2)
67: calorons of trivial holonomy form an adjoint
68: Higgs field (electric phase). Macroscopically, this field turns out to be thermodynamically and
69: quantum mechanically stabilized. As a consequence,
70: the problem of the infrared instability in the perturbative loop expansion of thermodynamical potentials,
71: generated by the soft magnetic modes, is resolved.
72: An evolution equation with two fixed points follows for the effective gauge coupling $e(T)$
73: from self-consistent thermodynamics involving the
74: ground-state and its quasiparticle excitations. A plateau value of $e(T)$, which is an
75: attractor of the evolution, is consistent
76: with the existence of isolated magnetic monopoles of conserved charge
77: being generated by dissociating calorons of nontrivial holonomy. The
78: (up to negligible corrections exact) one-loop and downward evolution of
79: $e(T)$ predicts the condensation of magnetic monopoles in a $2^{\tiny\mbox{nd}}$-order phase transition
80: at a critical temperature $T_{E,c}$. At $T_{E,c}$ tree-level
81: massive gauge modes decouple thermodynamically. This is the confinement phase transition identified in lattice simulations.
82: For N=2 we compute
83: the critical exponent taking the mass of
84: the dual photon as an order parameter. For arbitrary N we show the
85: restoration of the global electric $Z_{\tiny\mbox{N}}$
86: symmetry in the monopole condensed (magnetic) phase by investigating
87: the Polyakov loop in the effective theory. The magnetic gauge coupling $g(T)$ starts its downward
88: evolution from zero at
89: $T_{E,c}$ and runs into a logarithmic pole at $T_{M,c}<T_{E,c}$.
90: At $T_{M,c}$ center-vortex loops condense, the abelian gauge modes decouple
91: thermodynamically, and
92: the equation of state is $\rho=-P$ (zero entropy density). The Hagedorn transition
93: to the vortex condensing phase (center phase) goes with a complete breakdown of the local
94: magnetic $Z_{\tiny\mbox{N}}$ symmetry. After a rapid
95: reheating in terms of (intersecting) center-vortex loops has taken place
96: the ground-state pressure vanishes {\sl identically} on tree level.
97: This result is protected against radiative corrections. Throughout the electric
98: and the magnetic phase and for N=2,3 we compute the
99: temperature evolution of the (infrared sensitive) pressure and
100: energy density and for the (infrared insensitive) entropy density and
101: compare our results with lattice data. We show that the
102: disagreement for the two former quantities at low temperature
103: (negative pressure) originates from severe finite-size artefacts in lattice
104: simulations. For the entropy density we
105: obtain excellent agreement with lattice results. The implications of our results for
106: particle physics and cosmology are discussed.
107:
108:
109: \end{abstract}
110:
111: \end{titlepage}
112:
113: \section{Introduction}
114:
115:
116: The beauty and usefulness of the gauge principle for local field theories is
117: generally appreciated. Yet, in a perturbative approach to gauge theories
118: like the SM and its (non)supersymmetric extensions it is hard if not impossible
119: to convincingly address a number of recent experimental and
120: observational results in particle physics and
121: cosmology: nondetection of the Higgs particle at LEP \cite{Higgs2000},
122: indications for a rapid thermalization and strong collective behavior
123: in the early stage of an ultra-relativistic heavy-ion collision
124: \cite{RHIC2003,ShuryakTeaney2001}, dark energy and dark matter at present, a
125: strongly favored epoch of cosmological inflation in the early Universe
126: \cite{Linde1982,Guth1982,Starobinsky1982,WMAP2003},
127: and the likely existence of intergalactic magnetic
128: fields \cite{Dai2002,IMF}. An analytical and
129: nonperturbative approach to strongly interacting gauge
130: theories may further our understanding of these phenomena.
131:
132: It is difficult to gain insights in the dynamics of a strongly interacting
133: field theory by analytical means.
134: We conjecture with Ref.\,\cite{Hagedorn1965} that a thermodynamical approach
135: is an appropriate starting point for such an endeavor.
136: On the one hand, this conjecture is reasonable since a strongly interacting
137: system being in equilibrium communicates perturbations almost instantaneously due to
138: rigid correlations. Thus equilibrium is restored very rapidly.
139: On the other hand, the requirement of
140: thermalization poses strong constraints on the {\sl construction}
141: of a macroscopic, effective theory for the ground state and
142: its (quasiparticle) excitations.
143: The objective of the present paper is the thermodynamics of
144: SU(N) Yang-Mills theories in four dimensions.
145:
146: Let us very briefly recall some aspects of the analytical
147: approaches to thermal SU(N) Yang-Mills theory as they are
148: discussed in the literature. Because of asymptotic
149: freedom \cite{GrossWilczek1973,Politzer1973} one would naively
150: expect thermal perturbation theory to work well for temperatures
151: $T$ much larger than the Yang-Mills scale $\La_{\tiny\mbox{YM,N}}$ since the gauge coupling
152: constant $\bar{g}(T)$ logarithmically approaches zero
153: for $\frac{T}{\La_{\tiny\mbox{YM,N}}}\to\infty$.
154: It is known for a long time that this expectation is too optimistic since at
155: any temperature perturbation theory
156: is plagued by instabilities arising from the infrared sector (weakly screened,
157: soft magnetic modes \cite{Linde1980}). As a consequence, the pressure $P$ can be computed
158: perturbatively only up
159: to (and including) order $\bar{g}^5$. The effects of resummations of
160: one-loop diagrams (hard thermal loops), which rely on a
161: scale separation in terms of the small value of
162: the coupling constant $\bar{g}$, are summarized in terms of a
163: nonlocal effective theory for soft and
164: semi-hard modes \cite{BraatenPisarski1990}. In the computation of radiative
165: effects based on this effective theory
166: infrared effects due to soft modes still appear in an
167: uncontrolled manner. This has
168: lead to the construction of an effective theory where soft modes are collectively described in
169: terms of classical fields whose dynamics is influenced by integrated
170: semi-hard and hard modes \cite{Bodeker1998,Bodapps1998}.
171: In Quantum Chromodynamics (QCD) a perturbative calculation of $P$ was pushed up to order
172: $\bar{g}^6\log\,\bar{g}$, and an additive `nonperturbative'
173: term at this order was fitted to lattice results \cite{Kajantie2002}. Within the
174: perturbative orders a poor convergence of the expansion is observed for
175: temperatures not much larger than the $\overline{\mbox{MS}}$ scale. While
176: the work in \cite{Kajantie2002} is a computational masterpiece it could, by definition,
177: not shed light on the nonperturbative physics of
178: the infrared sector. Screened perturbation theory, which relies
179: on a separation of the tree-level
180: Yang-Mills action using variational parameters,
181: is a very interesting idea. Unfortunately, this approach generates
182: temperature dependent ultraviolet divergences in its presently used
183: form, see \cite{Blaizot2003} for a recent review.
184:
185: The purpose of this paper is to report in a detailed way\footnote{Some
186: aspects of the low-temperature physics
187: are revised in the present paper as compared to \cite{Hofmann2003F}.} on a nonperturbative and
188: analytical approach to the thermodynamics of SU(N) Yang-Mills theory
189: (see \cite{Hofmann2000t2003} for intermediate stages).
190: Conceptually, this approach is similar to the
191: macroscopic Landau-Ginzburg-Abrikosov (LGA) theory for superconductivity in metals
192: \cite{GinzburgLandau1950,Abrikosov1957}. Recall, that this theory does not derive
193: the condensation of Cooper pairs from first principles but rather describes
194: the condensate by a nonvanishing amplitude of a complex scalar field (local order parameter)
195: which is charged under the electromagnetic gauge group U(1). This
196: nonvanishing amplitude is driven by a phenomenologically introduced potential $V$.
197: As a consequence, a (macroscopic) U(1) gauge field $a_\rho$, which is deprived of the
198: microscopic gauge-field fluctuations associated with
199: the formation of Cooper pairs and their subsequent condensation, acquires mass, the U(1) symmetry is
200: spontaneously broken, and physical phenomena originating from
201: this breakdown can be explored in dependence of the parameters appearing in the effective action,
202: and in dependence of an external magnetic field and/or temperature.
203:
204: When applying this idea to the construction of a macroscopic
205: theory for SU(N) Yang-Mills thermodynamics (YMTD) one is in a much better
206: position as far as the uniqueness of the stabilizing potentials
207: in each phase of the theory is concerned. These potentials are
208: determined by thermodynamics and the requirement
209: that, in a first step of the construction, they admit
210: energy- and pressure-free macroscopic configurations describing the collective effects in
211: an ensemble of energy- and pressure-free, noninteracting, and self-dual topological field
212: configurations in the underlying theory.
213: If a particular phase supports propagating gauge modes then, in a second step,
214: the interactions between these topological defects
215: are treated by solving the macroscopic gauge-field equations in terms of a pure
216: gauge configuration in the background of the (inert) energy- and pressure-free
217: scalar field.
218:
219: More specifically, we assume that at a large temperature a macroscopic
220: adjoint scalar field $\phi$
221: is generated by a dilute gas of trivial-holonomy calorons
222: \footnote{We discuss in Sec.\,\ref{MPlanck} why the critical
223: temperature $T_P$ for the onset of the formation of $\phi$ should be comparable to
224: the cutoff-scale for the local field theory in four dimensions.}
225: \cite{HarrigtonShepard1977}. Calorons are Bogomolnyi-Prasad-Sommerfield (BPS)
226: saturated (or self-dual) solutions \cite{PrasadSommerfield1975}
227: to the classical Yang-Mills equations of motion in four-dimensional Euclidean spacetime\footnote{Whenever we speak of a topological soliton
228: this automatically includes the antisoliton.} (time coordinate $\tau$ is
229: compactified on a circle, $0\le\tau\le\frac{1}{T}$) with varying
230: topological charge and embedding in SU(N). Calorons are topologically
231: nontrivial, saturate the lowest possible value of the Euclidean action
232: in a given topological sector, and thus are energy- and pressure-free configurations.
233: Calorons with nontrivial holonomy have BPS magnetic monopole constituents
234: \cite{Nahm1984,KraanVanBaalNPB1998,vanBaalKraalPLB1998}. Their one-loop effective
235: action scales with the three volume of the system \cite{GrossPisarskiYaffe1981}, and
236: thus they should play no role in the thermodynamic limit.
237: This conclusion, however, is no longer valid if the
238: system generates domains of large but finite volume whose boundaries
239: are generated by discontinuous changes of the color orientation of
240: the field $\phi$. Microscopically, nontrivial-holonomy
241: calorons can be dynamically generated out
242: of trivial-holonomy calorons by macroscopic domain collisions. These calorons dissociate into
243: their magnetic monopole constituents subsequently, see \cite{Diakonov} for a
244: discussion of the destabilizing effects of quantum fluctuations in the case of
245: nontrivial holonomy. We thus anticipate the occurence of
246: isolated magnetic charge whose
247: abundance is governed by the ($T$ dependent)
248: typical volume of a domain.
249:
250: The property of vanishing energy and pressure of a caloron derives from its self-duality,
251: that is, the kinetic and the interaction part in the Euclidean energy-momentum tensor
252: precisely cancel when evaluated on a caloron. A potential $V_E$
253: (the subscript $E$ stands for electric phase) is constructed which stabilizes
254: the modulus $|\phi|$ for given $T$ quantum mechanically and thermodynamically
255: and which reflects the assumption that $\phi$ is composed of noninteracting, trivial-holonomy
256: calorons. We would like to stress at this point that the effects of calorons are reflected
257: as a $\frac{1}{\sqrt{T}}$ dependence of the modulus of $\phi$.
258: As a consequence, the nontrivial-topology sector of the
259: theory, indeed, is irrelevant at
260: asymptotically large temperatures.
261:
262: A unique decomposition of each
263: gauge-field configuration $A_\rho$ contributing to the partition function of the
264: fundamental Yang-Mills theory is
265: %************
266: \eqb
267: \label{decfl}
268: A_\rho=A_\rho^{\tiny\mbox{top}}+a_\rho\,.
269: \eqe
270: %************
271: In Eq.\,(\ref{decfl}) $A_\rho^{\tiny\mbox{top}}$ is a minimally (that is, BPS saturated)
272: topological part, represented by calorons,
273: and $a_\rho$ denotes a remainder which has trivial topology. The
274: configurations in $A_\rho^{\tiny\mbox{top}}$ having trivial holonomy would build up
275: the ground state described by $\phi$ if no holonomy-changing interactions
276: between them were allowed for. A change in holonomy by
277: interactions, mediated by the topologically trivial sector, will macroscopically manifest itself in terms of a
278: {\sl finite}, pure-gauge background $a_\rho^{bg}$. A fluctuation
279: $\delta a_\rho$ about this background
280: acquires mass by the adjoint Higgs mechanism if
281: $[\phi,\delta a_\rho]\not=0$ and thus the underlying gauge symmetry SU(N)
282: is spontaneously broken to U(1)$^{\tiny\mbox{N-1}}$ at most. The degree of gauge
283: symmetry breaking by calorons is a boundary
284: condition set at an asymptotically high temperature $T_P$
285: where the effect of $\phi\propto T^{-1/2}$ on the Yang-Mills spectrum
286: and its pressure $\sim T^4$ is very small since the ground state pressure scales as
287: $\propto T$. On the one hand, Higgs-mechanism
288: induced masses provide infrared cutoffs in the
289: loop expansions of thermodynamical quantities
290: which resolves the problem of the infrared
291: instability entcountered in perturbation theory.
292: On the other hand, the compositeness scale $|\phi|$ constrains the
293: hardness of quantum fluctuations, and so the usual
294: renormalization program needed to address ultra-violet divergences in perturbation theory is superfluous
295: in the effective theory. Notice that this way of introducing
296: a composite field $\phi$ in an effective description differs from the usual implementation of a
297: Wilsonian flow, where {\sl high-momentum} modes are successively
298: integrated out \cite{BraatenPisarski1990,Wetterich1996},
299: since $\phi$ is built of calorons with an `instanton' radius $\rho$ not being smaller than $|\phi|^{-1}$.
300: At the present stage the description of the ground state of an SU(N) Yang-Mills
301: theory at high temperatures in terms of
302: the field $\phi$ is self-consistent. The phase and the modulus of the field $\phi$ are derived
303: from a microscopic definition in \cite{HerbstHofmann2004}.
304:
305: The nonperturbative approach to SU(N) YMTD proposed here
306: implies the existence of three rather
307: than two phases: an electric phase at high temperatures, a magnetic
308: phase for a small range of temperatures comparable to the scale
309: $\La_{\tiny\mbox{YM,N}}$, and a center phase for
310: low temperatures. The ground state in the magnetic phase
311: confines fundamental, static test charges
312: but allows for the propagation of
313: massive, dual gauge bosons.
314: The center phase is thermodynamically disconnected
315: from the magnetic and the electric phase.
316: In the electric phase an evolution equation for the
317: effective gauge coupling constant $e$, which follows from the requirement of thermodynamical
318: self-consistency of the one-loop expression for the pressure, has two fixed points associated with
319: a highest and a lowest attainable temperature $T_P$ and $T_{E,c}$. It turns out that practically all strong-interaction
320: effects of the theory are described by a temperature dependent ground-state pressure and
321: tree-level masses for thermal quasiparticles such that higher loop corrections to
322: thermodynamical quantities are tiny.
323:
324: At $T_{E,c}$ the effective coupling $e(T)$ exhibits a thin divergence of the form
325: %*******
326: \eqb
327: \label{coupldiv}
328: e(T)\sim -\log(T-T_c)\,,
329: \eqe
330: %********
331: and the theory undergoes a 2$^{\mbox{\tiny{nd}}}$
332: order like phase transition to a magnetic phase which is driven by the
333: condensation of some of the magnetic monopoles
334: residing inside dissociating nontrivial-holonomy calorons. In this transition a part of the
335: continuous gauge symmetry, which survived the formation of the adjoint Higgs field $\phi$
336: in the electric phase, is
337: broken spontaneously and the tree-level
338: massive gauge modes of the electric phase decouple thermodynamically. In the case of submaximal gauge-symmetry
339: breaking by $\phi$ in the electric phase condensates of magnetic and
340: color-magnetic monopoles occur in the magnetic phase. The former are described by
341: complex scalar and the latter by adjoint Higgs fields. In the case of maximal
342: gauge symmetry breaking to U(1)$^{\tiny\mbox{N-1}}$, which we will only investigate in this paper,
343: the (local) magnetic center symmetry $Z_{\tiny\mbox{N/2,mag}}$ and the continuous gauge symmetry U(1)$^{\tiny\mbox{N/2-1}}$
344: survive the transition to the magnetic phase, the (global) electric center symmetry
345: $Z_{\tiny\mbox{N/2,elec}}$ is fully restored. An approach to the
346: thermodynamics in the magnetic phase, which is conceptually analoguous
347: to the one in the electric phase, yields an evolution equation
348: for the magnetic gauge coupling $g$ which has two
349: fixed points at $T_{E,c}$ and
350: $T_{M,c}$ (highest and lowest attainable temperature). Approaching $T_{M,c}$ from above,
351: the equation of state is increasingly dominated by the ground state contributions.
352: At $T_{M,c}$ we have
353: %********
354: \eqb
355: \label{eosintro}
356: \rho\sim -P\,.
357: \eqe
358: %*******
359: The theory undergoes a phase transition to a phase whose ground-state is
360: a condensate of center-vortex loops. In this phase
361: $Z_{\tiny\mbox{N,mag}}$ is entirely broken, and all gauge boson
362: excitations are thermodynamically decoupled. Once each of the vortex-loop condensates,
363: described by nonlocally defined
364: complex scalar fields, has relaxed to the one of the N degenerate minima
365: of its potential, the energy density and the pressure of the
366: ground state are precisely zero (no radiative corrections), and the system has created particles
367: by local $Z_{\tiny\mbox{N}}$ phase shifts of each vortex-condensate field which
368: are associated with localized (intersecting) center fluxes forming closed loops.
369: The corresponding density of states is over-exponentially
370: rising implying that the magnetic-center transiton is of the
371: Hagedorn type and thus nonthermal.
372:
373: There are many claims in the scenario outlined above. We will, step by step,
374: verify them as we proceed. The paper is organized as follows:
375:
376: In Sec.\,\ref{EP} we explain our approach
377: to the electric phase. We start with the basic assumption that it is noninteracting
378: trivial-holonomy calorons that form a macroscopic adjoint Higgs field $\phi$ at
379: high temperatures (electric phase)
380: and explore its consequences. A nonlocal definition for $\phi$ is given.
381: We then elucidate the details of the ground-state dynamics and the properties
382: of topology-free gauge modes. Subsequently,
383: an evolution equation for the effective gauge coupling constant $e$ is
384: derived and solved, interpretations of the solution are given, and an
385: argument is provided why the temperature $T_P$ for the onset of caloron 'condensation'
386: has to be comparable to the cutoff-scale for the
387: local field-theory description in four dimensions. In a next step, we perform the
388: counting of isolated magnetic
389: monopoles species in the effective theory for the electric phase
390: when assuming maximal gauge symmetry breaking
391: by $\phi$. The next part of Sec.\,\ref{EP} is devoted to a discussion of two-loop corrections to
392: thermodynamical quantities. For the SU(2) case we investigate
393: the simplest one-loop contribution to the `photon' polarization
394: and perform formal weak and strong coupling limits of this expression. We also discuss the
395: implementation of thermodynamical self-consistency when higher loop corrections to the pressure
396: are taken into account.
397:
398: In Sec.\,\ref{MP} we investigate the magnetic phase, again
399: assuming maximal gauge symmetry by caloron 'condensation': the pattern of gauge symmetry breaking
400: by monopole condensation is explored, the thermodynamics of the ground state and
401: its excitations is elucidated, an evolution equation for the magnetic gauge coupling constant
402: $g$ is derived. Solutions to this equations are obtained numerically and their
403: implications are discussed. Finally, we discuss the Polyakov loop
404: in the electric and the magnetic phase and compute the critical exponent of the phase transition for SU(2).
405:
406: In Sec.\,\ref{CVCM} we investigate the center phase. A nonlocal definition
407: for the local fields describing the condensed center-vortex loops is given, their transformation properties
408: under magnetic center rotations are determined, and their dynamics is
409: discussed.
410:
411: In Sec.\,\ref{MCC} we derive a matching condition for
412: the mass scales $\La_E$ and $\La_M$ which appear in the respective potentials for the caloron and magnetic
413: monopole condensates.
414:
415: In Sec.\,\ref{PEEN} we compute the temperature evolution of
416: the thermodynamical potentials pressure, energy density, and entropy density throughout the electric and
417: the magnetic phases at one loop for N=2,3 and compare our
418: results with lattice data.
419:
420: A conclusion and a discussion of likely implications of our results for
421: particle physics and cosmology are given in Sec.\,\ref{CO}.
422:
423: \section{The electric phase\label{EP}}
424:
425: \subsection{Conceptual framework\label{PR}}
426:
427: Our analysis is based on the following assumption about the ground-state physics
428: characterizing SU(N) YMTD at high temperatures.
429: \vspace{0.4cm}\\
430: {\sl At a temperature
431: $T_P\gg\Lambda_{\tiny\mbox{YM,N}}$ SU(N) YMTD, defined on a Euclidean, four-dimensional, and
432: flat spacetime, generates an adjoint Higgs field $\phi$ out of noninteracting (dilute), trivial-holonomy
433: SU(2) calorons.} \vspace{0.4cm}\\
434: \noindent Calorons are BPS saturated solutions
435: to the Euclidean equations of motion of SU(N) Yang-Mills theory at finite temperature
436: \cite{HarrigtonShepard1977,Nahm1984,KraanVanBaalNPB1998,vanBaalKraalPLB1998}.
437: One distinguishes SU(2) calorons according to their holonomy, that is, the behavior of the
438: Polyakov loop
439: %**********
440: \eqb
441: \label{polloop}
442: {\bf P}={\cal P}\exp\left[i\bar{g}\int_0^{1/T}d\tau A_4(\vec{x}, \tau)\right]\,
443: \eqe
444: %*********
445: at $|{\vec x}|\to \infty$ when evaluated on
446: the solution. In Eq.\,(\ref{polloop}) ${\cal P}$ denotes the
447: path-ordering symbol and $\bar{g}$ the gauge coupling constant
448: of the SU(2) Yang-Mills theory.
449: Trivial (nontrivial) holonomy means that
450: have ${\bf P}_{\tiny{|{\vec x}|\to \infty}}={\bf 1}\ \ \ (\not={\bf 1})$. In the former
451: case the SU(2) caloron has no isolated magnetic-monopole constituents, in the
452: latter case it exhibits a monopole and its
453: antimonopole. The masses of these constituents are determined
454: by the value of $A_4(|{\vec x}|\to \infty$. In the following
455: we only consider SU(2) nontrivial-holonomy calorons with no net magnetic charge.
456: Analytical expression for SU(2) caloron solutions
457: of trivial (nontrivial) holonomy can be found in \cite{HarrigtonShepard1977} (\cite{Nahm1984,KraanVanBaalNPB1998,Brower1998}),
458: see also Fig.\,1..
459: %***********************
460: \begin{figure}
461: \begin{center}
462: \leavevmode
463: %\epsfxsize=9.cm
464: \leavevmode
465: %\epsffile[80 25 534 344]{}
466: \vspace{4.3cm}
467: \special{psfile=Fig-1.ps angle=0 voffset=-140
468: hoffset=-205 hscale=55 vscale=75}
469: \end{center}
470: \caption{Action density of SU(2) nontrivial-holonomy calorons with
471: increasing 'instanton radius' (left to right) at fixed temperature. Figures are taken
472: from a paper by Kraan and van Baal. The peaks of the action density
473: coincide with the positions of constituent BPS monopoles.\label{instrad}}
474: \end{figure}
475: %************************
476: Since calorons are BPS saturated or self-dual their
477: energy-momentum tensor vanishes identically.
478:
479: An SU(2) caloron of topological charge $k$ has a classical
480: Euclidean action $S_{\mbox\tiny{cal},2}=\frac{8\pi^2}{\bar{g}^2}k$.
481: For trivial holonomy the one-loop effective action of a charge-one caloron
482: is given as \cite{GrossPisarskiYaffe1981}
483: %********
484: \eqb
485: \label{GPYSeffth}
486: S_{\tiny\mbox{eff}}=\frac{8\pi^2}{\bar{g}^2}+\frac{4}{3}(\pi\rho T)^2
487: \eqe
488: %*******
489: where $\rho$ and $T$ denote the 'instanton' radius
490: and temperature, respectively. For large $\bar{g}$ and small
491: enough $\rho$ the trivial-holonomy caloron
492: thus sizably contributes to the partition function of the theory.
493: The one-loop effective
494: action of a nontrivial-holonomy caloron is
495: %********
496: \eqb
497: \label{GPYSeffnth}
498: S_{\tiny\mbox{eff}}\propto T^3 V
499: \eqe
500: %*******
501: where $V$ denotes the spatial volume of the system.
502: In the thermodynamic limit $V\to\infty$ nontrivial-holonomy
503: calorons thus do not contribute to the partition function.
504: As we will show below, however, the thermodynamic limit is
505: not physical due to a domanization of the ground state
506: of the theory. The suppression of nontrivial-holonomy
507: calorons in the partition function is thus governed by the size
508: of a typical domain.
509:
510: It was shown in \cite{Diakonov} that SU(2) calorons are
511: unstable under one-loop quantum fluctuations.
512: Namely, for a holonomy close to trivial there is an
513: attractive potential between constituent monopole and antimonopole. In the opposite case
514: the potential is repulsive implying the dissociation of the caloron
515: into a monopole-antimonopole pair. In a mesoscopic level, an isolated
516: monopole arises at a point in space
517: where four or more Higgs-field domains meet \cite{Kibble1976}.
518:
519: Since the action {\sl density} of a caloron
520: is $T$-dependent the action density of the
521: macroscopic, adjoint Higgs field $\phi$ should be $T$-dependent through
522: the $T$-dependence of the configuration $\phi$.
523:
524: The effective
525: theory describing the (electric) phase macroscopically is an adjoint Higgs model:
526: %*********
527: \eqb
528: \label{actE}
529: S_E=\int_0^{1/T}d\tau\hspace{-0.1cm}\int d^3x\,\left(\frac{1}{2}\,\mbox{tr}_{\tiny\mbox{N}}
530: \,G_{\mu\nu}G_{\mu\nu}+\mbox{tr}_{\tiny\mbox{N}}\,{\cal D}_\mu\phi{\cal D}_\mu\phi+
531: V_E(\phi)\right)\,.
532: \eqe
533: %*********
534: In Eq.\,(\ref{actE}) $V_E$ denotes the potential responsible for the stabilization of
535: $\phi$. The covariant derivative is
536: defined as ${\cal D}_\rho\phi=\pd_\rho+ie[\phi,a_\rho]$, the field
537: strength as $G_{\mu\nu}=G^a_{\mu\nu}t^a$, where
538: $G^a_{\mu\nu}=\pd_\mu a^a_\nu-\pd_\nu a^a_\mu-ef^{abc}a^b_\mu a^c_\nu$,
539: $e$ denotes the effective gauge coupling constant, and $\mbox{tr}_{\tiny\mbox{N}}\,t^a t^b\equiv
540: 1/2\,\delta^{ab}$. While the effect of nontrivial topology is described by the scalar
541: sector of the effective theory the curvature $G_{\mu\nu}$ is
542: generated by the topologically trivial fluctuations.
543:
544: For future work \cite{HerbstHofmann2004}
545: we propose the following nonlocal definition for the phase of a given SU(2) block
546: $\tilde{\phi}$.
547: %*********
548: \eqb
549: \label{locdefphi}
550: \frac{\tilde{\phi}^a(\tau)}{|\tilde{\phi}|}\sim \mbox{tr}_{2} \int d\rho\, d^3x\,
551: \lambda^a\,F_{\mu\nu}((\tau,0))\,[(\tau,0),(\tau,\vec{x})]
552: F_{\mu\nu}((\tau,\vec{x}))[(\tau,\vec{x}),(\tau,0)]+\cdots\,.
553: \,
554: \eqe
555: %*********
556: The dots in Eq.\,(\ref{locdefphi}) denote the contributions of higher $n$-point functions, and
557: the $\sim$ sign indicates that this expansion very likely is asymptotic
558: at best as a powers series in a dimensionless parameter
559: $\xi$. This, however, is not an obstacle to determining $\tilde{\phi}'s$ phase and
560: modulus \cite{HerbstHofmann2004}.
561: Each block $\tilde{\phi}$ receives a nontrivial phase by the corresponding SU(2)-embedded trivial-holonomy
562: caloron $A^C_\beta$ (or anticaloron $A^A_\beta$) over which the correlator
563: in Eq.\,(\ref{locdefphi}) is evaluated\footnote{The topological
564: charges of $A^C_\beta$ or $A^A_\beta$ may, in principle, vary
565: from block to block.}. In the definition Eq.\,(\ref{locdefphi})
566: $[(\tau,0),(\tau,\vec{x})]$ denotes a Wilson line in the fundamental representation
567: which is taken to be along a
568: straight path connecting the two points $(\tau,0)$ and $(\tau,\vec{x})$:
569: %*********
570: \eqb
571: \label{Wilsonline}
572: [(\tau,0),(\tau,\vec{x})]
573: \equiv {\cal P}\,\exp\left[i\int_{(\tau,0)}^{(\tau,\vec{x})}dy_\beta\,A_\beta(y)\right]\,.
574: \eqe
575: %*********
576: In a lattice simulation at finite temperature $T$ the average in Eq.\,(\ref{locdefphi})
577: can be computed by using an ensemble of cooled configurations whose action
578: is an integer multiple of $8\pi^2/{\bar g}^2$\footnote{The nontrivial-holonomy part is then cooled away.}.
579: Local gauge-singlet composites such as the gluon condensate
580: %*******
581: \eqb
582: \label{gluoncondensate}
583: \la \mbox{tr}_{\tiny\mbox{N}}
584: F_{\mu\nu}(x)\,F_{\mu\nu}(x)\ra\,
585: \eqe
586: %********
587: are thermodynamically irrelevant for the following reasons:
588: Since they do not couple to the
589: topologically trivial sector they do not
590: influence the mass spectrum of fluctuations $\delta a_\rho$. Moreover, a singlet
591: composite, arising from noninteracting trivial-holonomy calorons,
592: would have zero energy density and pressure
593: because of the BPS saturation: a situation which cannot be changed by
594: interactions mediated by the trivial sector due to the missing
595: gauge charge. The situation is different though
596: if an axial anomaly, arising from integrated-over chiral fermions,
597: is operative. In this case the composite in
598: Eq.\,(\ref{gluoncondensate}) determines the
599: mass of the axion, and thus it is visible.
600:
601: The key question now is whether the potential $V_E$ in Eq.\,(\ref{actE}) is uniquely
602: determined by our basic assumption.
603: What are the properties of the field $\phi$ that can be deduced?
604: In thermal equilibrium $\phi$ must be periodic in
605: Euclidean time $\tau$ ($0\le\tau\le 1/T$).
606: Since $\phi$ describes the ground state of the thermal system its
607: modulus $|\phi|$ must not depend on $\tau,\vec{x}$ but should depend
608: on $T$. Since $\phi$ is built
609: of noninteracting, self-dual configurations (zero energy density and pressure)
610: it must also be pressure - and energy-free. This is the case
611: if and only if $\phi(\tau)$ (not its modulus!) is BPS saturated, that is,
612: it solves the following equation
613: %*******
614: \eqb
615: \label{BPSEP}
616: \pd_{\tau}\phi=v_E\,
617: \eqe
618: %*******
619: where $v_E$ is a `square root' of the potential $V_E$:
620: %******
621: \eqb
622: \label{squareroot}
623: V_E(\phi)=\mbox{tr}_{\tiny\mbox{N}}v_E^\dagger v_E\,.
624: \eqe
625: %*******
626: The above properties fix the potential uniquely
627: to be $V_E(\phi)\mbox{tr}_{\tiny\mbox{N}}=\Lambda_E^6\mbox{tr}_{\tiny\mbox{N}}(\phi^{2})^{-1}$.
628: As it turns out, a (winding) solution to Eq.\,(\ref{BPSEP})
629: is quantum mechanically and thermodynamically
630: inert and thus can be used as a background to the macroscopic equation of
631: motion for the trivial-topology sector of the theory.
632:
633: The equation of motion
634: %*******
635: \eqb
636: \label{GF}
637: {\cal D}_\mu G_{\mu\nu}=2ie[\phi,{\cal D}_\nu\phi]\,,
638: \eqe
639: %********
640: which follows from the effective action (\ref{actE}),
641: determines a configuration $a^{bg}_\rho$. For $a^{bg}_\rho$ to describe
642: the ground state of the theory it
643: needs to be pure gauge, that is, $G_{\mu\nu}[a^{bg}_\rho]\equiv 0$. Otherwise
644: the invariance of the thermal system under spatial rotations would be spontaneously broken.
645: It will turn out that such a pure-gauge solution $a^{bg}_\rho\propto \frac{T}{e}$
646: exists for ${\cal D}_\nu\phi\equiv 0$.
647: As a consequence, the action density in Eq.\,(\ref{actE}) when
648: evaluated on $\phi,a^{bg}_\rho$ reduces to the potential $V_E$. We thus describe
649: on a macroscopic level interactions between trivial-holonomy
650: calorons as mediated by the topologically trivial sector.
651: Namely, the vanishing ground-state energy density (pressure) of
652: noninteracting trivial-holonomy calorons is shifted from
653: zero to $V_E(\phi)$ ($-V_E(\phi)$). Moreover, a macroscopic
654: holonomy arises which indicates that (unstable) nontrivial-holonomy
655: calorons are generated by gluon exchange and, as a consequence, that isolated
656: magnetic monopoles occur. This precludes our conceptual
657: discussion of the ground-state physics.
658:
659: An adjoint Higgs field $\phi$ breaks the SU(N) gauge
660: symmetry down to U(1)$^{\tiny\mbox{N-1}}$ at most. Whether SU(N) gauge
661: symmetry is broken maximally or submaximally is decided by a boundary
662: condition to the BPS equation (\ref{BPSEP}) set at an
663: asymptotically high temperature. Interacting calorons emit and absorb
664: gauge-field fluctuations $\delta a_\rho$. To discuss their
665: quasiparticle mass spectrum a gauge transformation to $a^{bg}_\rho\equiv 0$ (unitary gauge)
666: must be performed. We will show explicitly for the SU(2) case that such a
667: transformation is on the one hand nonperiodic but on the other hand a
668: symmetry transformation for all thermodynamical quantities. This is true since the transformation does
669: not affect the periodicity
670: of the fluctuations $\delta a_\rho$ (no Hosotani mechanism \cite{Hosotani1983}). The nonperiodic gauge transformation maps the
671: Polyakov loop from ${\bf -1}$ to ${\bf 1}$, therefore generates a
672: global electric center transformation and thus interpolates between the
673: two physically equivalent ground states of the theory. As a consequence,
674: the global symmetry $Z_{\tiny\mbox{2,elec}}$ is spontaneously broken and
675: hence the electric phase is deconfining. The generalization to arbitrary
676: N is straight forward.
677:
678: There are tree-level heavy
679: (TLH) and tree-level massless (TLM) modes in $\delta a_\rho$.
680: Due to the $T$ dependence of $\phi$
681: on-shell TLH modes are thermal quasiparticles.
682:
683: Due to the $T$ dependent Higgs mechanism and the $T$ dependent
684: ground-state energy, which are
685: both generated by the macroscopic field $\phi$, implicit temperature
686: dependences arise in a loop expansion of thermodynamical quantities.
687:
688: To guarantee in the effective theory the validity of the Legendre transformations between
689: thermodynamical quantities, as they can be derived from the partition function of the underlying
690: theory, thermodynamical self-consistency has to be demanded. This condition determines the temperature
691: evolution of the effective gauge coupling constant $e$ with temperature.
692: As we will see, there is an attractor to this evolution which is the constancy of
693: $e$ except for a logarithmic pole at a temperature $T_{c,E}$. We thus recover in the effective theory
694: the ultraviolet-infrared decoupling that follows from the
695: renormalizability of the underlying theory.
696:
697: The approach to the ground-state dynamics is an inductive one.
698: Namely, we first define a potential $V_E(\phi)$ and subsequently show that this definition
699: implies the above properties of $\phi$, a small action for
700: calorons at the temperature where they are assumed to first form the field $\phi$, and the
701: existence of a pure-gauge solution $a^{bg}_\rho$ to Eq.\,(\ref{GF}). The thermal system
702: decomposes into a ground state and a part represented by very weakly interacting
703: quasiparticle fluctuations \footnote{We compute two-loop correction to the pressure in
704: \cite{HerbstHofmannRohrer2004}.}. One can consider the former as a heat bath for the latter at low temperatures
705: and vice versa at high temperatures.
706:
707:
708: \subsection{Caloron `condensate', macroscopically \label{CCM}}
709:
710: \subsubsection{The case of even N: Ground-state physics}
711:
712: We first address the macroscopic dynamics
713: of the adjoint Higgs field $\phi$ when N is even.
714: The case of odd N is discussed in Sec.\,\ref{oddN}.
715: We may always work in a gauge where $\phi$ is SU(2) block diagonal:
716: %********
717: \eqb
718: \label{SU2diag}
719: \phi=\left(\begin{array}{cccc}
720: \tilde{\phi}_1 & {\bf 0} & {\bf 0} &\cdots\\
721: {\bf 0}&\tilde{\phi}_2& {\bf 0} &\cdots\\
722: {\bf 0}&{\bf 0}& \ddots & \\
723: \vdots& \vdots& &
724: \end{array}\right)\,.
725: \eqe
726: %********
727: In Eq.\,(\ref{SU2diag}) each field
728: $\tilde{\phi}_l,\,(l=1,\cdots,\mbox{N}/2)$,
729: lives in an independent SU(2) subalgebra of SU(N), and we define the SU(2) modulus as
730: %********
731: \eqb
732: \label{modulus}
733: |\tilde{\phi}_l|^2\equiv \frac{1}{2}\,\mbox{tr}_2\,\tilde{\phi_l}^2\,.
734: \eqe
735: %*********
736: The potential $V_E$ in Eq.\,(\ref{actE}) is defined as
737: %********
738: \eqb
739: \label{potentialE}
740: V_E=\mbox{tr}_{\tiny\mbox{N}} v^\dagger_E v_E\equiv\Lambda_E^6\,
741: \mbox{tr}_{\tiny\mbox{N}}\,(\phi^2)^{-1}\,
742: \eqe
743: %********
744: where $\La_E$ is a fixed mass scale generated by dimensional transmutation.
745: It is important to note already at this point
746: that there is only one independent mass scale describing the
747: thermodynamics in {\sl all} phases of the theory.
748:
749: We define $v_E$ as follows:
750: %********
751: \eqb
752: \label{sqpotentialE}
753: v_E\equiv i\La_E^3\left(\begin{array}{cccc}
754: \lambda_1\tilde{\phi}_1/|\tilde{\phi}_1|^2& {\bf 0} & {\bf 0} &\cdots\\
755: {\bf 0}&\lambda_1\tilde{\phi}_2/|\tilde{\phi}_2|^2& {\bf 0} &\cdots\\
756: {\bf 0}& {\bf 0}& \ddots & \\
757: \vdots& \vdots& &
758: \end{array}\right)\,
759: \eqe
760: %********
761: where $\lambda_i,\,(i=1,2,3)$, denote the Pauli matrices. This definition is modulo
762: global SU(2)-block gauge transformations. This
763: global symmetry (the `direction' of winding along a U(1) circle around the group manifold $S^3$ od SU(2))
764: will translate into a gauge symmetry once the theory condenses magnetic monopoles,
765: see Sec.\,\ref{windingmag}.
766:
767: In SU(2) decomposition
768: the solution $\tilde{\phi}_l$ to the BPS equation
769: (\ref{BPSEP}) reads
770: %********
771: \eqb
772: \label{solBPSE}
773: \hspace{-0.5cm}\tilde{\phi}_l(\tau)=\sqrt{\frac{\Lambda_E^3}{2\pi T |K(l)|}}\,\lambda_3
774: \exp(-2\pi i T K(l)\lambda_1\tau)\,
775: \eqe
776: %*********
777: where $K(l)$ is a non-zero integer. The solution in Eq.\,(\ref{solBPSE})
778: is periodic in $\tau$ and depends on $T$.
779: The set $\{K(1),\dots,K(\mbox{N}/2)\}$, which is a boundary condition to
780: Eq.\,(\ref{BPSEP}) at the large temperature $T_P$, determines the value of the potential
781: at a given temperature. It also specifies to what extent the
782: SU(N) gauge symmetry is spontaneously broken by caloron 'condensation'. For example,
783: the sets $\{1,1\}$ and $\{1,2\}$
784: break SU(4) down to SU(2)$\times$SU(2)$\times$U(1)
785: and U(1)$^3$, respectively. Out of 15 gauge-field
786: modes 7 modes remain massless in the former and 3 modes in the latter case.
787: For a description in terms of a given SU(N) Yang-Mills theory the
788: set $\{K(1),\dots,K(\mbox{N}/2)\}$ has to be measured,
789: see also the discussion in Sec.\,\ref{MPlanck}. For definiteness and simplicity we
790: assume in the following that the gauge symmetry breaking
791: is maximal in such a way that the
792: potential $V_E(\phi)$ is minimal (MGSB).
793: This corresponds to the boundary condition $\{K(1),\dots,K(\mbox{N}/2)\}=\{1,2,\dots,\mbox{N}/2\}$
794: or a (local) permutation thereof.
795:
796: Let us now verify that the solution in Eq.\,(\ref{solBPSE}) is quantum mechanically and thermodynamically
797: stabilized. Assuming MGSB, the following ratios are obtained
798: %*********
799: \eqb
800: \label{ratios}
801: \frac{\pd^2_{|\tilde{\phi}_l|}V_E}{T^2}=12\pi^2\,l^2\,, \ \ \ \ \ \
802: \frac{\pd^2_{|\tilde{\phi}_l|}V_E}{|\tilde{\phi}_l|^2}=3l^3\,\lambda_E^3\,
803: \eqe
804: %*********
805: where the dimensionless temperature $\lambda_E$ is defined
806: as $\lambda_E\equiv 2\pi T/\La_E$. For N not too large we have
807: $\lambda_E\gg 1$, see Sec.\,\ref{TSCE}. As one can infer from Eq.\,(\ref{ratios}),
808: the mass $m_l^2\equiv \pd^2_{|\tilde{\phi}_l|}V_E$ of collective caloron fluctuations is much larger than $T$ and
809: the compositeness scale $|\tilde{\phi}_l|$. The off-shellness of
810: quantum fluctuations of the field $\tilde{\phi}_l$ is
811: cut off at this scale in Minkowskian or Euclidean signature as
812: %**********
813: \eqb
814: \label{Mineuosn}
815: |p^2-m_l^2|\le |\tilde{\phi}_l|^2\,,\ \ \ \ \ \ \mbox{or} \ \ \ \ \
816: \ p_e^2+m_l^2\le |\tilde{\phi}_l|^2\,,\ \ \ \ (p_e^2\ge 0)\,.
817: \eqe
818: %*********
819: So if no off-shellness in Minkowskian or Euclidean signature is
820: allowed for on the one hand. On the other hand, statistical fluctuations of
821: on-shell $\phi$-particles are strongly Boltzmann suppressed and thus negligible.
822: We conclude
823: that the solution $\phi$ in Eq.\,(\ref{solBPSE}) is
824: stabilized against fluctuations $\delta\phi$ and the potential $V_E$ in Eq.\,(\ref{potentialE})
825: is a truly effective one. Thus $\phi$ is nothing but a background for the
826: thermodynamics of the topologically trivial sector of the theory.
827: As we will see in Sec.\,\ref{TSCE}, topologically trivial quantum fluctuations $\delta a_\rho$
828: generate only a tiny correction to the tree-level value $V_E(\phi)$.
829:
830: Before we investigate the properties of the fluctuations $\delta a_\rho$
831: let us complete our construction of the ground state.
832: The ground-state configurations $a_\rho^{bg}$
833: needs to be a pure-gauge solution to the classical equation of motion Eq.\,(\ref{GF}). In order for
834: $a_\rho^{bg}$ not to break the rotational invariance of the system
835: it needs to be pure gauge. Inserting the background (\ref{solBPSE})
836: (winding numbers: $\{l=1,\dots,l=N/2\}$ for MGSB) into Eq.\,(\ref{GF}),
837: we obtain the following pure-gauge solution:
838: %**********
839: \eqb
840: \label{curvfree}
841: a^{bg}_\rho=\frac{\pi}{e}\,T \delta_{\rho 4}\left(\begin{array}{cccc}
842: \lambda_1& {\bf 0} & {\bf 0} &\cdots\\
843: {\bf 0}&2\,\lambda_1& {\bf 0} &\cdots\\
844: {\bf 0}& {\bf 0}& \ddots & \\
845: \vdots& \vdots& &
846: \end{array}\right)\,.
847: \eqe
848: %***********
849: Moreover, we have
850: %**********
851: \eqb
852: \label{covder0}
853: {\cal D}_\rho\phi=0\,
854: \eqe
855: %***********
856: on $\phi, a^{bg}_\rho$. A remarkable thing has happened:
857: On a macroscopic level we describe the generation
858: of a nontrivial holonomy by interactions
859: between trivial holonomy calorons, mediated by trivial-topology fluctuations,
860: in terms of a macroscopic holonomy associated with $a^{bg}_\rho$!
861: For the microscopic physics this implies the generation of (rare) nontrivial-holonomy calorons
862: and their subsequent dissociation into magnetic monopoles. On a mesoscopic level,
863: this is nothing but the Kibble mechanism for monopole creation \cite{Kibble1976} arising from the
864: domanization of color orientations of $\phi$.
865: Moreover, the vanishing pressure and energy density of a hypothetical
866: ground state composed of noninteracting trivial-holonomy calorons,
867: is shifted to $\mp V_E(\phi)$ with
868: %*********
869: \eqb
870: \label{gspot}
871: V_E(\phi)=\frac{\pi}{2}\,\Lambda_E^3 T \mbox{N(N+2)}\,
872: \eqe
873: %*********
874: by gluon exchange (insert Eqs.\,(\ref{curvfree}) and (\ref{covder0}) into Eq.\,(\ref{actE})).
875:
876: Let us now split the topologically trivial part in Eq.\,(\ref{decfl})
877: further into the ground-state part $a^{bg}_\rho$ and fluctuations $\delta a_\rho$:
878: %******
879: \eqb
880: \label{gffluct}
881: a_\rho=a^{bg}_\rho+\delta a_\rho\,.
882: \eqe
883: %******
884: To make the mass spectrum of the fluctuations $\delta a_\rho$
885: visible it would be desirable to work in unitary
886: gauge where $a^{bg}_\rho\equiv 0$ and thus no coupling
887: of $\delta a_\rho$ to the background $a^{bg}_\rho$ takes place.
888:
889: The gauge rotation $\Omega\equiv e^{i\theta}$,
890: which transforms $\phi$ and $a_\rho$ according to
891: %**********
892: \eab
893: \label{gurot}
894: \phi &\to& \Omega^\dagger\,\phi\,\Omega\nonumber\\
895: a_\rho&\to& \Omega^\dagger\,a_\rho\,\Omega+\frac{i}{e}\,(\pd_\rho\Omega^\dagger)\,\Omega
896: \eae
897: %***********
898: from winding gauge to unitary gauge is for MGSB given as
899: %***********
900: \eqb
901: \label{thetamat}
902: \theta=\left(\begin{array}{cccc}
903: -\pi\lambda_1T\tau & {\bf 0 }& {\bf 0} &\cdots\\
904: {\bf 0}&-2\pi\lambda_1T\tau & {\bf 0} &\cdots\\
905: {\bf 0}&{\bf 0}& \ddots & \\
906: \vdots& \vdots& &
907: \end{array}\right)\equiv \left(\begin{array}{cccc}
908: \theta_1 & {\bf 0 }& {\bf 0} &\cdots\\
909: {\bf 0}&\theta_2& {\bf 0} &\cdots\\
910: {\bf 0}&{\bf 0}& \ddots & \\
911: \vdots& \vdots& &
912: \end{array}\right)\,.
913: \eqe
914: %*********
915: Notice that the gauge transfromation $\Omega$ as parametrized by Eq.\,(\ref{thetamat}) is {\sl not} periodic due to its
916: first, third, fifth, ... block being {\sl antiperiodic} in $\tau$.
917: Is a nonperiodic
918: gauge transformation physically admissible? Let us discuss this
919: for the SU(2) case only. We can make $\Omega=\exp[\frac{-i\pi\lambda_1}{T}]$
920: periodic at the expense of sacrificing its smoothness at the point $\tau=\frac{1}{2T}$
921: by
922: %********
923: \eqb
924: \label{omegatrafo}
925: \Omega\rightarrow \tilde{\Omega}=\Omega Z(\tau)\,
926: \eqe
927: %********
928: where $Z(\tau)$ is a local (electric) $Z_2$ transformation of the form
929: %********
930: \eqb
931: \label{loccentertr}
932: Z(\tau)=2\Theta(\tau-\frac{1}{2T})-1\,,
933: \eqe
934: %********
935: and $\Theta$ denotes the Heavyside step function:
936: %*********
937: \eqb
938: \label{Hevayside}
939: \Theta(x)=\left\{\begin{array}{c}
940: 0\,,\ \ \ \ (x<0)\,,\\
941: \frac{1}{2}\,,\ \ \ \ (x=0)\,,\\
942: 1\,,\ \ \ \ \ (x>0)\,.
943: \end{array}\right.\,.
944: \eqe
945: %***********
946: Applying $\Omega^\prime$
947: to $a_\mu=a^{bg}_\rho+\delta a_\rho$,
948: where $\delta a_\rho$ is a periodic fluctuation in winding gauge, we have
949: %**********
950: \eab
951: \label{trafiwtoug}
952: a_\rho&\rightarrow &\tilde{\Omega}^\dagger(a^{bg}_\rho+\delta a_\rho)\tilde{\Omega}+
953: \frac{i}{e}\pd_\rho \tilde{\Omega}^\dagger\tilde{\Omega}\nonumber\\
954: &=& \Omega^\dagger(a^{bg}_\rho+\delta a_\rho)\Omega+
955: \frac{i}{e}\left((\pd_\rho\Omega^\dagger)\Omega+(\pd_\rho Z(\tau))Z(\tau)\right)\nonumber\\
956: &=&\Omega^\dagger\delta a_\rho\Omega+\frac{2i}{e}\delta(\tau-\frac{1}{2T})Z(\tau)\nonumber\\
957: &=&\Omega^\dagger\delta a_\rho\Omega\,.
958: \eae
959: %***********
960: Since $\Omega^\dagger(0)=-\Omega^\dagger(\frac{1}{T})=\Omega(0)=-\Omega(\frac{1}{T})$
961: we conclude that if the fluctuation $\delta a_\rho$
962: is periodic in winding gauge it is also periodic in unitary
963: gauge. It thus is irrelevant whether we integrate out the
964: fluctuations $\delta a_\rho$ in winding or unitary gauge in a loop expansion
965: of thermodynamical quantities:
966: Hosotani's mechanism \cite{Hosotani1983} does not take place. What changes though
967: is the Polyakov loop evaluated on
968: the background configuration $a^{bg}_\rho$:
969: %*********
970: \eqb
971: \label{Polch}
972: P[a^{bg}_\rho]=-{\bf 1} \rightarrow P[0]={\bf 1}\,.
973: \eqe
974: %***********
975: We conclude that the theory has two equivalent ground states and
976: that the global electric $Z_{\tiny\mbox{2,elec}}$ symmetry is spontaneously broken. We thus have shown that
977: the elecric phase is, indeed, {\sl deconfining}.
978: The generalization of this result to SU(N) with N even is straight forward.
979:
980: In unitary gauge
981: we have
982: %**********
983: \eqb
984: \label{solBPSEug}
985: \hspace{-0.5cm}\tilde{\phi}_l(\tau)\equiv\sqrt{\frac{\Lambda_E^3}{2\pi T l}}\,\lambda_3\,.
986: \eqe
987: %*********
988: Thus the field $\phi$ is constant and diagonal. Moreover, we have
989: %******
990: \eqb
991: \label{Gug}
992: G_{\mu\nu}^a[a_\rho]=G_{\mu\nu}^a[\delta a_\rho]\,.
993: \eqe
994: %**************
995: The gauge-covariant kinetic term for $\phi$ in the action Eq.\,(\ref{actE}) reduces to a sum over
996: mass terms for the TLH modes contained in $\delta a_\rho$. The TLH (TLM)
997: modes are massive (massless) quasiparticles
998: associated with three (two) polarization states. As we will show
999: in \cite{HerbstHofmannRohrer2004} by computing the two-loop correction to the thermodynamical pressure for N=2
1000: these quasiparticles are practically noninteracting for sufficiently large temperatures.
1001:
1002: \subsubsection{The case of odd N\label{oddN}}
1003:
1004: If N is odd then a decomposition of $\phi$ into SU(2) blocks only is no longer possible. One of the SU(2) blocks
1005: in Eq.\,(\ref{SU2diag}) is then replaced by
1006: an SU(3) block. Imposing MGSB and counting the number of independent
1007: stable and unstable magnetic monopoles microscopically on the one hand and macroscopically
1008: in the electric and the magnetic phase on the other hand,
1009: see Sec.\,(\ref{CMP}), we conclude that temporal winding should
1010: take place within the SU(3) block as in Eq.\,(\ref{solBPSE}) but now in
1011: each of the two independent SU(2) subalgebras only
1012: for half the time. The first SU(2) subgroup is generated by
1013: $\bar{\lambda}_i$, ($i=1,\cdots,3$) where
1014: %*********
1015: \eqb
1016: \label{barla}
1017: \bar{\lambda}_1=\left(\begin{array}{ccc}0&0&0\\
1018: 0&0&1\\
1019: 0&1&0\end{array}\right)\,,\ \ \
1020: \bar{\lambda}_2=\left(\begin{array}{ccc}0&0&0\\
1021: 0&0&-i\\
1022: 0&i&0\end{array}\right)\,,\ \ \
1023: \bar{\lambda}_3=\left(\begin{array}{ccc}0&0&0\\
1024: 0&1&0\\
1025: 0&0&-1\end{array}\right)\,.
1026: \eqe
1027: %*********
1028: The second SU(2) subgroup is generated by
1029: $\tilde{\lambda}_i$, ($i=1,\cdots,3$) where
1030: %*********
1031: \eqb
1032: \label{tildela}
1033: \tilde{\lambda}_1=\left(\begin{array}{ccc}0&0&1\\
1034: 0&0&0\\
1035: 1&0&0\end{array}\right)\,,\ \ \
1036: \tilde{\lambda}_2=\left(\begin{array}{ccc}0&0&-i\\
1037: 0&0&0\\
1038: i&0&0\end{array}\right)\,,\ \ \
1039: \tilde{\lambda}_3=\left(\begin{array}{ccc}1&0&0\\
1040: 0&0&0\\
1041: 0&0&-1\end{array}\right)\,.
1042: \eqe
1043: %*********
1044: The solution of the BPS equation (\ref{BPSEP}) for the SU(3) block can be obtained by applying the following prescription to
1045: a single block (\ref{solBPSE}) of {\sl odd} winding number
1046: $K(l)$ in Eq.\,(\ref{SU2diag}): generate two configurations $\tilde{\phi}_{l,1}$ and $\tilde{\phi}_{l,2}$
1047: by replacing the
1048: Pauli matrix $\lambda_i$ in Eq.\,(\ref{solBPSE}) by $\bar{\lambda}_i$ in the first half period, $0\le\tau\le 1/(2T)$, and
1049: by zero in the second half period, $1/(2T)<\tau<1/T$, and by replacing the Pauli matrix $\lambda_i$
1050: in Eq.\,(\ref{solBPSE}) by zero in the first half period, $0<\tau<1/(2T)$, and
1051: by $\tilde{\lambda}_i$ in the second half period, $1/(2T)\le\tau\le 1/T$. Add
1052: $\tilde{\phi}_{l,1}$ and $\tilde{\phi}_{l,2}$ to generate a new
1053: solution to the BPS equation (\ref{BPSEP}) over the entire period. Within this block a member
1054: of the third, dependent SU(2) algebra is generated at $\tau=0,1/(2T)$. To which block
1055: $l$ of odd winding number $K(l)$
1056: this prescription is applied is a boundary condition to the BPS equation.
1057: For simplicity we will proceed in this paper by only quoting results for the case N=3.
1058:
1059: It is clear that for even $N\ge 6$ a decomposition of $\phi$ may contain an even number of SU(3)
1060: blocks besides SU(2) blocks. For definiteness,
1061: we only consider the decomposition into SU(2) blocks as proposed in Eq.\,(\ref{SU2diag}).
1062:
1063: \subsection{Tree-level mass spectrum of TLH modes}
1064:
1065: In absence of radiative corrections only N(N$-$1) TLH modes acquire
1066: mass by the adjoint Higgs mechanism. TLM modes aquire tiny
1067: screening masses radiatively.
1068: This effect will be discussed in Sec.\,\ref{Radcor}.
1069:
1070: In unitary gauge, the TLH mass spectrum calculates as
1071: %***********
1072: \eqb
1073: \label{massspectrum}
1074: m_k^2=-2e^2\,\mbox{tr}\,[\phi,t^{k}][\phi,t^{k}]\,,\ \ \ \ \ \ (k=1,\cdots,\mbox{N(N$-$1)})\,.
1075: \eqe
1076: %***********
1077: The SU(N) generators $t^k$, which are associated with the
1078: TLH modes, are
1079: %********
1080: \eab
1081: \label{TLgen}
1082: t^{IJ}_{rs}&=&1/2\,(\delta_r^I\delta_s^J+\delta_s^I\delta_r^J)\,,\ \ \
1083: \bar{t}^{IJ}_{rs}=-i/2\,(\delta_r^I\delta_s^J-\delta_s^I\delta_r^J)\,,\nonumber\\
1084: (I&=&1,\cdots,\mbox{N};\,J>I;\,r,s=1,\cdots,\mbox{N})\,.
1085: \eae
1086: %*********
1087: By virtue of Eqs.\,(\ref{solBPSEug}), (\ref{massspectrum}), (\ref{TLgen}) we obtain:
1088: %********
1089: \eqb
1090: \label{TLHmasses}
1091: m_{IJ}^2=\bar{m}_{IJ}^2=e^2(\phi_I-\phi_J)^2=
1092: e^2\,\frac{\Lambda_E^3}{2\pi T}\left\{\begin{array}{cc}
1093: \left(\frac{1}{\sqrt{I/2}}-
1094: \frac{1}{\sqrt{J/2}}\right)^2\,,&\hspace{0.5cm}(I\,\mbox{even}, J\,\mbox{even})\\
1095: \left(\frac{1}{\sqrt{(I+1)/2}}-
1096: \frac{1}{\sqrt{(J+1)/2}}\right)^2\,,&\hspace{0.5cm}(I\,\mbox{odd}, J\,\mbox{odd})\\
1097: \left(\frac{1}{\sqrt{(I+1)/2}}+
1098: \frac{1}{\sqrt{J/2}}\right)^2\,,&\hspace{0.5cm}(I\,\mbox{odd}, J\,\mbox{even})\\
1099: \left(\frac{1}{\sqrt{I/2}}+
1100: \frac{1}{\sqrt{(J+1)/2}}\right)^2\,,&\hspace{0.5cm}(I\,\mbox{even}, J\,\mbox{odd})\,\,.
1101: \end{array}\right.
1102: \eqe
1103: %*********
1104: For N=3 we have
1105: %******
1106: \eab
1107: \label{n3}
1108: m_{12}^2&=&m_{13}^2=\bar{m}_{12}^2=\bar{m}_{13}^2=e^2\,\frac{\Lambda_E^3}{2\pi T}\nonumber\\
1109: m_{23}^2&=&\bar{m}_{23}^2=4e^2\,\frac{\Lambda_E^3}{2\pi T}\ \ \ \ \ \mbox{or}\nonumber\\
1110: m_{12}^2&=&m_{23}^2=\bar{m}_{12}^2=\bar{m}_{23}^2=e^2\,\frac{\Lambda_E^3}{2\pi T}\nonumber\\
1111: m_{13}^2&=&\bar{m}_{13}^2=4e^2\,\frac{\Lambda_E^3}{2\pi T}\,.
1112: \eae
1113: %******
1114:
1115: \subsection{Thermodynamical self-consistency and
1116: $e(T)$ at one loop\label{TSCE}}
1117:
1118: The TLH modes $\delta a^k_\rho$ are thermal quasiparticle fluctuations on tree-level
1119: since their masses $m_{IJ}$ and $\bar{m}_{IJ}$, given in Eqs.\,(\ref{TLHmasses}) and (\ref{n3}),
1120: are $T$ dependent. Moreover, the ground-state pressure, given by $-V_E$, is linearly
1121: dependent on $T$, see Eq.\,(\ref{gspot}). Thermodynamical quantities such as the pressure,
1122: the energy density, or the
1123: entropy density are interrelated by Legendre transformations as
1124: derived from the partition function associated with
1125: the underlying SU(N) Yang-Mills Lagrangian. To assure
1126: that the same Legendre transformations are valid in the effective electric theory, where ground-state pressure
1127: and particle masses are temperature dependent, a condition for thermodynamic self-consistency needs
1128: to be imposed. In general, this condition assures that $T$-derivatives of quantities that enter the
1129: action {\sl density} of the effective theory (in our case the TLH masses and the ground-state pressure)
1130: cancel one another.
1131:
1132: Let us formulate this condition at one-loop accuracy. It is convenient to
1133: work with dimensionless quantities. The quantity $a_k$ (mass over temperature)
1134: is defined as
1135: %*************
1136: \eqb
1137: \label{dimlessdef}
1138: a_k\equiv c_k a\,\ \ \ \ \ \mbox{where} \ \ \ \ \ a\equiv e\sqrt{\frac{\Lambda_E^3}{2\pi T^3}}\,,
1139: \ \ \ \ (k=1,\cdots,\mbox{N}(\mbox{N}-1))\,,
1140: \eqe
1141: %*************
1142: and the coefficient $c_k$ is equal to the square root of one of the numbers
1143: appearing to the right of the curly bracket in Eq.\,(\ref{TLHmasses}) or, for N=3, to
1144: 1 or 2, compare with Eq.\,(\ref{n3}). Recalling
1145: our definition of a dimensionless temperature
1146: %********
1147: \eqb
1148: \label{lambdaE}
1149: \lambda_E\equiv \frac{2\pi T}{\La_E}\,,
1150: \eqe
1151: %********
1152: we have
1153: %*******
1154: \eqb
1155: \label{massP}
1156: a=2\pi e\lambda_E^{-3/2}\,.
1157: \eqe
1158: %*******
1159: Let us also define the (negative definite) function $\bar{P}(a)$ as
1160: %*********
1161: \eqb
1162: \label{P(a)}
1163: \bar{P}(a)\equiv\int_0^\infty dx \,x^2\,\log[1-\exp(-\sqrt{x^2+a^2})]\,.
1164: \eqe
1165: %*********
1166: Ignoring higher loop corrections, the total thermodynamical pressure
1167: $P(\lambda_E)$ associated with the diagrams in
1168: Fig.\,(\ref{diagpress}) calculates as
1169: %********
1170: \eqb
1171: \label{treelP}
1172: P(\lambda_E)=-\La_E^4\left\{\frac{2\lambda_E^4}{(2\pi)^6}\left[2(\mbox{N}-1)\bar{P}(0)+
1173: 3\sum_{k=1}^{\tiny\mbox{N(N-1)}}\bar{P}(a_k)\right]+
1174: \frac{\lambda_E}{2}\left(\frac{\mbox{N}}{2}+1\right)\mbox{N}\right\}\,\ \ (\mbox{N\ \ even})\,.
1175: \eqe
1176: %*******
1177: In Eq.\,(\ref{treelP}) we have neglected the `nonthermal' contribution $-\Delta V_E$ to the
1178: one-loop bubbles in Fig.\,\ref{diagpress} which is estimated as (quantum fluctuations are cut
1179: off at the compositeness scale $|\phi|$)
1180: %*********
1181: \eqb
1182: \label{1loopvacbubble}
1183: \Delta V_E<(\mbox{N}^2-1)\frac{3}{8\pi^2}\int_0^{|\phi|}dp\, p^3\log\left(\frac{p}{|\phi|}\right)=
1184: -3(\mbox{N}^2-1)\frac{|\phi|^4}{128\pi^2}\,.
1185: \eqe
1186: %********
1187: To neglect $-\Delta V_E$ is well justified for sufficiently small N, say N$<$20, since we have
1188: %**********
1189: \eqb
1190: \label{DeltaV}
1191: \left|\frac{\Delta V_E}{V_E}\right|<
1192: (\mbox{N}^2-1)\frac{3}{128\pi^2}\left(\frac{|\phi|}{\La_E}\right)^6\sim
1193: (\mbox{N}^2-1)\frac{3}{128\pi^2}\,\lambda_E^{-3}\,,
1194: \eqe
1195: %**********
1196: and we will show later that the minimal temperature in the electric phase
1197: $\lambda_{c,E}$ is much larger than unity.
1198:
1199: For N=3 we obtain
1200: %********
1201: \eqb
1202: \label{treelP3}
1203: P(\lambda_E)=-\La_E^4\left\{\frac{2\lambda_E^4}{(2\pi)^6}\left[4\bar{P}(0)+3\left(
1204: 4\bar{P}(a)+2\bar{P}(2a)\right)\right]+
1205: 2\lambda_E\right\}\,\ \ \ (\mbox{N=3})\,.
1206: \eqe
1207: %*******
1208: %***********************
1209: \begin{figure}
1210: \begin{center}
1211: \leavevmode
1212: %\epsfxsize=9.cm
1213: \leavevmode
1214: %\epsffile[80 25 534 344]{}
1215: \vspace{2.5cm}
1216: \special{psfile=Fig-4a.ps angle=0 voffset=-75
1217: hoffset=-100 hscale=60 vscale=60}
1218: \end{center}
1219: \caption{Diagrams contributing to the pressure when radiative corrections are
1220: ignored. A thick line corresponds to TLH modes and a
1221: thin one to TLM modes. The cross depicts the ground-state contribution arising
1222: from caloron `condensation'.
1223: \label{diagpress}}
1224: \end{figure}
1225: %************************
1226: A particular Legendre transformation following from the partition function of the underlying
1227: theory maps pressure into energy density as
1228: %*********
1229: \eqb
1230: \label{rhoPree}
1231: \rho=T\frac{dP}{dT}-P\,.
1232: \eqe
1233: %*********
1234: For Eq.\,(\ref{rhoPree}) to hold also in the effective electric theory the following situation
1235: has to be {\sl arranged for}: only the explicit
1236: T-dependence in $P$, arising from the explicit T-dependence of the
1237: Boltzmann weight, should contribute to the derivative $\frac{dP}{dT}$ while implicit $T$ dependences
1238: of gauge-boson masses and the ground-state pressure ought to cancel one other.
1239: This condition is expressed as \cite{Gorenstein1995}
1240: %**************
1241: %*********
1242: \eqb
1243: \label{TSC}
1244: \pd_a P=0\,.
1245: \eqe
1246: %*********
1247: Before we proceed let us recall that the $\lambda_E$ dependence of the ground-state pressure, as indicated
1248: in Eqs.\,(\ref{gspot}) and (\ref{treelP}), could be expressed in terms of a dependence on the
1249: mass parameter $a$ by virtue of Eq.\,(\ref{massP}) if the $\lambda_E$ dependence of the gauge
1250: coupling constant $e$ was known. Keeping this in mind, we derive from Eqs.\,(\ref{TSC}), (\ref{treelP}), and
1251: (\ref{treelP3}) the following evolution equation
1252: %********
1253: \eqb
1254: \label{eeq}
1255: \pd_a \lambda_E=-\frac{24\,\lambda_E^4\,a}{(2\pi)^6
1256: \mbox{N(N+2)}}\sum_{k=1}^{\tiny\mbox{N(N-1)}}c_k^2 D(a_k)\,,\ \ (\mbox{N\ \ even})\,.
1257: \eqe
1258: %*********
1259: For N=3 we have
1260: %********
1261: \eqb
1262: \label{eeq3}
1263: \pd_a \lambda_E=-\frac{12\,\lambda_E^4\,a}{(2\pi)^6}
1264: \left(D(a)+2D(2a)\right)\,,\ \ (\mbox{N=3})\,.
1265: \eqe
1266: %*********
1267: The function $D(a)$ is defined as
1268: %**********
1269: \eqb
1270: \label{DA}
1271: D(a)\equiv \int_0^{\infty} dx\,
1272: \frac{x^2}{\sqrt{x^2+a^2}}\frac{1}{\exp(\sqrt{x^2+a^2})-1}\,.
1273: \eqe
1274: %**********
1275: Eqs.\,(\ref{eeq}) and (\ref{eeq3}) describe the evolution of temperature as a
1276: function of tree-level gauge boson mass. The right-hand sides of these equations
1277: are negative definite since the function $D(a)$ in Eq.\,(\ref{DA})
1278: is positive definite, see Fig.\,\ref{DAf}.
1279: %***********************
1280: \begin{figure}
1281: \begin{center}
1282: \leavevmode
1283: %\epsfxsize=9.cm
1284: \leavevmode
1285: %\epsffile[80 25 534 344]{}
1286: \vspace{4.5cm}
1287: \special{psfile=Fig-2.ps angle=0 voffset=-140
1288: hoffset=-120 hscale=50 vscale=40}
1289: \end{center}
1290: \caption{The function $a\,D(a)$.\label{DAf}}
1291: \end{figure}
1292: %************************
1293: As a consequence, the solutions $\lambda_E(a)$
1294: to Eqs.\,(\ref{eeq}) and (\ref{eeq3}) can be inverted to $a(\lambda_E)$. In Fig.\,\ref{laofa} a solution
1295: for N=2 subject to the initial condition $\lambda_{E,P}\equiv\lambda_E(a=0)=10^3$ is shown.
1296: We have noticed numerically that the low-temperature behavior of $\lambda_E(a)$
1297: is practically independent of the value $\lambda_{E,P}$ as long as $\lambda_{E,P}$ is
1298: sufficiently large. Let us show this analytically. For $a$ sufficiently smaller than unity we may
1299: expand the right-hand side of Eq.\,(\ref{eeq}) only taking the
1300: linear term in $a$ into account. The inverse of the solution is then of the following form
1301: %********
1302: \eqb
1303: \label{anasolv}
1304: a\propto \lambda_E^{-3/2}\sqrt{1-\left(\frac{\lambda_E}{\lambda_{E,P}}\right)^3}\,.
1305: \eqe
1306: %********
1307: For $\lambda_E$ sufficiently smaller than $\lambda_{E,P}$ this can be approximated as
1308: %********
1309: \eqb
1310: \label{anasolapp}
1311: a\propto\lambda_E^{-3/2}\,.
1312: \eqe
1313: %********
1314: The dependence in Eq.\,(\ref{anasolapp}) thus
1315: is an {\sl attractor}. So at whatever asymptotically high temperature the formation of the
1316: adjoint Higgs field $\phi$ out of noninteracting trivial-holonomy calorons
1317: is assumed does not influence the behavior of the theory at much
1318: lower temperatures. This result is reminiscent of the
1319: ultraviolet-infrared decoupling property of the renormalizable,
1320: underlying theory.
1321:
1322: Notice that Fig.\,(\ref{DAf}) and Eq.\,(\ref{eeq}) imply that there are
1323: fixed points of the evolution $\lambda_E(a)$ at $a=0$ and $a=\infty$.
1324: The points $\lambda_{E,P}\equiv\lambda_E(a=0)$ and $\lambda_{E,c}\equiv\lambda_E(a=\infty)$ are
1325: associated with the highest and the lowest attainable temperaturesin the electric phase, respectively.
1326: In a bottom-up evolution no information can be obtained about $\lambda_{E,P}$ if the temperature that is maximally
1327: reached is sufficiently smaller than $\lambda_{E,P}$.
1328: In a top-down evolution, where $\lambda_{E,P}$ is set as a boundary value, the prediction of
1329: $\lambda_{E,c}$ is independent of $\lambda_{E,P}$. These two
1330: statements are immediate consequences of the existence of an
1331: attractor in the thermodynamical evolution.
1332:
1333: Numerically inverting the solution $\lambda_E(a)$ to $a(\lambda_E)$, the evolution of the gauge coupling constant
1334: $e$ can be computed using Eq.\,(\ref{massP}):
1335: %*******
1336: \eqb
1337: \label{coupP}
1338: e(\lambda_E)=\frac{1}{2\pi}a(\lambda_E)\lambda_E^{3/2}\,.
1339: \eqe
1340: %*******
1341: We show the result in Fig.\,\ref{eoflam} for N=2,3. Before we interprete this result
1342: a remark on the interpretation of the effective gauge coupling constant $e$ for $T\sim T_P$ is in order.
1343: Since $e$ determines the
1344: strength of the interaction between nontopological gauge field fluctuations $\delta a_\rho$
1345: and the {\sl coherent} caloron state $\phi$ there is, in general, no reason for it
1346: to be equal to the gauge coupling constant $\bar{g}$ of the fundamental Yang-Mills theory.
1347: However, for temperatures very close to $T_P$, where $\phi$ is assumed to form
1348: (see also Sec.\,\ref{MPlanck}), the coupling constant $e$ should be roughly equal to $\bar{g}$.
1349: From Fig.\,\ref{eoflam} we see that the effective gauge coupling constant $e\sim\bar{g}$
1350: evolves to values larger than unity
1351: shortly below the initial temperature $\lambda_{E,P}$. This is in agreement
1352: with our assumption that trivial-holonmy SU(2) {\sl calorons} (of sufficiently small `instanton' radius)
1353: have a large action and thus contribute sizably to the partion function of the underlying theory,
1354: compare with Eq.\,(\ref{GPYSeffth}). For a grandly unifying SU(N) gauge theory we argue
1355: in Sec.\,\ref{MPlanck} that $\lambda_{E,P}\sim \frac{M_P}{\La_E}$ where $M_P\gg \La_E$ denotes the
1356: cutoff scale for the description in terms of a local, four-dimensional field theory.
1357:
1358: There is no handle on the relation between the fundamental gauge coupling
1359: $\bar{g}$ and $e$ after the condensate has formed and is sustained by interactions
1360: between the trivial-holonomy calorons. Since collective effects are strong
1361: they can make up for a large value of the action of an isolated caloron generated by a
1362: small value of $\bar{g}$. This may justify the use of universal
1363: perturbative expressions for the beta function in lattice
1364: simulation at low temperatures.
1365:
1366: Notice that the dependence of $a$ on $\lambda_E$ in
1367: Eq.\,(\ref{anasolapp}) is canceled in the dependence of $e$ on $\lambda_E$
1368: such that a plateau is reached quickly in Eq.\,(\ref{coupP}). We interprete the fact that the
1369: gauge coupling constant $e$ remains constant for a large range of temperatures
1370: as another indication for the existence of spatially isolated and conserved
1371: magnetic charges in the system, see also Sec.\,\ref{CMP}.
1372: %***********************
1373: \begin{figure}
1374: \begin{center}
1375: \leavevmode
1376: %\epsfxsize=9.cm
1377: \leavevmode
1378: %\epsffile[80 25 534 344]{}
1379: \vspace{4.5cm}
1380: \special{psfile=Fig-4.ps angle=0 voffset=-150
1381: hoffset=-130 hscale=50 vscale=50}
1382: \end{center}
1383: \caption{The solution $\lambda_E(a)$ to Eq.\,(48) for N=2 subject to the
1384: boundary condition $\lambda_{E,P}\equiv\lambda_E(a=0)=10^3$.\label{laofa}}
1385: \end{figure}
1386: %************************
1387: During the relaxation of $e$ to its plateau value
1388: constituent BPS magnetic monopoles residing in dissociating nontrivial-holonomy
1389: calorons form as isolated defects
1390: \cite{KraanVanBaalNPB1998,vanBaalKraalPLB1998,Diakonov}. Since the
1391: interaction between a monopole and an antimonopole, as mediated by TLM modes,
1392: is screened \cite{KorthalsAltes,HoelbingRebbiRubakov2001} these defects need
1393: not be considered explicitly in the effective, thermal theory discussed in the present work. Implicitly,
1394: their presence is accounted for here by the holonomy of the
1395: background field $a^{bg}_\rho$.
1396:
1397: The effective gauge coupling constant
1398: $e$ runs into a logarithmic (needle) pole at $\lambda_{E,c}$ of the form
1399: %********
1400: \eqb
1401: \label{logpolee}
1402: e(\lambda_E)\propto-\log(\lambda_E-\lambda_{E,c})\,,
1403: \eqe
1404: %********
1405: compare with Fig.\,\ref{laofa}.
1406:
1407: The plateau values for $e$ are $e\sim 5.1$ and $e\sim 4.2$
1408: for N=2 and N=3, respectively. For a given N they do not depend on where the
1409: boundary condition $\lambda_{E,P}\equiv\lambda_E(a=0)$ is set if $\lambda_{E,P}$ is sufficiently larger than
1410: $\lambda_{E,c}$. For N=2 we have $\lambda_{E,c}=11.65$ and for N=3 we have $\lambda_{E,c}=8.08$.
1411: It thus is self-consistently justified to neglect the one-loop quantum corrections
1412: to $V_E(\phi)$ as they are estimated in Eq.\,(\ref{DeltaV}).
1413: %***********************
1414: \begin{figure}
1415: \begin{center}
1416: \leavevmode
1417: %\epsfxsize=9.cm
1418: \leavevmode
1419: %\epsffile[80 25 534 344]{}
1420: \vspace{5.5cm}
1421: \special{psfile=Fig-5.ps angle=0 voffset=-170
1422: hoffset=-130 hscale=60 vscale=60}
1423: \end{center}
1424: \caption{The low-temperature evolution of the gauge
1425: coupling $e$ in the electric phase for
1426: N=2 (grey line) and N=3 (black line). The gauge coupling
1427: diverges logarithmically, $e\propto -\log(\lambda_{E}-\lambda_{c,E})$, at
1428: $\lambda_{E,c}=11.65\ (\mbox{N=2})$ and $\lambda_{E,c}=8.08\ (\mbox{N=3})$.
1429: The plateau values are $e=5.1\ (\mbox{N=2})$ and $e=4.2\ (\mbox{N=3})$.\label{eoflam}}
1430: \end{figure}
1431: %************************
1432: Naively, one would conclude that the large plateau values would render the
1433: fluctuations $\delta a_\rho$ to be very strongly coupled and that radiative
1434: corrections to the thermodynamical potentials would thus be uncontrolled. This, however,
1435: does not happen due to the fact that the TLH modes acquire masses by the Higgs mechanism
1436: which are proportional to $e$ and due to the existence of a compsiteness scale
1437: $|\phi|$. The latter constrains the momentum $p$ of quantum fluctuations
1438: in $\delta a_\rho$ as
1439: %********
1440: \eab
1441: \label{qfconstr}
1442: |p^2-m^2_k|&\le &|\tilde{\phi}_l|^2\,,\ \ (\mbox{TLH, Minkowskian})\,,\ \
1443: |p^2|\le |\tilde{\phi}_l|^2\,, \ \ (\mbox{TLM, Minkowskian})\,,\nonumber\\
1444: \ p_e^2+m_k^2&\le&|\tilde{\phi}_l|^2\,,\ \ (\mbox{TLH, Euclidean})\,, \ \
1445: p_e^2\le |\tilde{\phi}_l|^2\,, \ \ (\mbox{TLM, Euclidean})\,.
1446: \eae
1447: %********
1448: Since in nonlocal two-loop contributions, see Fig.\,\ref{looppress},
1449: each TLH line is on-shell for $e>1$ the effect
1450: of a strongly coupled vertex is compensated by the very
1451: small phase space that is allowed for the progagation of
1452: a TLM mode coupling to the TLH mode. The center-of-mass energy flowing into or out of a four-vertex
1453: is constrained in addition to Eq.\,(\ref{qfconstr})
1454: to be smaller than $|\tilde{\phi}_l|$. We show in \cite{HerbstHofmannRohrer2004}
1455: that the two-loop contributions to
1456: the pressure for SU(2) are, depending on temperature, at most $\sim 0.1$\%
1457: of the one-loop result.
1458:
1459: \subsection{What is $T_P$?\label{MPlanck}}
1460:
1461: At temperatures
1462: larger than the highest attainable temperature $T_P$ in the electric phase (corresponding to $a=0$)
1463: a grandly unifying SU(N) gauge symmetry, which generates all matter and its (nongravitational)
1464: interactions in the Universe at lower temperatures, would be unbroken. We assume here
1465: that gravity is a perfectly classical theory up to the Planck
1466: mass $M_P$ \footnote{This assumption is usually made
1467: in field-theory models of cosmological
1468: inflation.}. The perturbative phase of the SU(N) Yang-Mills theory at
1469: $T>T_P$ would have a trivial vacuum state represented by weakly interacting
1470: quantum fluctuations. The momenta associated with these
1471: fluctuations can be maximally as hard as some cutoff scale at which the
1472: four-dimensional setup ceases to be reliable. Common belief is that this cutoff scale is
1473: $M_P$.
1474:
1475: The highest temperature $T_{\tiny\mbox{cutoff}}$
1476: attainable in the perturbative phase thus is comparable to $M_P$, $T_{\tiny\mbox{cutoff}}\sim M_P$.
1477: At temperatures rnaging between $T_{\tiny\mbox{cutoff}}$ and $T_P$ perturbative
1478: vacuum fluctuations would generate a cosmological
1479: constant $\La_{\tiny\mbox{cosmo}}$ given as
1480: %*********
1481: \eqb
1482: \label{wbLambdacosmo}
1483: \La_{\tiny\mbox{cosmo}}\sim M_P^4\,.
1484: \eqe
1485: %**********
1486: At $T_{\tiny\mbox{cutoff}}$ the vacuum energy density $M_P^4$
1487: would be comparable to the thermal energy density of on-shell fluctuations
1488: $\sim T_{\tiny\mbox{cutoff}}^4$. While the former is constant
1489: the later dies off quickly as the Universe cools down. Thus for $T$
1490: slightly smaller than $T_{\tiny\mbox{cutoff}}$ the vacuum energy density dominate the expansion hence the
1491: Universe would rapidly decrease its temperature as
1492: %*********
1493: \eqb
1494: \label{decT}
1495: T\sim \exp[-M_P \Delta t] T_{\tiny\mbox{cutoff}}\,
1496: \eqe
1497: %*********
1498: where $\Delta t\equiv t-t_P$ and $t_P\sim M^{-1}_P$. A sudden termination
1499: of this Planck-scale inflation would occur at $T=T_P$ where
1500: the field $\phi$ comes into existence. While the {\sl radiation} component of the total
1501: energy density is continuous across the phase boundary
1502: at $T_P$ the energy density of the {\sl ground-state}
1503: would be discontinuously reduced from $M_P^4$ to
1504: $\sim T_P \La_{YM,\tiny\mbox{N}}^3$, see Eq.(\ref{gspot}). On the one hand,
1505: this release of latent heat is a characteristic for a (strong if $T_P\ll T_{\tiny\mbox{cutoff}}$)
1506: 1$^{\tiny\mbox{st}}$ order transition. On the other hand, the order parameter $a$ for the onset of the electric
1507: phase is continuous, see Fig.\,(\ref{eoflam}) (screening masses are comparable
1508: on both sides of the phase boundary, see Eq.\,(\ref{thermPiweakcoupl})).
1509: But this is signalling a 2$^{\tiny\mbox{nd}}$ order phase transition. There is only one way
1510: to avoid this
1511: contradiction: The phase boundary at
1512: $T_P$ needs to be hidden beyond the point
1513: $T_{\tiny\mbox{cutoff}}\sim M_P$, that is, $T_P\ge T_{\tiny\mbox{cutoff}}$.
1514:
1515: The reader may
1516: object that our conclusion about the 2$^{\tiny\mbox{nd}}$ order
1517: of the phase transition (order parameter $a$) is resting on a
1518: one-loop analysis of the gauge-coupling evolution. Usually,
1519: it is understood that such a mean-field treatment breaks down close
1520: to a 2$^{\tiny\mbox{nd}}$ order transition due to fact that long range
1521: correlations mediated by low-momentum quantum fluctuations become important.
1522: In the electric phase these long-range correlations are, however,
1523: contained in the field $\phi$ which does not fluctuate at
1524: any temperature $T_{E,c}<T\le T_P$. Due to $|\phi|$ being a
1525: cutoff for the quantum fluctuations of the TLM
1526: and TLH modes and due to the fact that $|\phi|$ dies off as $T^{-1/2}$ the long-range correlating effects of
1527: TLM and TLH quantum fluctuations can safely be
1528: neglected if $T_P\gg\La_E$. The above discussion and conclusion thus are valid.
1529:
1530: So far we had in mind the simplified case of MGSB at $T_P$ in
1531: grandly unifying SU(N) Yang-Mills theory. We know from experiment, however,
1532: that gauge-symmetry breakdown at $T_P$ is {\sl not} maximal in Nature. If the gauge-symmetry breaking
1533: by $\phi$ at $T_P$ is submaximal, however, the
1534: same contradiction between the orders of the
1535: phase transition arises for $T_P<M_P$. In this case a
1536: fundamental gauge symmetry SU(N) would be broken
1537: to a product of group with factors SU(M) M$<$N or U(1). Recall,
1538: that submaximal gauge-symmetry breaking in the electric phase takes
1539: place if SU(2) blocks in $\phi$ with equal winding number
1540: are generated at $T_P$, see Eq.\,(\ref{SU2diag}).
1541:
1542: The theory would then condense magnetic SU(M) color and magnetic U(1) monopoles at the
1543: temperature $T_{E,c}$. While condensates of the latter are described by
1544: complex scalar fields, see Sec.\,\ref{magPT}, condensates of the
1545: former are, again, described by adjoint Higgs fields. Maximal or submaximal
1546: breakings of the residual SU(M) gauge symmetries would be possible at the electric-magnetic
1547: transition of the fundamental SU(N) theory. For the effective
1548: description of SU(M) thermodynamics at $T<T_{E,c}$ this
1549: boundary condition effectively is set during the phase transition at $T_{E,c}$ and not at $T_P$.
1550: By matching the thermodynamical pressures at $T_{E,c}$ the scale $\La_E^\prime$ of the
1551: effective theory SU(M) is determined in terms of the scale $\La_E$
1552: of the fundamental SU(N) theory and the pattern of symmetry breaking at $T_P$ and $T_{E,c}$.
1553: After a sequence of such matching procedures has taken place (in which residual U(1)
1554: factors have thermodynamically decoupled) a hierarchy
1555: between $\La_E$ and the scale $\La_E^{\prime\cdots\prime}$ of an effective
1556: SU(L) theory ($L\ll N$), seen experimentally at low energies (or local
1557: temperatures for that matter), is generated.
1558:
1559: To summarize,
1560: we provided an argument that in a grandly unifying and four-dimensional SU(N) Yang-Mills theory the dynamical
1561: generation of the adjoint background field $\phi$ must
1562: take place at a temperature $T_{\tiny\mbox{cutoff}}\sim M_P$ where the local
1563: field-theory description breaks down. Any lower SU(L) gauge symmetry, which is
1564: generated from the SU(N) theory by a sequence of condensations of (color) monopoles and confining
1565: transitions, is matched to its `predessesor' theory in
1566: 2$^{\tiny\mbox{nd}}$ order like phase transitions. The scale of this SU(L) gauge theory can be much
1567: lower than the scale $\Lambda_E$ of the fundamental SU(N) theory.
1568:
1569:
1570: \subsection{Stable und unstable magnetic monopoles\label{CMP}}
1571:
1572: Due to the presence of an adjoint Higgs field $\phi$
1573: in the electric phase there are 't Hooft-Polyakov magnetic monopoles
1574: \cite{'tHooft1974,Polayakov1974} which are centered at
1575: the isolated zeros of $\phi$. On a mesoscopic level, these zeros occur at points in space
1576: where four or more color-orientation domains of a
1577: given block $\tilde{\phi}_l$ meet \cite{Kibble1976}. Microscopically,
1578: BPS monopoles \cite{PrasadSommerfield1974} are contained within
1579: decaying nontrivial-holonomy calorons \footnote{A perturbative analysis of 1-loop radiative corrections to an isolated
1580: instanton were performed in \cite{'tHooft1976}. In \cite{Diakonov} this was done for
1581: the nontrivial-holonomy caloron. As a result a repulsive potential for the constituent monopole and
1582: antimonopole was obtained for a sufficiently large holonomy.}.
1583:
1584: Close to $T_P$ we have $\bar{g}\sim e$, and the
1585: following processes take place:
1586: Trivial-holonomy calorons grow rapidly in size,
1587: start to overlap, and thus generate calorons with holonomy by their
1588: interactions. If this holonomy is sufficiently large then two following processes take place: (i)
1589: SU(2) nontrivial-holonomy calorons of the same embedding in SU(N) decay independently into
1590: their constituent magnetic monopoles and antimonopoles, and (ii) SU(2)
1591: nontrivial-holonomy calorons generated from trivial-holonomy calorons
1592: of different SU(2) embeddings\footnote{Recall, that we have assumed that these calorons
1593: come with different topological charge to
1594: obtain maximal symmetry breaking.} in SU(N) do not decay into constituent
1595: monopoles and antimonopoles since they would have to live in instable
1596: superpositions of the embeddings of the asymptotic
1597: trivial-holonomy calorons. While the former process generates stable magnetic
1598: dipoles the latter generates instable monopoles and antimonopoles.
1599:
1600: In our macroscopic approach it is hard to see how the size of a typical trivial-holonomy
1601: caloron changes with temperature after $e$ has reached its plateau value
1602: since $e$ plays a different role than the fundamental coupling constant
1603: $\bar{g}$. We may, however, infer from lattice simulations that trivial-holonomy calorons
1604: are large enough to not generate a topological susceptibility on lattices of
1605: presently feasible sizes.
1606:
1607: A magnetic monopole-antimonopole pair, which is connected by a magnetic flux line, becomes a stable dipole
1608: if the pair has a sufficiently large spatial separation.
1609: In this case the monopole and the antimonopole are practically
1610: noninteracting \cite{HoelbingRebbiRubakov2001} and
1611: thus are stable defects. If a single monopole is produced then it
1612: is unstable unless it connects with its antimonopole, produced in
1613: an independent collision. N$-$1 independent SU(2) subgroups exist and so
1614: N$-$1 independent magnetic monopoles may occur in the case of MGSB.
1615: Since the monopole constituents in a caloron are BPS
1616: saturated we also expect an isolated monopole in a stable monopole-antimonopole pair
1617: to be BPS saturated. The analytical expression
1618: for an SU(2) charge-one BPS monopole in a gauge where the Higgs field $\phi$
1619: winds around the group manifold at spatial infinity \cite{PrasadSommerfield1974} is given as
1620: %**********
1621: \eqb
1622: \label{BPSmonop}
1623: A^a_0=0\,,\ \ \ \ A^a_i=\epsilon_{aij}\hat{r}_j\frac{1-K(r)}{er}\,,\ \ \ \ \phi^a=\hat{r}_a \frac{H(r)}{er}\,,
1624: \eqe
1625: %**********
1626: where $r\equiv \sqrt{\vec{r}^2}$, and $\hat{r}$ is a spatial unit vector.
1627: The form of the functions $K(r)$ and $H(r)$ is
1628: %********
1629: \eqb
1630: \label{KandH}
1631: K(r)=\frac{Cr}{\sinh(Cr)}\,,\ \ \ \ \ \ \ \ \ H(r)=Cr\coth(Cr)-1\,.
1632: \eqe
1633: %********
1634: In Eqs.\,(\ref{KandH}) the mass scale $C$ is proportional to the asymptotic Higgs
1635: modulus $|\phi(|\vec x|\to\infty)|$ and the gauge coupling constant $e$.
1636: The mass of a BPS monopole is given as \cite{PrasadSommerfield1974}
1637: %**********
1638: \eqb
1639: \label{monopmass}
1640: M=\frac{8\pi}{\sqrt{2}\,e} |\phi(|\vec x|\to\infty)|\,.
1641: \eqe
1642: %***********
1643: A dual, abelian field strength $\tilde{G}_{\mu\nu}$ can be defined as
1644: \cite{'tHooft1974}
1645: %********
1646: \eqb
1647: \label{tT}
1648: \tilde{G}_{\mu\nu}=\frac{\phi^a G_{\mu\nu}^a}{|\phi|}-
1649: \frac{\epsilon_{abc}}{e|\phi|^3}\phi_a ({\cal D}_\mu \phi)_b ({\cal D}_\mu \phi)_c\,.
1650: \eqe
1651: %***********
1652: The expression in Eq.\,(\ref{tT}) reduces to $\tilde{G}_{\mu\nu}=\pd_\mu a^3_\nu-\pd_\nu a^3_\mu$
1653: in unitary gauge $\phi^a=\delta^{a3}|\phi|$. Eq.\,(\ref{tT}) defines the field strength of a
1654: dual photon which couples
1655: to the magnetic charge $4\pi/e$ of the monopole.
1656: Both the gauge dynamics involving only dual photons and magnetic monopoles and
1657: the entire gauge dynamics in the electric phase are blind with respect to the
1658: magnetic (local in space) center symmetry $Z_{\tiny\mbox{N,mag}}$
1659: (the field strength ${G}_{\mu\nu}$, the field $\phi$ and the covariant derivative
1660: ${\cal D}_\mu \phi$ are invariant under center transformations and,
1661: as a consequence of Eq.\,(\ref{tT}) so is the dual field strength
1662: $\tilde{G}_{\mu\nu}$). However, a local-in-time
1663: transformation $\in Z_{\tiny\mbox{N,elec}}$ may transform the Dirac string between
1664: a static monopole and a static antimonopole in a given SU(2) embedding
1665: into a Dirac string belonging to a dipole in a different SU(2)
1666: embedding. This does not violate the
1667: conservation of total magnetic charge and certainly has no effect
1668: on any gauge invariant quantity. In an effective theory,
1669: where monopoles are condensed degrees of freedom, a local $Z_{\tiny\mbox{N,elec}}$
1670: transformation should thus be represented by a local permutation of the
1671: fields describing the monopole condensates and the gauge-field fluctuations
1672: which couple to them. In such an effective theory the action thus ought to be
1673: invariant under these local permutations.
1674:
1675: How can we see the occurrence of stable and unstable magnetic monopoles in the macroscopic,
1676: effective theory for the electric phase? In winding gauge the temporal winding
1677: of the $l^{\tiny\mbox{th}}$ SU(2) block in $\phi$
1678: is complemented by spatial winding at isolated points in 3D space. By a large, $\tau$
1679: dependent gauge transformation the monopole's spatially asymptotic SU(2)
1680: Higgs field is rotated to spatial constancy and into the direction given
1681: by the temporal winding of the unperturbed block $\tilde{\phi}_l$. Its
1682: Dirac string rotates as a function of Euclidean time. Since this monopole is stable,
1683: a correlated antimonopole must exist. We arrive at a dipole rotating
1684: about its center of mass at an
1685: angular frequency $2\pi l T$. This rotation is an
1686: artifact of our choice of gauge. Rotating the dipole to unitary gauge by the gauge
1687: function $\theta_l$ in Eq.\,(\ref{thetamat}), we arrive
1688: at a (quasi)static dipole. There are isolated
1689: coincidence points (CPs) in time where the lower right (upper left) corner
1690: of the $l^{\tiny\mbox{th}}$ ($(l+1)^{\tiny\mbox{th}}$)
1691: SU(2) block (now ($l=1,\cdots,\mbox{N}/2-1$))
1692: together with the number zero of its
1693: right-hand (left-hand) neighbour are proportional to the generator $\lambda_3$.
1694: Coincidence also takes place between the first and last diagonal entry in $\phi$.
1695: At a CP, spatial winding may take place at isolated points in 3D space. Moreover,
1696: coincidence also takes place between nonadjacent
1697: SU(2) blocks and the first and last diagonal entry in $\phi$. The associated
1698: monopoles are, however, not independent. The spatial winding associated with the additional
1699: SU(2) generators `flashing out' at the CPs corresponds to unstable magnetic monopoles.
1700:
1701: Summing up all independent monopole species, we have:
1702: %*********
1703: \eqb
1704: \label{counting}
1705: \frac{\mbox{N}}{2}+\frac{\mbox{N}}{2}-1=\mbox{N}-1\,.
1706: \eqe
1707: %*********
1708: In the case N=3 the field $\phi$ winds with winding number one in each of the
1709: two independent SU(2) subalgebras for half the time, see Sec\,\ref{oddN}. The match between these
1710: subalgebras happens at the CPs $\tau_{CP}=0,1/(2T)$ where an element of the
1711: third, dependent SU(2) algebra is generated, see Sec.\,\ref{oddN}. Due to these CPs we have
1712: unstable monopoles.
1713:
1714: \subsection{Outlook on radiative corrections\label{Radcor}}
1715:
1716: \subsubsection{Contributions to the TLM self-energy}
1717:
1718: Let us now investigate for N=2 and at one loop the simplest contribution to
1719: the polarization tensor for the TLM mode. A complete investigation of
1720: two-loop contributions to the pressure is the objective of \cite{HerbstHofmannRohrer2004}. We work in unitary
1721: gauge $\phi=\mbox{diag}(\phi_1,\phi_2)$, $a_\rho^{bg}=0$.
1722: This condition fixes the gauge up to U(1) rotations
1723: generated by $\lambda_3$. This remaining gauge freedom can be used to gauge the TLM mode to
1724: transversality: $\pd_i \delta a^{\tiny\mbox{TLM}}_i=0$ (radiation or Coulomb gauge).
1725: No ghost fields need to be introduced
1726: in unitary-Coulomb gauge.
1727:
1728: After an analytical continuation to Minkowskian signature\footnote{For the purpose of the present
1729: work we do not need the matrix formulation of the real-time propagators.}
1730: the asymptotic propagator of a free TLM mode is given as \cite{LandsmanWeert1987}
1731: %**************
1732: \eqb
1733: \label{TLMprop}
1734: D^{\tiny\mbox{TLM},0}_{\mu\nu,ab}(k,T)=-\delta_{ab}P^T_{\mu\nu}
1735: \left(\frac{i}{k^2}+2\pi\delta(k^2)n_B(|k_0|/T)\right)\,
1736: \eqe
1737: %**************
1738: where
1739: %***********
1740: \eab
1741: \label{PT}
1742: P^T_{00}&=&P^T_{0i}=P^T_{i0}=0\,,\nonumber\\
1743: P^T_{ij}&=&\delta_{ij}-\frac{k_ik_j}{\vec{k}^2}\,,
1744: \eae
1745: %***********
1746: and $n_B(x)\equiv\frac{1}{\exp[x]-1}$ denotes the Bose distribution function.
1747: The analytically continued asymptotic propagator of
1748: a free TLH mode $D^{\tiny\mbox{TLH},IJ,0}_{\mu\nu,ab}(k,T)$
1749: is that of a massive vector boson
1750: %***********
1751: \eqb
1752: \label{TLHprop}
1753: D^{\tiny\mbox{TLH},IJ,0}_{\mu\nu,ab}(k,T)=-\delta_{ab}
1754: \left(g_{\mu\nu}-\frac{k_\mu k_\nu}{m_{IJ}^2}\right)
1755: \left[\frac{i}{k^2-m_{IJ}^2}+2\pi\delta(k^2-m_{IJ}^2)n_B(|k_0|/T)\right]\,.
1756: \eqe
1757: %**********
1758: The vertices for the interactions of TLH and TLM modes are the usual ones.
1759: In unitary-Coulomb gauge the 4D loop integrals over quantum fluctuations
1760: are cut off at the compositeness scale $|\phi|(T)$ of the effective theory.
1761: Thermal fluctuations, associated with 3D loop integrals, are automatically cut off
1762: by the distribution function $n_B$.
1763: %***********************
1764: \begin{figure}
1765: \begin{center}
1766: \leavevmode
1767: %\epsfxsize=9.cm
1768: \leavevmode
1769: %\epsffile[80 25 534 344]{}
1770: \vspace{3.5cm}
1771: \special{psfile=Fig-6.ps angle=0 voffset=-90
1772: hoffset=-130 hscale=60 vscale=60}
1773: \end{center}
1774: \caption{A tadpole contribution to the self-energy of the TLH mode.\label{tadpole}}
1775: \end{figure}
1776: %************************
1777: Let us now look at the tadpole contribution to $\Pi_{a=3,\mu\rho}$ as shown in Fig.\,\ref{tadpole}.
1778: This diagram decomposes into a part for the vacuum fluctuations in the loop,
1779: which has a $T$ dependence only due to the $T$ dependence of TLH masses,
1780: and a thermal part. Contracting the Lorentz indices, the former can be calculated as
1781: %***********
1782: \eqb
1783: \label{vacPi}
1784: \Pi^{\tiny\mbox{vac},\mu}_{a=3,\mu}=-\frac{3(e|\tilde{\phi}|)^2}{2\pi^2}\int_0^{\sqrt{1-(2e)^2}}dx\,
1785: \frac{x^3(4+\frac{x^2}{(2e)^2})}{x^2+(2e)^2}\,
1786: \eqe
1787: %***********
1788: while the thermal part reads
1789: %**********
1790: \eqb
1791: \label{thermPi}
1792: \Pi^{\tiny\mbox{therm},\mu}_{a=3,\mu}=\frac{18}{\pi^2}(e|\tilde{\phi}|)^2\int_0^\infty dy\,
1793: \frac{y^2}{\sqrt{y^2+(2e)^2}}\,\,\frac{1}{\exp\left[2\pi\lambda_E^{-3/2}\sqrt{y^2+(2e)^2}\right]-1}\,.
1794: \eqe
1795: %************
1796: It is instructive to perform the weak and strong coupling limits
1797: in Eqs.\,(\ref{vacPi}) and (\ref{thermPi}).
1798:
1799: \noindent For $e<\frac{1}{\sqrt{2}}$, we obtain
1800: %***********
1801: \eqb
1802: \label{vacPiweakcoupl}
1803: \Pi^{\tiny\mbox{vac},\mu}_{a=3,\mu}=-\frac{3|\tilde{\phi}|^2}{32\pi^2}\left(1+16\,e^2-
1804: 80\left(1-\frac{12}{5}\log(2e)\right)\,e^4\right)\,
1805: \eqe
1806: %***********
1807: and for $e\ll 1$
1808: %**********
1809: \eqb
1810: \label{thermPiweakcoupl}
1811: \Pi^{\tiny\mbox{therm},\mu}_{a=3,\mu}\to 3\pi^2\,(eT)^2+O(e^4)\,.
1812: \eqe
1813: %************
1814: For $e\gg \frac{1}{\sqrt{2}}$ there is no vacuum contribution, $\Pi^{\tiny\mbox{vac},\mu}_{a=3,\mu}=0$.
1815: The thermal part reads
1816: %**********
1817: \eqb
1818: \label{thermstrongcoupl}
1819: \Pi^{\tiny\mbox{therm},\mu}_{a=3,\mu}\to\frac{18}{\pi^3}\,e^3\,\Lambda_E^2\,\lambda_E^{1/2}\,
1820: K_1(4\pi\,e\,\lambda_E^{-3/2})\,
1821: \eqe
1822: %************
1823: where $K_1(x)$ denotes a modified Bessel function.
1824: The weak coupling result for the thermal part in Eq.\,(\ref{thermPiweakcoupl}) coincides,
1825: up to a numerical factor, with the perturbative expression for the electric
1826: screening (or Debye) mass-squared, as it should. In the limit of infinite coupling,
1827: which is reached due to the logarithmic pole for
1828: $\lambda_E\searrow\lambda_{E,c}$, see Eq.\,(\ref{logpolee}),
1829: the thermal part in Eq.\,(\ref{thermstrongcoupl}) vanishes.
1830: This agrees qualitatively with results obtained in
1831: thermal quasiparticle models fitted to lattice data
1832: \cite{Biro1990,Peshier1995,LevaiHeinz1998}. It was found in these models that the Debye mass
1833: vanishes for $T\searrow T_{E,c}$ \footnote{We foretake at this point that the deconfinement phase transiiton
1834: seen on the lattice is the electric-magnetic transition at $T_{E,c}$. We will discuss
1835: in Sec.\,\ref{complat} why the lattice is not capable of measuring infrared
1836: sensitive quantities such as the pressure at temperatures below $T_{E,c}$.}.
1837: A large and {\sl constant} value of $e$, as it is generated
1838: by one-loop evolution (compare with Eqs.\,(\ref{anasolapp}) and (\ref{coupP})),
1839: implies that the approximation leading to Eq.\,(\ref{thermstrongcoupl}) breaks down for
1840: high temperatures. It is, however, clear from Eq.\,(\ref{thermPi}) that the
1841: weak coupling result at $O(e^2)$ in Eq.\,(\ref{thermPiweakcoupl})
1842: gives an upper bound on the contribution to the screening mass-squared at any
1843: temperature and any value of the
1844: coupling constant.
1845:
1846: We expect that the situation is similar for the nonlocal one-loop diagrams.
1847: As for the tadpole correction in
1848: the polarization operator of a TLH
1849: mode there is a contribution $\propto e^2$ for strong coupling which arises from the
1850: vacuum part with the TLM mode in the loop. This contribution is, however,
1851: suppressed due to the constraint that the center-of-mass energy flowing into or out of
1852: the vertex must be smaller than $|\phi|$ \cite{HerbstHofmannRohrer2004}.
1853:
1854:
1855: \subsubsection{Loop expansion of the pressure}
1856:
1857: Two-loop diagrams contributing to the pressure in a real-time formulation
1858: are indicated in Fig.\,\ref{looppress}. We do not compute them here but in
1859: \cite{HerbstHofmannRohrer2004} for N=2.
1860: A general remark concerning thermodynamical self-consistency
1861: is in order already here. Recall, that on one-loop level
1862: we have obtained an evolution equation from the requirement of thermal
1863: self-consistency $\pd_a P=0$. This gave a functional relation between temperature and mass
1864: which could be inverted for all temperatures in the electric phase.
1865: After the relation Eq.\,(\ref{coupP}) between
1866: coupling constant $e$ and mass $a$ was exploited we obtained
1867: a functional dependence of the effective gauge coupling constant
1868: $e$ on temperature. Equivalently, we could have
1869: demanded $\pd_e P=0$ since $e$ is the only variable parameter of our
1870: effective theory for the electric phase. This would have {\sl directly}
1871: generated an evolution equation for temperature
1872: as a function of $e$.
1873:
1874: Radiative corrections $\Delta P$ to the
1875: pressure have a separate dependence on $a$ and $e$,
1876: %*********
1877: \eqb
1878: \label{radpressure}
1879: \Delta P=T^4 \Delta\tilde{P}(e,a,\lambda_E)\,,
1880: \eqe
1881: %**********
1882: where $\Delta\tilde{P}$ is a dimensionless function of its dimensionless arguments.
1883: To implement thermodynamical self-consistency by demanding $\pd_a P=0$ one has to
1884: express the explicitly appearing $e$ in Eq.\,(\ref{radpressure})
1885: in terms of $a$ by means of Eq.\,(\ref{coupP}) and distinguish temperature
1886: dependences arising from a simple rescaling and those
1887: arising from the $T$ dependent ground-state physics. For SU(2) we have
1888: %*********
1889: \eqb
1890: \label{e(a)}
1891: \frac{m^2}{|\phi|^2}\equiv e^2(a,\lambda_E)=\frac{T^2}{2}\,\times\frac{a^2}{|\phi|^2}=
1892: \frac{\lambda_E^2}{8\pi^2}\times a^2\lambda_E\,.
1893: \eqe
1894: %*******
1895: The first factor on the right-hand sides of Eq.\,(\ref{e(a)}) arises
1896: from rescaling, so only the second factor needs to be differentiated:
1897: %*********
1898: \eqb
1899: \label{e(a)D}
1900: \pd_a e(a,\lambda_E)=\frac{\lambda_E^2}{8\pi^2}\times\left(2a\lambda_E+a^2\pd_a\lambda_E\right)\,.
1901: \eqe
1902: %*******
1903: After solving $\pd_a P=0$ for the term $\pd_a\lambda_E$
1904: we obtain a modified right-hand side of the
1905: evolution equation Eq.\,(\ref{eeq}). The inverted solution to this
1906: evolution equation describes the dependence of mass on temperature or,
1907: after applying Eq.\,(\ref{massP}), the dependence of $e$ on temperature
1908: when two-loop diagrams for the pressure are taken into account.
1909: %***********************
1910: \begin{figure}
1911: \begin{center}
1912: \leavevmode
1913: %\epsfxsize=9.cm
1914: \leavevmode
1915: %\epsffile[80 25 534 344]{}
1916: \vspace{3.5cm}
1917: \special{psfile=Fig-7.ps angle=0 voffset=-90
1918: hoffset=-100 hscale=60 vscale=60}
1919: \end{center}
1920: \caption{Two-loop diagrams contributing to the pressure. Thick lines denote propagators of TLH modes,
1921: thin lines those of TLM modes.\label{looppress}}
1922: \end{figure}
1923: %************************
1924: The Euclidean momenta $p_e$ of off-shell fluctuations of the TLH-modes are constrained by the condition
1925: %*********
1926: \eqb
1927: \label{osmomTLH}
1928: p^2_e\le|\phi|^2(1-c_k^2\,e^2\,\frac{|\tilde{\phi}_1|}{|\phi|^2})\,,\ \ \ \
1929: (k=1,\cdots,\mbox{N}(\mbox{N}-1))\,,
1930: \eqe
1931: %***********
1932: and by the requirement that the total momentum squared
1933: flowing into or our of a four-vertex cannot be larger than $|\phi|^2$, s
1934: ee Eq.\,(\ref{Mineuosn}) and Eq.\,(\ref{dimlessdef}). Since at one loop $e$ is smaller than unity
1935: for $T\sim T_P$ only, we expect TLH-mode quantum
1936: fluctuations to be absent at temperatures lower than $T_P$ also at
1937: higher-loop accuracy.
1938:
1939:
1940: \section{The magnetic phase\label{MP}}
1941:
1942: \subsection{The electric-magnetic phase transition\label{magPT}}
1943:
1944: In Sec.\,\ref{CMP} we have discussed how stable and unstable BPS monopoles
1945: are generated as isolated objects in the electric phase. For definiteness we have assumed MGSB
1946: by the adjoint scalar $\phi$. The mass of the N/2 stable BPS monopole species is given as in
1947: Eq.\,(\ref{monopmass}) when replacing $\phi\to\tilde{\phi}_l\,,\ (l=1,\cdots,\mbox{N}/2)$.
1948: As a consequence of the evolution of the gauge
1949: coupling $e(\lambda_E)$ following from Eq.\,(\ref{eeq}) the mass of a stable
1950: monopole vanishes at $\lambda_E=\lambda_{E,c}$ due
1951: the logarithmic pole of $e$. Stable monopoles do not
1952: carry any Euclidean action at
1953: this point, and thus they condense.
1954:
1955: TLM modes, which couple to isolated
1956: monopoles with strength $g=\frac{4\pi}{e}$, become dual gauge bosons. They couple to the
1957: monopole {\sl condensates} with a strength $g$, which may now
1958: continuously vary with temperature, starting with $g=0$ at $\lambda_E=\lambda_{E,c}$.
1959: The TLH modes of the electric phase decouple kinematically at
1960: $\lambda_E=\lambda_{E,c}$ since their masses, $\propto e|\phi|$, diverge.
1961: At the onset of the {\sl magnetic} phase,
1962: where N/2 species of stable monopoles are condensed, we are thus left with an
1963: effective Abelian theory of $\mbox{N}-1$ dual gauge fields $a^D_{\mu,k}$, $(k=1,\cdots,\mbox{N}-1)$
1964: and N/2 condensates of stable monopoles described by complex scalar fields
1965: $\varphi_l\,,\ (l=1,\cdots,\mbox{N}/2)$. The temporal winding of these fields is the
1966: same as that of the associated SU(2) blocks $\tilde{\phi}_l$ in the electric phase.
1967: For N=3 there are two independent condensates of stable monopoles.
1968:
1969: What happens to the N/2$-$1 independent unstable monopoles (N even, $\mbox{N}>3$?
1970: Unstable monopoles are generated by gluon exchanges
1971: between trivial-holonomy calorons in different SU(2) embeddings. We conclude,
1972: that at $\lambda_E=\lambda_{E,c}$ the intact continuous gauge symmetry
1973: U(1)$^{\tiny\mbox{N-1}}$ of the electric phase becomes a gauge symmetry
1974: U(1)$_D^{\tiny\mbox{N-1}}$ which is spontaneously
1975: broken as
1976: %***********
1977: \eqb
1978: \label{breaking}
1979: \mbox{U}(1)_D^{\tiny\mbox{N-1}}\to \mbox{U}(1)_D^{\tiny\mbox{N/2-1}}
1980: \ (\mbox{N}>3)\,
1981: \eqe
1982: %************
1983: for $\lambda_E<\lambda_{E,c}$ in the magnetic phase. For $N=2,3$ the spontaneous
1984: breakdown of continuous gauge symmetry is maximal. The monopole condensate
1985: $\bar{\varphi}_k$, which is associated with the
1986: dual gauge-field fluctuation $\delta a^D_{\mu,k}$, is
1987: defined as
1988: %*********
1989: \eqb
1990: \label{monopcondsit}
1991: \bar{\varphi}_k=\left\{\begin{array}{c} \hspace{-1.1cm}\varphi_i\,,\ \ \ (k=1,\cdots,\mbox{N}/2)\,,\nonumber\\
1992: 0\,,\ \ \ (k=\mbox{N}/2+1,\cdots,\mbox{N}-1)\,,\end{array}\right.\,.
1993: \eqe
1994: %**********
1995: The local $Z_{\tiny\mbox{N,elec}}$ symmetry acts
1996: on $\delta a_{\mu,k}^D, \bar{\varphi}_k$ as a
1997: local-in-time permutation (see the discussion in Sec.\,\ref{CMP})
1998: %********
1999: \eqb
2000: \label{localcenter}
2001: (\delta a^D_{\mu,k},\bar{\varphi}_k)\to
2002: (\delta a^D_{\mu,(k+j(\tau))\tiny{\mbox{mod\, (N-1)}}},\bar{\varphi}_{(k+j(\tau))
2003: \tiny{\mbox{mod\, (N-1)}}})\,,\ \ \ \ (j\in \bf{Z})\,.
2004: \eqe
2005: %********
2006: In Eq.\,(\ref{localcenter}) the integer-valued functions $j$ are piecewise constant
2007: on extended regions of Euclidean spacetime. The symmetry defined in Eq.\,(\ref{localcenter})
2008: leaves the ground state of the system invariant, and thus
2009: the discrete, local symmetry $Z_{\tiny\mbox{N},elec}$ is unbroken in
2010: the magnetic phase. As a consequence the {\sl global}
2011: $Z_{\tiny\mbox{N},elec}$ associated with the Polyakov loop as an
2012: order parameter is also unbroken, see also Sec.\,\ref{polyaloop}.
2013:
2014:
2015: \subsection{Monopole condensates, macroscopically\label{windingmag}}
2016:
2017: The effective theory describing the magnetic phase is
2018: constructed in close analogy to the effective theory
2019: describing the electric phase. Recall, that we assume
2020: MGSB in the electric phase. Since the condensation of monopoles is driven by their
2021: masslessness the complex scalar fields $\varphi_l$,
2022: which describe the monopole condensates,
2023: are energy- and pressure-free in the absence of
2024: monopole interactions mediated by dual gauge-field fluctuations in the topologically trivial sector
2025: of the theory. For the N=2 case the exponent of the phase of the local field $\varphi$ is defined as
2026: %*********
2027: \eqb
2028: \label{defvarphiloc}
2029: i\log\left[\frac{\varphi}{|\varphi|}\right]=\la \int d\Sigma_{\mu\nu} \tilde{G}_{\mu\nu}
2030: \ra_{\tiny\mbox{z. m. of n.i. magn.
2031: monop.}}\,.
2032: \eqe
2033: %*********
2034: In Eq.\,(\ref{defvarphiloc}) the dual field strength $\tilde{G}_{\mu\nu}$ is the 't Hooft tensor of
2035: Eq.\,(\ref{tT}), the (surface-) integal is over a spatial
2036: 2-sphere of infinite radius, and the average is over the zero-mode deformations of a
2037: noninteracting magnetic monopole. Again, Eq.\,(\ref{defvarphiloc}) defines a dimensionless
2038: entity in accord with the fact that the Yang-Mills scale is a parameter to be
2039: measured and not to be calculated. The right-hand side of Eq.\,(\ref{defvarphiloc}) measures the magnetic
2040: flux. If monopoles are point-like, that is, if
2041: they are massive, then the right-hand side of
2042: Eq.\,(\ref{defvarphiloc}) vanishes identically due to cancellation of ingoing and outgoing
2043: fluxes. If monopoles are massless (condensed), that is, if their charges are
2044: spread over the entire Universe,
2045: then this cancellation does not take place, see \cite{HerbstHofmann2004} for a more
2046: detailed investigation. It is clear that definition (\ref{defvarphiloc}) relies on
2047: definitions (\ref{locdefphi}) and (\ref{tT}). So when expressed in terms of
2048: fundamental caloron and topologically trivial fields it looks quite involved.
2049: No (nonlocally defined) lattice operator of this type has ever been constructed.
2050: One more point needs to be discussed: The phase in Eq.\,(\ref{defvarphiloc}) should be a function of Euclidean time $\tau$,
2051: see Eq.\,(\ref{BPSsolM}). From the definition of the 't Hooft tensor, Eq.\,(\ref{tT}), we
2052: see that such a time dependence manifests itself in terms a $\tau$ dependent
2053: angle in adjoint color space between
2054: the fields $\phi^a$ and $G_{\mu\nu}^a$. Deep in the electric phase, where monopoles are
2055: isolated defects, this angle is subject to a
2056: global gauge choice for the direction of winding of the field $\phi^a$. In the magnetic phase, where
2057: monopoles are condensed, the global gauge choice in the electric phase
2058: is promoted to a {\sl local} gauge
2059: choice for the composite field $\varphi$. This situation is reminiscent of
2060: Kaluza-Klein like gemoetarical compactifications where
2061: global space-time symmetries along `extra' dimensions become gauge symmetries upon
2062: compactification \cite{Kaluza1921,Klein1926} within a low-energy formulation of the theory.
2063: This also seems to happens in the low-energy formulation of the Yang-Mills
2064: theory being in its magnetic phase.
2065:
2066: The fields $\varphi_l$ are
2067: energy- and pressure-free if and only if their Euclidean time dependence is BPS saturated.
2068: Moreover, the $\varphi_l$ must be periodic in time, and their gauge invariant
2069: modulus must not depend on spacetime. Since BPS monopoles have resonant
2070: excitations \cite{ForgasVolkov2003}, which are activated by the exchanges of
2071: dual gauge bosons, one expects the ground-state energy
2072: of the system and the tree-level mass spectrum of dual
2073: gauge-field fluctuations to be $T$ dependent.
2074: The local permutation symmetry discussed in Secs.\,\ref{CMP} and
2075: \ref{magPT} is respected by the effective potential
2076: $\tilde{V}_{M}(\varphi_1,\cdots,\varphi_{\tiny\mbox{N}/2})$ if it
2077: decomposes into a {\sl sum over potentials} $V_{M}(\varphi_l)$:
2078: %********
2079: \eqb
2080: \label{potmagn}
2081: \tilde{V}_M(\varphi_1,\cdots,\varphi_{\tiny\mbox{N}/2})\equiv\sum_{l=1}^{\tiny\mbox{N}/2}
2082: V_{M}(\varphi_l)\,.
2083: \eqe
2084: %*********
2085: The potential $V_{M}$ is uniquely determined by the above conditions. We have
2086: %********
2087: \eqb
2088: \label{singlepotM}
2089: V_{M}(\varphi_l)\equiv\overline{v_{M}(\varphi_l)}v_{M}(\varphi_l)
2090: \, \ \ \ \ \mbox{and} \ \ \ \ v_{M}(\varphi_l)=i\La_M^3/\varphi_l\,.
2091: \eqe
2092: %********
2093: In Eq.\,(\ref{singlepotM}) $\La_M$ denotes a mass scale which is
2094: related to the mass scale $\La_E$ in the electric phase by a matching
2095: condition, see Sec.\,\ref{MCC}. The effective action for the magnetic phase
2096: reads
2097: %**********
2098: \eqb
2099: \label{effactM}
2100: S_M=\int_0^{1/T}
2101: d\tau\int d^3x\,\left[\frac{1}{4}\sum_{k=1}^{\tiny\mbox{N}-1}\,
2102: \tilde{G}_{\mu\nu,k}\tilde{G}_{\mu\nu,k}+
2103: \frac{1}{2}\left(\sum_{l=1}^{\tiny\mbox{N}/2}\overline{\tilde{{\cal D}}_{\mu,l}\varphi_l}
2104: \tilde{{\cal D}}_{\mu,l}\varphi_l+\tilde{V}_M(\varphi_1,\cdots,\varphi_{\tiny\mbox{N}/2})\right)\right]\,.
2105: \eqe
2106: %**********
2107: In Eq.\,(\ref{effactM}) $\tilde{G}_{\mu\nu,l}$
2108: denotes the Abelian field strength of the dual field $a^D_{\mu,l}$, $\tilde{G}_{\mu\nu,l}\equiv
2109: \pd_\mu a^D_{\nu,l}-\pd_\nu a^D_{\mu,l}$, the covariant derivative is defined as
2110: $\tilde{{\cal D}}_{\mu,l}\equiv \pd_\mu+ig\,a^D_{\mu,l}$, and $g$ denotes
2111: the magnetic gauge coupling constant. One remark concerning the normalization of the kinetic
2112: terms for the fields $\varphi_l$ is in order. The ratio between the gauge kinetic term $\frac{1}{4}\,
2113: \tilde{G}_{\mu\nu,l}\tilde{G}_{\mu\nu,l}$ and the kinetic term
2114: $\frac{1}{2}\,\tilde{{\cal D}}_{\mu,l}\varphi_l\tilde{{\cal D}}_{\mu,l}\varphi_l$
2115: defines the mass spectrum of the gauge-field fluctuations $\delta a^D_{\mu,l}$ in unitary gauge.
2116: A redefinition of the factor in front of
2117: $\tilde{{\cal D}}_{\mu,l}\varphi_l\tilde{{\cal D}}_{\mu,l}\varphi_l$ (and $\tilde{V}_M$)
2118: changes this ratio. At the same time, however, the scale $\La_M$ is changed because the
2119: matching condition - equality of the pressure in the magnetic and electric phase at the phase boundary -
2120: is unchanged.
2121: The canonical normalization used in Eq.\,(\ref{effactM})
2122: is thus nothing but a convention for defining the scale $\La_M$ in terms of $\La_E$.
2123:
2124: \noindent The solutions to the BPS equations
2125: %********
2126: \eqb
2127: \label{BPSM}
2128: \pd_\tau \varphi_l=\bar{v}_{M}(\bar{\varphi}_l)
2129: \eqe
2130: %********
2131: read
2132: %********
2133: \eqb
2134: \label{BPSsolM}
2135: \varphi_l=\sqrt{\frac{\La_M^3}{2\pi T K(l)}}\exp[-2\pi i T K(l)\tau]
2136: \eqe
2137: %********
2138: where $K(l)$ is an integer. The condition of MGSB by
2139: minimal ground-state energy in the electric phase translated into
2140: %********
2141: \eqb
2142: \label{maxsymbr}
2143: K(l)=l\,.
2144: \eqe
2145: %********
2146: Since the $l^{\tiny\mbox{th}}$ stable monopole species in the electric phase is
2147: associated with zeros of the $l^{\tiny\mbox{th}}$ SU(2) block in $\phi$ the temporal
2148: winding of its {\sl condensate} $\varphi_l$ is accordingly.
2149: From Eqs.\,(\ref{BPSsolM}) and (\ref{maxsymbr}) we derive a potential
2150: %*********
2151: \eqb
2152: \label{potMsol}
2153: 1/2\,\tilde{V}_M=\frac{\pi\mbox{N(N+2)}}{8}\,T\Lambda_M^3
2154: \eqe
2155: %*********
2156: for even N. For N=3 we obtain
2157: %*********
2158: \eqb
2159: \label{potMsolN3}
2160: 1/2\,\tilde{V}_M=\pi\,T\Lambda_M^3\,.
2161: \eqe
2162: %*********
2163: Again, statistical fluctuations of the fields $\varphi_l$ are negligible and quantum fluctuations
2164: are absent:
2165: %*********
2166: \eqb
2167: \label{qsflphiM}
2168: \frac{\pd^2_{|\varphi_l|} V_{M}(\varphi_l)}{T^2}=24\pi^2l^2\,,\ \ \ \ \
2169: \frac{\pd^2_{|\varphi_l|} V_{M}(\varphi_l)}{|\varphi_l|^2}=6\,l^3\lambda_M^3\,.
2170: \eqe
2171: %*********
2172: In Eq.\,(\ref{qsflphiM}) we have defined $\lambda_M\equiv \frac{2\pi T}{\La_M}$. For $\mbox{N}=2,3$
2173: $\lambda_M$ is larger than unity throughout the magnetic phase, see Sec.\,\ref{TSCM}.
2174: Since the fields $\varphi_l$ do not fluctuate they are a background
2175: to the macroscopic gauge-field equations of
2176: motion
2177: %*********
2178: \eqb
2179: \label{eomdualG}
2180: \pd_\mu \tilde{G}_{\mu\nu,l}=ig\left[\overline{\tilde{{\cal D}}_{\nu,l}\varphi_l}\varphi_l-\bar{\varphi_l}
2181: \tilde{{\cal D}}_{\nu,l}\varphi_l\right]\,.
2182: \eqe
2183: %*********
2184: There exist pure-gauge solutions to Eq.\,(\ref{eomdualG}) given as
2185: %*******
2186: \eqb
2187: \label{pgsolM}
2188: a^{D,bg}_{\mu,l}=\delta_{\mu 4}\frac{2\pi l}{g}\,T\,.
2189: \eqe
2190: %********
2191: Again, a macroscopic `holonomy' is created by interacting monopoles which can be
2192: related to the existence of isolated loops of magnetic flux: ANO vortices.
2193: We have $\tilde{{\cal D}}_{\nu,l}\varphi_l=0$.
2194: As a consequence, the action in Eq.\,(\ref{effactM}) reduces to the potential term
2195: on the ground-state solutions $\varphi_l, a^{D,bg}_{\mu,l}$. This results in a shift of
2196: the ground-state energy density and pressure to $\pm 1/2\,\tilde{V}_M$, respectively.
2197:
2198: \noindent Nonperiodic gauge functions
2199: %*******
2200: \eqb
2201: \label{grM}
2202: \theta_l=2\pi l T\tau
2203: \eqe
2204: %*******
2205: transform each pair $\varphi_l,\, a^{D,bg}_{\mu,l}$ to unitary gauge
2206: %********
2207: \eqb
2208: \label{unitgaugeM}
2209: \varphi_l=|\varphi_l|\,,\ \ \ \ \ a^{D,bg}_{\mu,l}=0\,.
2210: \eqe
2211: %********
2212: In analogy to the electric phase one shows that
2213: the gauge rotations $\Omega_l=e^{\theta_l}$ leave intact the
2214: periodicity of the fluctuations $\delta a^D_{\mu,l}$. These gauge rotations thus
2215: do not change the physics upon integrating out the fields $\delta a^D_{\mu,l}$
2216: at one loop. Due to the effective theory being Abelian and the monopole condensates $\varphi_l$
2217: inert the one-loop calculation is exact. There is, however, a pronounced difference
2218: to the electric phase. In winding as well as in unitary gauge the
2219: Polyakov loop evaluated on each of the background fields $a^{D,bg}_{\mu,l}$
2220: is {\sl unity}. We conclude that the ground state of the system is unique,
2221: much in contast to the electric phase, and thus that the global
2222: $Z_{\tiny\mbox{N,elec}}$ symmetry is {\sl restored}. For a discussion of the
2223: full Polyakov loop see Sec.\,\ref{polyaloop}. Hence the magnetic phase confines
2224: fundamental test charges at the same time as it allows for the propagation of
2225: massive, Abelian gauge modes!
2226:
2227:
2228: \subsection{Gauge-field excitations and \\ thermodynamical self-consistency\label{TSCM}}
2229:
2230:
2231: The Abelian Higgs mechanism generates a tree-level
2232: mass spectrum for the fluctuations $\delta a^D_{\mu,l}$. It is given as
2233: %********
2234: \eqb
2235: \label{massspecM}
2236: m_l=g|\varphi_l|\equiv a_l\,T\,.
2237: \eqe
2238: %*********
2239: Due to the noncondensation of unstable monopoles there are N/2$-1$
2240: dual gauge field fluctuations
2241: %********
2242: \eqb
2243: \label{masslessM}
2244: \delta a^D_{\mu,i}\,,\ \ (i=\mbox{N}/2+1,\cdots,\mbox{N})\,
2245: \eqe
2246: %********
2247: which are massless. In analogy to the electric phase we derive
2248: an evolution equation for temperature as a function of mass from the
2249: requirement of thermodynamical self-consistency, $\pd_a P=0$
2250: (for the definition of $a$ in the magnetic phase
2251: see Eq.\,(\ref{defM})). Notice that the
2252: one-loop expression for thermodynamical quantities and the tree-level masses of
2253: the dual gauge boson fluctuations $\delta a^D_{\mu,k}\,,\ (k=1,\mbox{N}-1)$ are exact due to
2254: the effective theory being Abelian. We obtain
2255: %********
2256: \eqb
2257: \label{evolMeven}
2258: \pd_a\lambda_M=-\frac{96}{(2\pi)^6 \mbox{N(N+2)}}\lambda_M^4\, a
2259: \sum_{l=1}^{\tiny{\mbox{N/2}}}c_l^2 D(a_l)\,,
2260: \ \ \ (\mbox{N\ \ even})\,.
2261: \eqe
2262: %**********
2263: For N=3 we have
2264: %********
2265: \eqb
2266: \label{evolM3}
2267: \pd_a\lambda_M=-\frac{12}{(2\pi)^6}\lambda_M^4\, a D(a)\,.
2268: \eqe
2269: %**********
2270: In Eqs.\,(\ref{evolMeven}) and (\ref{evolM3}) we have defined:
2271: %********
2272: \eab
2273: \label{defM}
2274: c_l&\equiv&\frac{1}{\sqrt{l}}\,,\ \ \ \
2275: \lambda_M\equiv\frac{2\pi T}{\La_M}\,,\nonumber\\
2276: a&\equiv& \frac{g}{T}|\varphi_1|=2\pi\,g\,\lambda_M^{-3/2}\,,\ \ \ \ a_l\equiv c_l a\,.
2277: \eae
2278: %*********
2279: In deriving Eqs.\,(\ref{evolMeven}) and (\ref{evolM3})
2280: we have neglected the `nonthermal' contribution $-\Delta \tilde{V}_M$ to the pressure which is
2281: very small for sufficiently small N (quantum fluctuations are cut
2282: off at the compositeness scales $|\varphi_l|$):
2283: %**********
2284: \eqb
2285: \label{DeltaVM}
2286: \left|\frac{\Delta \tilde{V}_M}{\tilde{V}_M}\right|<
2287: 3\,\frac{\mbox{N}-1}{128\pi^2}\left(\frac{|\varphi_1|}{\La_M}\right)^6\sim
2288: 3\,\frac{\mbox{N}-1}{128\pi^2}\,\lambda_M^{-3}\,.
2289: \eqe
2290: %**********
2291: The function $D(a)$ is defined in Eq.\,(\ref{DA}). The evolution equations
2292: (\ref{evolMeven}) and (\ref{evolM3}) have fixed points at $a=0,\infty$ which correspond to
2293: the highest and lowest attainable temperatures in the magnetic phase,
2294: $\lambda_{E,c}$ and $\lambda_{M,c}$, respectively.
2295:
2296: The $\lambda_M$ dependence of the gauge coupling constants $g$
2297: is obtained by inverting the solutions to Eqs. (\ref{evolMeven}) and
2298: (\ref{evolM3}) and using the relation between $g$, $\lambda_M$ and $a$ in
2299: Eq.\,(\ref{defM}) afterwards. Results for N=2,3 are shown in Fig.\,\ref{gevol}.
2300: %***********************
2301: \begin{figure}
2302: \begin{center}
2303: \leavevmode
2304: %\epsfxsize=9.cm
2305: \leavevmode
2306: %\epsffile[80 25 534 344]{}
2307: \vspace{4.5cm}
2308: \special{psfile=Fig-8.ps angle=0 voffset=-140
2309: hoffset=-140 hscale=60 vscale=60}
2310: \end{center}
2311: \caption{The evolution of the gauge
2312: coupling constant $g$ in the magnetic phase for
2313: N=2 (thick grey line) and N=3 (thick black line). The gauge coupling constant
2314: diverges logarithmically, $g\propto -\log(\lambda_{M}-\lambda_{M,c})$, at
2315: $\lambda_{M,c}=6.41$ (N=2) and $\lambda_{M,c}=5.82$ (N=3). \label{gevol}}
2316: \end{figure}
2317: %************************
2318: The gauge coupling constant $g$ increases continuously from $g=0$ at the electric-magnetic
2319: phase boundary ($\lambda_M=\lambda_{E,c}$) until it diverges logarithmically at $\lambda_{M,c}$.
2320: A continuous behavior of the magnetic
2321: coupling is not in contradiction with magnetic charge conservation
2322: since no isolated magnetic charges appear
2323: in the magnetic phase: magnetic monopoles are only present in condensed
2324: form, and there are no collective monopole excitations, see Eq.\,(\ref{qsflphiM}). The
2325: continuous rise of $g$, starting from zero at $\lambda_{M,c}$, implies a continuous behavior of the
2326: mass parameter $a$ in Eq.\,(\ref{defM}). Since $a$ is the measurable order parameter
2327: for monopole condensation this situation is reminiscent of a
2328: 2$^{\tiny\mbox{nd}}$ order phase transition. We compute the critical exponent of this
2329: transition for N=2 in Sec.\,\ref{critexpo}.
2330:
2331: Lowering $\lambda_M$ towards $\lambda_{M,c}$ the core size of ANO vortices
2332: becomes large, the monopole condensates are more and
2333: more distorted by magnetic flux lines, and thus it becomes increasingly
2334: irrelevant in what SU(2) embedding a particular monopole lives. Formerly
2335: unstable monopoles acquire longevity and thus additional monopole
2336: condensates form at $\lambda_{M,c}$: $\bar{\varphi}_i\not=0\,,\ (i=\mbox{N}/2+1,\cdots,\mbox{N}-1)$ for
2337: $\lambda_M\sim \lambda_{c,M}$.
2338: As a consequence, {\sl all} dual gauge-boson fluctuations $\delta a^D_{\mu,k}\,,\ (k=1,\cdots,\mbox{N}-1)$
2339: are very massive close to $\lambda_{M,c}$ and decouple thermodynamically at $\lambda_{M,c}$.
2340: The equation of state at $\lambda_M=\lambda_{c,M}$ thus is
2341: %********
2342: \eqb
2343: \label{eosMc}
2344: \rho(\lambda_{M,c})=-P(\lambda_{M,c})\,.
2345: \eqe
2346: %********
2347: At $\lambda_{M,c}$, where all dual gauge-field fluctuations are very massive, the continuous
2348: dual gauge symmetry $U(1)_D^{\tiny\mbox{N-1}}$ is broken
2349: completely.
2350:
2351: \subsection{Polyakov loop in the electric and the magnetic phase\label{polyaloop}}
2352:
2353: In this section we show that the Polyakov loop, which is an order parameter
2354: for the deconfining transition, indeed is finite in the electric
2355: and close to zero in the magnetic phase. In each phase the Polyakov loop of the full
2356: effective theory (now we also consider the fluctuations $\delta a_\rho$)
2357: formulated in Euclidean spacetime and unitary gauge is defined as
2358: %***********
2359: \eab
2360: \label{Ployeff}
2361: {\bf P}&=&Z^{-1}\times\exp[-S_{cl}]\times
2362: \int {\cal D}\delta b_\rho\,\exp\left[-i\gamma\int_0^{T^{-1}} d\tau\,\delta b_4^{b.g.}\right]\times\nonumber\\
2363: & &\exp\left[-\int_0^{T^{-1}}d\tau d^3x\left\{\frac{1}{4}G^2[\delta b_\rho]+
2364: \sum_k m^2_k\,\delta b_{\mu,k}\delta b_{\mu,k}\right\}\right]\,.
2365: \eae
2366: %**********
2367: where $\gamma=\{e,g\}$ and $\delta b_\rho=\{\delta a_\rho,\delta a^D_\rho\}$
2368: depending on whether we discuss the electric or the magnetic phase. The term
2369: $\exp[-S_{cl}]$ refers to the vanishing weight of the ground state in
2370: the partition function $Z$ in either phase. This weight, however, cancels in expectation
2371: values.
2372: %***********************
2373: \begin{figure}
2374: \begin{center}
2375: \leavevmode
2376: %\epsfxsize=9.cm
2377: \leavevmode
2378: %\epsffile[80 25 534 344]{}
2379: \vspace{3.5cm}
2380: \special{psfile=Fig-8b.ps angle=0 voffset=-120
2381: hoffset=-240 hscale=60 vscale=70}
2382: \end{center}
2383: \caption{The data set used for the fit of the critical exponent $\nu$.\label{nudata}}
2384: \end{figure}
2385: %************************
2386: Since the fluctuations $\delta b_\rho$
2387: are periodic in time they are decomposed as
2388: %********
2389: \eqb
2390: \label{fourierb}
2391: \delta b_\rho(\tau,{\vec x})=\sum_{n=-\infty}^{n=\infty}
2392: \exp\left[2\pi in \frac{\tau}{T}\right]\bar{b}_{\rho,n}({\vec x})\,.
2393: \eqe
2394: %********
2395: Modes with $n\not=0$ make the Polyakov-loop
2396: phase in Eq.\,(\ref{Ployeff}) unity, are action-suppressed,
2397: and thus they are irrelevant. Zero modes ($n=0$) make a contribution
2398: to the Polyakov-loop phase if they are not action-suppressed. This is the case
2399: if both of the following conditions are met: (i) there is no mass term for these
2400: fluctuations and (ii) we have $\pd_i\bar{b}_{\rho,0}({\vec x})\sim 0$ where $\pd_i$
2401: denotes a spatial derivative. In the electric
2402: phase TLM modes are massless and thus their space-indepent zero-mode fluctuations
2403: generate the bulk of the (finite) Polyakov loop. For N=2,3 there are no massless
2404: fluctuations in the magnetic phase and thus condition (i) is violated. As a consequence, none of
2405: the fluctuations $\delta b_\rho$ can contribute
2406: to the Polyakov loop in a substantial way in the magnetic phase and thus we have
2407: ${\bf P}\sim 0$. For N$>$3 the presence of unstable magnetic monopoles in the
2408: magnetic phase prevents some of the TLM modes to pick up a mass by the Abelian
2409: Higgs mechanism. Thus the Polyakov loop should be small but nonvanishing in
2410: the magnetic phase.
2411:
2412: \subsection{Critical exponent for the SU(2) electric-magnetic transition\label{critexpo}}
2413:
2414: In this section we compute the critical exponent $\nu$ for the
2415: electric-magnetic transition for N=2. The obvious order parameter for this transition is
2416: the `photon' mass.
2417:
2418: The data set in Fig.\,\ref{nudata} is generated from an inversion of the numerical solution to the
2419: evolution equation Eq.\,(\ref{evolMeven}). The following model is used to fit the data
2420: %********
2421: \eqb
2422: \label{modeldata}
2423: a(\lambda_M)=C\times |\lambda_M-\lambda_{M,em}|^\nu
2424: \eqe
2425: %********
2426: where $C$ and $\nu$ are constants, $\lambda_{M,em}$ denotes the critical
2427: temperature $T_{e,c}$ in units of $\frac{\Lambda_M}{2\pi}$, and $a$ is the dimensionless `photon' mass.
2428: Recall, that the value $\lambda_{E,c}=11.65$ is obtained from the position of the logarithmic pole
2429: of the coupling constant $e$ in the {\sl electric phase}. By matching the pressure at
2430: the electric-magnetic phase boundary, see Sec.\ref{MCC}, this translates into a
2431: value $\lambda_{M,em}=7.337$.
2432:
2433: To perform the actual fit we
2434: have used Mathematica's NonlinearFit function which is contained in the
2435: statistics package. In Fig.\,\ref{shapeFIT} the
2436: critical behavior of the `photon' mass $a$ is shown.
2437: %***********************
2438: \begin{figure}
2439: \begin{center}
2440: \leavevmode
2441: %\epsfxsize=9.cm
2442: \leavevmode
2443: %\epsffile[80 25 534 344]{}
2444: \vspace{3.5cm}
2445: \special{psfile=Fig-8a.ps angle=0 voffset=-130
2446: hoffset=-140 hscale=60 vscale=50}
2447: \end{center}
2448: \caption{The function $a(\lambda_M)$ in the vicinity of the electric-magnetic phase transition.
2449: The region between the dashed vertical lines corresponds to the data set in Tab.\,1 which generates the
2450: least sensitivity of $\nu$ on the length of the
2451: fitting interval $\Delta=|\lambda_{M,\tiny\mbox{min,fit}}-\lambda_{M,em}|$.\label{shapeFIT}}
2452: \end{figure}
2453: %************************
2454: To determine the length of the fitting interval
2455: $\Delta=|\lambda_{M,\tiny\mbox{min,fit}}-\lambda_{M,em}|$ where $\nu$ is
2456: least sensitive to changes in $\lambda_{M,\tiny\mbox{min,fit}}$ we numerically
2457: determine the inflexion point $\Delta_{\tiny\mbox{inflex}}$ of the function $\nu=\nu(\Delta)$,
2458: see Fig.\,\ref{inflexionpoint}. We obtain
2459: %*********
2460: \eqb
2461: \label{inflexpo}
2462: \Delta_{\tiny\mbox{inflex}}=0.29\pm0.05\,.
2463: \eqe
2464: %*********
2465: We emphasize that this interval is well contained in the fitting
2466: interval used in \cite{deForcrandEliaPepe2001} where the critical exponent was determined
2467: from the expectation of the dual string tension. Their fit interval $0\le t\le 1$ with
2468: $t\equiv\frac{T_{M,em}-T}{T_{M,em}}$ correponds to an interval length
2469: $\Delta=\lambda_{M,em}\sim 7.34$!
2470:
2471: The interval of least senstivity $\Delta_{\tiny\mbox{inflex}}=0.29\pm0.05$
2472: translates into
2473: %**********
2474: \eqb
2475: \label{nuFITst}
2476: \nu=0.61+0.02-0.01\,.
2477: \eqe
2478: %**********
2479: Alternatively, we can determine $\nu$ by a fit in
2480: very small intervals around $\lambda_{M,em}$. In this case we obtain the result expected from a naive
2481: mean-field analysis, $\nu\to 0.5$. However, the fitted
2482: prefactor $C$ varies considerably for very small intervals around $\lambda_{M,em}$.
2483: It is worth mentioning at this point that the `would-be' critical
2484: exponent for N=3 has at least
2485: two inflexion points as a function of $\Delta$. This makes a unique determination
2486: of the physical value of $\nu$
2487: impossible and clearly indicates that the phase transition is not second order anymore. A small latent
2488: heat is associated with the electric-magnetic transition for N=3, see Fig.\,\ref{rho},
2489: which makes it
2490: weakly first order. This is seen on the lattice \cite{LuciniTeperWenger2003}.
2491:
2492: The critical exponent for the 3D Ising
2493: model (same universality class as SU(2) Yang-Mills \cite{SvetitskyYaffe1982-1,SvetitskyYaffe1982-2})
2494: is $\nu_{\tiny\mbox{Ising}}\sim 0.63$. As it seems, the effective theories for the electric and the
2495: magnetic phases have passed an important test!
2496:
2497:
2498: \section{The center phase\label{CVCM}}
2499:
2500: \subsection{Isolated center vortices in the magnetic phase}
2501:
2502: Away from the points $\lambda_{E,c}$ and $\lambda_{M,c}$ there
2503: are isolated vortices in the magnetic
2504: phase which form closed loops due to the
2505: conservation of magnetic flux. Along the
2506: core region of a vortex, where the monopole condensate vanishes,
2507: $\bar{\varphi}_k\approx 0$, monopoles and antimonopoles form a closed
2508: chain and move into opposite directions \cite{Olejnik1997}.
2509: Along a vortex loop the magnetic flux is $2\pi\,k/N\,,\ (k=1,\cdots,\mbox{N}-1)$
2510: with respect to the dual gauge field $a^D_{\mu,k}$. This coins the
2511: name center vortex loop.
2512: %***********************
2513: \begin{figure}
2514: \begin{center}
2515: \leavevmode
2516: %\epsfxsize=9.cm
2517: \leavevmode
2518: %\epsffile[80 25 534 344]{}
2519: \vspace{3.5cm}
2520: \special{psfile=Fig-8c.ps angle=0 voffset=-120
2521: hoffset=-140 hscale=60 vscale=50}
2522: \end{center}
2523: \caption{Dependence of the critical exponent $\nu$ on the length of the fitting
2524: interval $\Delta=|\lambda_{M,\tiny\mbox{min}}-\lambda_{M,c}|$. The inflexion point
2525: of the curve is at $\Delta_{\tiny\mbox{inflex}}=
2526: 0.29\pm0.05$ corresponding to $\nu=0.61+0.02-0.01$.\label{inflexionpoint}}
2527: \end{figure}
2528: %************************
2529: The typical action $S_{\tiny\mbox{ANO}}$ and energy $E_{\tiny\mbox{ANO}}$ of a center vortex loop can be
2530: estimated since an analytical solution to the Abelian Higgs
2531: model is known \cite{NielsenOlesen1973}:
2532: %********
2533: \eqb
2534: \label{ANOana}
2535: S_{\tiny\mbox{ANO}}\sim\frac{1}{g^2}\,,\ \ \ \ \ \ \ E_{\tiny\mbox{ANO}}\propto \frac{1}{g}\,.
2536: \eqe
2537: %********
2538: As a consequence of Eq.\,(\ref{ANOana}) center vortex
2539: loops do not carry action and are
2540: massless at $\lambda_{M,c}$ where $g$
2541: diverges. They condense to form a new ground state of
2542: the system: a phase transition takes place. As we will show below this phase
2543: transition is nonthermal and of the Hagedorn type.
2544:
2545: \subsection{Center vortex condensates, macroscopically}
2546:
2547: Since center vortex loops are extended, one-dimensional
2548: objects the local scalar fields $\Phi_k(x)\,,\ (k=1,\cdots,\mbox{N}-1)$
2549: describing their respective condensates have to be
2550: defined in a nonlocal way. We formally define the fields $\Phi_k(x)$
2551: in terms of an average over the dual Abelian gauge
2552: fields $A^D_{\mu,k}$ of the magnetic phase
2553: (the part belonging to an ANO or center vortex is included in $A^D_{\mu,k}$!) as
2554: %*********
2555: \eqb
2556: \label{Wilsonloopx}
2557: \frac{\Phi_k(x)}{|\Phi_k(x)|}=\la \exp\left[ig\oint dz_\mu A^D_{\mu,k}\right]\ra_{A^D_{\mu,k}}\,.
2558: \eqe
2559: %*********
2560: In Eq.\,(\ref{Wilsonloopx}) the integration contour is spatial and circular,
2561: its center is at $x$, and circle's diameter is infinite. In absence of a Yang-Mills scale
2562: $\Lambda_C$, which is a relevant situation
2563: when investigating the ground-state structure in the center phase due
2564: to the missing gauge-mode propagation, only dimensionless quantities
2565: like in Eq.\,(\ref{Wilsonloopx}) can be computed. As we shall see below, the presence of
2566: $\Lambda_C$ can only be observed in the excitation spectrum (selfintersections of center-vortices).
2567: We may always chose a parametrization of the integration
2568: contour $z(\xi)$ in Eq.\,(\ref{Wilsonloopx})
2569: such that $0\le\xi\le 1$.
2570:
2571: A (local) magnetic center rotation can be expressed as
2572: %*********
2573: \eqb
2574: \label{centerrot}
2575: \Omega(z)=\exp\left[\frac{2\pi i}{\mbox{N}} \sum_{k=1}^{\tiny\mbox{N-1}}\chi_k(z)\right]\,
2576: \eqe
2577: %*********
2578: where $\chi_k(z)$ is either $k$ or zero. The dual gauge field $A^D_{\mu,k}$ transforms
2579: under $\Omega(z)$ as
2580: %*********
2581: \eqb
2582: \label{trafocenterrot}
2583: A^D_{\mu,k}\to A^D_{\mu,k}-g^{-1}\frac{2\pi}{\mbox{N}}\pd_\mu \chi_k(z)\,.
2584: \eqe
2585: %*********
2586: %***********************
2587: \begin{figure}
2588: \begin{center}
2589: \leavevmode
2590: %\epsfxsize=9.cm
2591: \leavevmode
2592: %\epsffile[80 25 534 344]{}
2593: \vspace{4.5cm}
2594: \special{psfile=Fig-9.ps angle=0 voffset=-140
2595: hoffset=-140 hscale=50 vscale=50}
2596: \end{center}
2597: \caption{The phase of a local magnetic center transformation along a
2598: circular contour parametrized by $0\le \xi\le 1$ and its
2599: decomposition into boxes. \label{locCP}}
2600: \end{figure}
2601: %************************
2602: According to the definition (\ref{Wilsonloopx}) a
2603: magnetic center rotation $\Omega(z)$ may locally add a
2604: magnetic flux quantum \cite{'tHooft1978} to the flux contained
2605: in $\Phi_k$. This is possible since we may have
2606: $\Omega(z(0))\not=\Omega(z(1))$. The action of the local magnetic
2607: center rotation $\Omega(z)$ on the complex field $\Phi_k(x)$ is as follows:
2608: %********
2609: \eqb
2610: \label{centerphi}
2611: \Phi_k(x)\to\left\{\begin{array}{c}\exp[\pm\frac{2\pi i k}{\tiny\mbox{N}}]\,\Phi_k(x),\ \ \
2612: (|\chi_k(z(1))-\chi_k(z(0))|=k)\,,\\
2613: \,\,\,\,\,\Phi_k(x)\,,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \mbox{otherwise} \end{array}\right.\,.
2614: \eqe
2615: %*********
2616: Obviously, the action of
2617: $\Omega(z)$ on $\Phi_k(x)$ may locally change the phase of the field $\Phi_k$,
2618: and thus the ground state is not invariant under local magnetic center
2619: transformations: $Z_{\mbox{N,mag}}$ as a discrete gauge symmetry is spontaneously
2620: broken in the center phase.
2621:
2622: The local center transformation $\Omega$ singles out possible
2623: `boundaries of the circle' at the positions $z(\xi_n)$, ($\xi_0=0,\cdots$),
2624: where it jumps, see Fig.\,\ref{locCP}. The size and direction of an
2625: $\Omega$-induced magnetic flux quantum - an observable quantity -
2626: should not depend on the admissible re-parametrizations
2627: $\xi(\zeta)$, $(0\le\zeta\le 1)$ with $\xi(0)\in\{\xi_n\}$.
2628: For example, a translation
2629: %********
2630: \eqb
2631: \label{zeta}
2632: \xi(\zeta)=\zeta+\xi_1
2633: \eqe
2634: %********
2635: would have shifted the `boundary of the circle' at $z(\xi=0)$ to $z(\xi=\xi_1)$. Thus
2636: it would have generated a different flux quantum. To avoid such an ambiguity
2637: we have to impose that a
2638: reparametrization of the circle is compensated for by an according local
2639: permutation of the fields $\Phi_k(x)$: if flux quanta $\frac{2\pi k}{\mbox{N}}$ and
2640: $\frac{2\pi l}{\mbox{N}}$, ($l\not=k$),
2641: were generated by $\Omega$ in the parameterizations
2642: $\xi$ and $\zeta$, respectively, then a
2643: permutation with $\Phi_l\to\Phi_k$ needs to be
2644: performed {\sl after} the reparametrization $\zeta$ was implemented.
2645: This yields the same flux quantum as in parametrization
2646: $\xi$. Now, the discontinuous phase
2647: change in Eq.\,(\ref{centerphi}) is
2648: provided by the {\sl dynamics} in an effective theory
2649: and is not externally imposed. As a consequence,
2650: the demand for invariance of a generated flux quantum
2651: under admissible contour-reparametrizations translates
2652: into an invariance of the effective Lagrangian under local permutations
2653: of the set $\{\Phi_1(x),\cdots,\Phi_{\tiny\mbox{N-1}}(x)\}$. The following Lagrangian
2654: satisfies this requirement:
2655: %************
2656: \eqb
2657: \label{actcenter}
2658: {\cal L}_C=\frac{1}{2}\sum_{k=1}^{\tiny\mbox{N-1}}
2659: \left[\pd_\mu\Phi_k \pd_\mu\Phi_k+V_C(\Phi_k)\right]\,.
2660: \eqe
2661: %************
2662: It is clear from Eq.\,(\ref{actcenter}) that the transformation
2663: in Eq.\,(\ref{centerphi}) is only a symmetry of the potential term
2664: due to the noninvariance of the kinetic terms under the jumps in the fields $\Phi_k$.
2665: However, because of the conservation of magnetic flux (only closed loops of center vortex
2666: flux can be generated) one quantum of center flux created by a jump at one point in spacetime
2667: is compensated by an opposite quantum of
2668: center flux created by the opposite jump at another point.
2669: The spacetime integral over ${\cal L}_C$, the action,
2670: is therefore invariant under local center rotations which induce magnetic
2671: flux in a closed flux line. Based on the continuum Lagrangian
2672: in Eq.\,(\ref{actcenter}) it would be interesting to perform a
2673: lattice simulation of the transition to the center phase taking any member of the set $\{|\Phi_k|\}$ as an
2674: order parameter.
2675: %***********************
2676: \begin{figure}
2677: \begin{center}
2678: \leavevmode
2679: %\epsfxsize=9.cm
2680: \leavevmode
2681: %\epsffile[80 25 534 344]{}
2682: \vspace{6.5cm}
2683: \special{psfile=Fig-9-10-10.ps angle=0 voffset=-200
2684: hoffset=-140 hscale=50 vscale=50}
2685: \end{center}
2686: \caption{The potential $V_C=\overline{v_C(\Phi)}v_C(\Phi)$ corresponding
2687: to the definition in Eq.\,(\ref{vCright})
2688: for N=2,3,4 and $\La_C=\Lambda^\prime_C$.
2689: $|\Phi|$ is given in units of $\La_C$ and $V_C$ in units of $\La_C^4$. Notice the minima $V_C=0$ at
2690: the N$^{\tiny\mbox{th}}$ unit roots. \label{pot234}}
2691: \end{figure}
2692: %************************
2693: The function $V_C(\Phi)$ in Eq.\,(\ref{actcenter}) can be written as
2694: $V_C(\Phi)=\overline{v_C(\Phi)}v_C(\Phi)$. Let us now discuss the properties of
2695: the function $v_C$. If a matching to the magnetic phase would
2696: take place in thermodynamical equilibrium and a classical treatment of the ground-state dynamics could be
2697: justified at this point then the Euclidean time
2698: dependence of the (periodic) fields $\Phi_k$ would have to be BPS
2699: saturated. Only in this case do the fields $\Phi_k$ describe the condensates of
2700: zero-energy center-vortex loops \footnote{Recall, that all gauge-boson fluctuations are
2701: decoupled in the center
2702: phase. As a consequence, interaction between center vortices are extremely local.}.
2703:
2704: At first sight a candidate for $v_C$ would
2705: be $v_{C,\tiny\mbox{trial}}(\Phi_k)=i\Lambda_C^3/\Phi_k$. The potential $V_C$ would then not only
2706: be $Z_{\tiny\mbox{N,mag}}$ but also $U(1)^{\tiny\mbox{N-1}}$ symmetric. The
2707: latter (global) symmetry, however, does not exist in SU(N) Yang-Mills theory
2708: and thus $v_{C,\tiny\mbox{trial}}$ is excluded.
2709:
2710: A function $v_C(\Phi_k)$, which is covariant {\sl only}
2711: under $Z_{\tiny\mbox{N}}$ transformations and, at the same time, allows for periodic and
2712: BPS saturated solutions along the Euclidean time coordinate $\tau$ \cite{Hofmann2002},
2713: see Fig.\,\ref{H2000},
2714: %***********************
2715: \begin{figure}
2716: \begin{center}
2717: \leavevmode
2718: %\epsfxsize=9.cm
2719: \leavevmode
2720: %\epsffile[80 25 534 344]{}
2721: \vspace{4.5cm}
2722: \special{psfile=Fig-9n.ps angle=0 voffset=-140
2723: hoffset=-190 hscale=60 vscale=40}
2724: \end{center}
2725: \caption{Numerical solutions to the BPS equation $\pd_\tau\Phi =\overline{v_C(\Phi)}$
2726: for N=2,3,4. \label{H2000}}
2727: \end{figure}
2728: %************************
2729: \footnote{The latter requirement derives from a consideration of
2730: the limit $\mbox{N}\to\infty$ where these solutions are of
2731: physical relevance, see below.}, is uniquely given as
2732: %***********
2733: \eqb
2734: \label{vCright}
2735: v_C(\Phi_k)=i\left(\frac{\Lambda_C^3}{\Phi_k}-\frac{\Phi_k^{\tiny\mbox{N-1}}}
2736: {(\Lambda^\prime_C)^{\tiny\mbox{N-3}}}\right)\,
2737: \eqe
2738: %***********\
2739: where $\Lambda_C$ and $\Lambda^\prime_C$ denote mass scales that are a priori independent.
2740: The N degenerate minima of the potential $V_C(\Phi_k)$ are at
2741: %*********
2742: \eqb
2743: \label{minPot}
2744: |\Phi^{\tiny\mbox{min}}_k|=\left[\La_C^3 (\Lambda^\prime_C)^{\tiny\mbox{N-3}}\right]^{1/\tiny\mbox{N}}\,.
2745: \eqe
2746: %********
2747: At these minima the potential $V_C$ {\sl vanishes}. Periodic solutions to the BPS
2748: equations
2749: %*********
2750: \eqb
2751: \label{PBScenter}
2752: \pd_\tau\Phi_k=\bar{v_C(\Phi_k)}
2753: \eqe
2754: %*********
2755: are parameterized by a winding number $\in\bf{Z}$ in analogy to the situation in the magnetic phase.
2756: The fields $\Phi_l\,,\ \ (l=1,\cdots,\mbox{N}/2)$, which
2757: are associated with vortices formed from the stable dipoles of the electric phase, are
2758: winding accordingly. For the fields $\Phi_i\,,\ \ (i=\mbox{N}/2+1,\cdots,\mbox{N}-1)$, being associated
2759: with vortices formed from independent monopoles, which are unstable in the electric phase, no winding numbers can be
2760: derived. This uncertainty in assigning winding numbers for the fields $\Phi_i$
2761: is reflected in the appearance of an additional mass scale $\Lambda^\prime_C$ in
2762: Eq.\,(\ref{vCright}). In the cases N=2,3, however, only {\sl stable}, independent monopoles exist,
2763: and thus we can set $\La_C=\Lambda^\prime_C$.
2764:
2765: In Fig.\,\ref{pot234} the
2766: graphs of the potential $V_C$ for N=2,3,4 and for $\La_C=\Lambda^\prime_C$ are shown.
2767: Notice the ridges and the valleys of
2768: negative and positive tangential curvature, respectively.
2769: Due to the BPS saturation a solution
2770: $\Phi_k$ to Eq.\,(\ref{PBScenter}) is guaranteed to carry no energy. This statement is, however, only then useful
2771: for the description of the ground state if the classical approximation can be
2772: justified. At finite N the modulus of BPS saturated solutions
2773: is no longer $\tau$ independent, see Fig.\,\ref{H2000}.
2774: On the one hand, this situation is in contradiction
2775: to thermal equilibrium. On the other hand, we can not trust the classical approximation
2776: since tangential fluctuations $\theta_k$, defined as
2777: %***********
2778: \eqb
2779: \label{TanModes}
2780: \Phi_k=|\Phi_k|\exp[i\frac{\theta_k}{\Lambda_C}]\,,
2781: \eqe
2782: %************
2783: can be tachyonic and therefore destabilize the classical solution to Eq.\,(\ref{PBScenter}), see
2784: Fig.\,\ref{tac}. In the electric and the magnetic phase tangential fluctuations of the caloron and the monopole
2785: condensates are would-be Goldstone modes giving rise to longitudinal polarizations
2786: of gauge boson fluctuations. Due to the absence of a continuous gauge
2787: symmetry in the center phase no gauge-field fluctuations exist
2788: which could `eat' the tangential fluctuations. How to integrate out tachyonic tangential
2789: modes analytically is conceptually unclear. The only definite statement we
2790: can make at present is that they rapidly drive the expectation of the fields
2791: $\Phi_k$ towards the minima of $V_C$: $\Phi_k$ relaxes to one of the minima
2792: of $V_C$ along an outward directed spiral in the complex plane. During this process magnetic flux quanta
2793: are locally generated by discontinuous phase changes of $\Phi_k$ due the tunneling through the
2794: regions where the tangential fluctuations are tachyonic.
2795: %***********************
2796: \begin{figure}
2797: \begin{center}
2798: \leavevmode
2799: %\epsfxsize=9.cm
2800: \leavevmode
2801: %\epsffile[80 25 534 344]{}
2802: \vspace{4.5cm}
2803: \special{psfile=Fig-9-10-1.ps angle=0 voffset=-140
2804: hoffset=-120 hscale=50 vscale=50}
2805: \end{center}
2806: \caption{The ratio $\left.\pd^2_{\theta_k} V_C(\Phi_k)/|\Phi_k|^2\right|_{|\Phi_k(0)|=0.8\La_C}$ ($V_C$ defined
2807: in Eq.\,(\ref{vCright}) and $\Lambda_C=\La_C^\prime$) as a function of $\frac{\theta_k}{\La_C}$
2808: for N=2 (thick grey line), N=3 (thick black line), N=4 (dashed line), and N=10
2809: (thin solid line). \label{tac}}
2810: \end{figure}
2811: %************************
2812: Once the field $\Phi_k$ has settled into one of the minima
2813: of $V_C$ the situation is classical again. At the minima we have
2814: %*******
2815: \eqb
2816: \label{minimacur}
2817: \left.\frac{\pd^2_{\theta_k} V_C(\Phi_k)}{|\Phi_k|^2}\right|_{\Phi^{\tiny\mbox{min}}_k}=
2818: \left.\frac{\pd^2_{|\Phi_k|} V_C(\Phi_k)}
2819: {|\Phi_k|^2}\right|_{\Phi^{\tiny\mbox{min}}_k}
2820: =2\mbox{N}^2>1\,.
2821: \eqe
2822: %*******
2823: The approximation $\La_C=\La_C^\prime$ used in Eq.\,(\ref{minimacur}) is exact for N=2,3.
2824: Since radial {\sl and} tangential quantum fluctuations are heavier than the
2825: compositeness scale $\Phi^{\tiny\mbox{min}}_k$ they are absent in the effective theory.
2826: As a consequence, the vanishing value of $V_C$ at the minima receives {\sl no}
2827: radiative corrections.
2828:
2829: In the limit $\mbox{N}\to\infty$ only the pole term of the potential
2830: $V_C$ survives for $|\Phi_k|<|\Phi^{\tiny\mbox{min}}_k|$ and
2831: thus we recover the situation of a global
2832: $U(1)^{\tiny\mbox{N}-1}$ symmetry. The $\tau$ dependence of the solutions to the BPS equation
2833: (\ref{PBScenter}) is then a pure phase and no
2834: tachyonic but only massless tangential fluctuations exist. These fluctuations
2835: lead to an instantaneous reheating at the phase boundary. The order parameter $|\Phi_k|$ jumps
2836: from zero to a finite value across the phase boundary. While the energy density of the
2837: ground-state jumps to zero at the phase boundary (BPS saturation),
2838: the ground-state pressure jumps to zero only after
2839: $|\Phi_k|=|\Phi^{\tiny\mbox{min}}_k|$ is reached.
2840:
2841: At finite N a flux quantum and a radial displacement $\Delta|\Phi_k|$
2842: are generated in each tunneling process. This reduces the ground-state energy
2843: density locally and brings the field $\Phi_k$ closer to one of the
2844: minima of $V_C$. The closer $\Phi_k$ to a minimum the less likely a
2845: tunneling process since the
2846: associated energy gain is quickly reduced ($|\pd_{|\Phi_k|}V_C|$ decreases!).
2847: If the number of tunneling processes taking place until
2848: relaxation would be roughly independent of N then
2849: for small N many unit-root directions will be reached in a relaxation process
2850: and thus it is likely that the
2851: average value of $\Phi_k$ close to the phase boundary is
2852: close to zero. For large N only a small angular sector
2853: would be reached by tunneling and thus close to the phase
2854: boundary the average value of $\Phi_k$ is not close to zero: a strong discontinuity of $\Phi_k$ across the
2855: transition should be observed.
2856:
2857: Closed center-vortex flux lines can self-intersect and in this way form
2858: isolated $Z_{\tiny\mbox{N}}$ monopoles which reverse the flux, see Fig.\,\ref{intersect} for the N=2 case.
2859: %***********************
2860: \begin{figure}
2861: \begin{center}
2862: \leavevmode
2863: %\epsfxsize=9.cm
2864: \leavevmode
2865: %\epsffile[80 25 534 344]{}
2866: \vspace{4.5cm}
2867: \special{psfile=Fig-18.ps angle=0 voffset=-120
2868: hoffset=-170 hscale=60 vscale=50}
2869: \end{center}
2870: \caption{The creation of an isolated $Z_2$ monopole by self-intersection
2871: of a $Z_2$ center vortex. \label{intersect}}
2872: \end{figure}
2873: %************************
2874: Let us discuss the case N=2 more explicitly.
2875: During the phase transition the process leading to self intersection can be performed $L$ times
2876: on a nonintersecting vortex loop. This generates $L$ isolated $Z_2$
2877: monopoles, each of mass $\sim \Lambda_C$, such that the mass $M_L$
2878: of the $L$-monopole state is $M_L\sim L\Lambda_C$. Since the number of possible
2879: $L$-monopole states roughly grows as $L!$
2880: (see \cite{BenderWu1976} for a more precise estimate) we conclude that the density of
2881: states $\rho(E)$ for the particle excitations in the center phase of an SU(2) theory
2882: is over-exponentially growing in energy $E$:
2883: %*********
2884: \eqb
2885: \label{rhoenergydesn}
2886: \rho(E)>\Lambda_C\exp\left[\frac{E}{\Lambda_C}\right]\,.
2887: \eqe
2888: %**********
2889: For N$>$2 the situation is similar. So we conclude that there exists a
2890: highest temperature $T_{\tiny\mbox{H}}\sim \Lambda_C$ in the center
2891: phase of an SU(N) Yang-Mills theory: a situation which was
2892: anticipated for any strongly interacting four dimensional field theory a
2893: long time ago \cite{Hagedorn1965}, see also \cite{HongLiu2004} for a discussion of the Hagedorn transition in
2894: an SU(N) matrix model. Moreover, we conclude that
2895: the center-magnetic phase transition can by no means be
2896: thermal (the spatial homogeneity of the system is violated during the transition)
2897: and thus that the thermodynamical pressure may jump across the transition.
2898: A single and a one-time self-intersecting center-vortex loop for N=2 are
2899: {\sl fermions} \cite{Hofmann20033}. This result is crucial for
2900: our understanding of the nature of leptons in the present Standard Model of
2901: Particle Physics. The time scale for the relaxation to one of the minima of $V_C$
2902: is roughly given by $|(\Phi^{\tiny\mbox{vac}}_k)|^{-1}$. A quantitative investigation
2903: of this reheating process would need methods of nonequilibrium field theory, see
2904: for example \cite{nonequ}. In a thermal approach it would be interesting
2905: to check in a lattice simulation of the {\sl effective theory}
2906: in the center phase Eq.\,(\ref{actcenter}) whether a density of states
2907: as in Eq.\,(\ref{rhoenergydesn}) is indeed seen.
2908:
2909: The fact that the pressure and the energy density of the ground state are
2910: precisely vanishing at the minima $\Phi^{\tiny\mbox{vac}}_k$ and that there
2911: are no radiative corrections to this situation clearly is of
2912: cosmological relevance.
2913:
2914:
2915: \section{Scale matching\label{MCC}}
2916:
2917: The scales $\La_E$ and $\La_M$, which determine the magnitudes of the
2918: adjoint Higgs field $\phi$ and the
2919: monopole condensates $\varphi_l$ at a given temperature in the electric and the magnetic phase,
2920: can be
2921: related by imposing the condition that the thermodynamical pressure $P$ be continuous
2922: across the thermal electric-magnetic phase transition. Disregarding the mismatch in the number of
2923: polarizations for some TLM modes in the electric
2924: phase and some dual gauge bosons in the magnetic phase, we derive
2925: %********
2926: \eqb
2927: \label{laElaM}
2928: \Lambda_E=(1/4)^{1/3}\Lambda_M\,\ \ \ \ (\mbox{N\ \ even})\,.
2929: \eqe
2930: %********
2931: For N=3 we have $\Lambda_E=(1/2)^{1/3}\Lambda_M$. The match between the
2932: magnetic and the center phase is less determined for the following reasons:
2933: (i) For N$>$3 the
2934: winding numbers of the fields $\Phi_i\,\ \ (i=\mbox{N}/2+1,\cdots,\mbox{N}-1)$
2935: close to the center-magnetic phase boundary are dynamically generated during the phase transition
2936: and so cannot be derived from the boundary conditions for
2937: the electric phase. For N=2 and N=3 the winding numbers of $\Phi_1$ and
2938: $(\Phi_1,\Phi_2)$ are unity, and we can set
2939: $\Lambda_C=\Lambda^\prime_C$ in Eq.\,(\ref{vCright}).
2940: (ii) An analytical description based on the potential $V_C$
2941: of the center-magnetic phase transition at finite N breaks down
2942: at the phase boundary, see Sec.\,(\ref{CVCM}).
2943:
2944: In the electric and in the magnetic phase (recall that the latter confines fundamental test charges)
2945: we encounter the thermodynamical analogue to dimensional transmutation in perturbation theory:
2946: Assuming MGSB, a single, fixed mass scale determines the thermodynamics of the SU(N) Yang-Mills
2947: theory in these phases. At a given temperature
2948: this mass scale can be experimentally inferred from the mass
2949: spectrum of the gauge boson fluctuations.
2950:
2951: \section{Pressure, energy density, and entropy density\label{PEEN}}
2952:
2953: \subsection{Numerical results}
2954:
2955: In this section we present our numerical results for one-loop
2956: temperature evolutions of thermodynamical potentials through the
2957: electric and magnetic phase. For the actual computations we consider N=2,3 only.
2958:
2959: \noindent In the electric phase the pressure $P$ divided by $T^4$ is given as
2960: %**********
2961: \eqb
2962: \label{pressureEP}
2963: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_E^4}
2964: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left\{2\,(\mbox{N}-1)\bar{P}(0)+
2965: 3\sum_{k=1}^{\tiny\mbox{N(N-1)}}\bar{P}(a_k)\right\}+
2966: \frac{\lambda_E}{2}\left(\frac{\mbox{N}}{2}+1\right)\mbox{N}\right]\,,\ \ (\mbox{N\ \ even})\,,
2967: \eqe
2968: %**********
2969: where the function $\bar{P}(a)$ and the dimensionless mass parameters $a_k$ are
2970: defined in Eqs.\,(\ref{P(a)}) and (\ref{dimlessdef}), respectively, and MGSB is assumed.
2971: In the magnetic phase we have
2972: %**********
2973: \eqb
2974: \label{pressureMP}
2975: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_M^4}
2976: \left[\frac{2\lambda_M^4}{(2\pi)^6}\left\{2\,(\frac{\mbox{N}}{2}-1)\bar{P}(0)+
2977: 3\sum_{l=1}^{\tiny\mbox{N}/2}\bar{P}(a_l)\right\}+
2978: \frac{\lambda_M}{16}\mbox{N}(\mbox{N}+2)\right]\,,\ \ (\mbox{N\ \ even})\,.
2979: \eqe
2980: %**********
2981: The dimensionless mass parameters $a_l$ are defined in
2982: Eqs.\,(\ref{massspecM}) and Eq.\,(\ref{defM})).
2983:
2984: For N=3 we have in the electric and the magnetic phase, respectively:
2985: %**********
2986: \eab
2987: \label{pressure3EPpressure3MP}
2988: \frac{P}{T^4}&=&-\frac{(2\pi)^4}{\lambda_E^4}
2989: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left\{4\bar{P}(0)+
2990: 3\left(4\,\bar{P}(a)+2\,\bar{P}(2a)\right)\right\}+
2991: 2\lambda_E\right]\,,\ \ (\mbox{N=3})\,,\nonumber\\
2992: \frac{P}{T^4}&=&-\frac{(2\pi)^4}{\lambda_M^4}
2993: \left[\frac{12\lambda_M^4}{(2\pi)^6}
2994: \bar{P}(a)+\lambda_M\right]\,,\ \ (\mbox{N=3})\,.
2995: \eae
2996: %**********
2997: The mass parameters $a_k$ and $a_l$ evolve with temperature according to the (inverted)
2998: solutions to Eqs.\,(\ref{eeq}), (\ref{eeq3}), (\ref{evolMeven}), and (\ref{evolM3}).
2999:
3000: Our results for $\frac{P}{T^4}$ as
3001: a function of temperature in the electric and
3002: magnetic phase are shown in Fig.\,\ref{pressure}.
3003: %***********************
3004: \begin{figure}
3005: \begin{center}
3006: \leavevmode
3007: %\epsfxsize=9.cm
3008: \leavevmode
3009: %\epsffile[80 25 534 344]{}
3010: \vspace{4.5cm}
3011: \special{psfile=Fig-10.ps angle=0 voffset=-160
3012: hoffset=-250 hscale=70 vscale=70}
3013: \end{center}
3014: \caption{$\frac{P}{T^4}$ as a function of temperature for N=2,3. The horizontal lines denote
3015: the respective asymptotic values. For N=2 we have $\lambda_{E,c}=11.65$
3016: and for N=3 $\lambda_{E,c}=8.08$. \label{pressure}}
3017: \end{figure}
3018: %************************
3019: %***********************
3020: \begin{figure}
3021: \begin{center}
3022: \leavevmode
3023: %\epsfxsize=9.cm
3024: \leavevmode
3025: %\epsffile[80 25 534 344]{}
3026: \vspace{4.5cm}
3027: \special{psfile=Fig-22.ps angle=0 voffset=-160
3028: hoffset=-180 hscale=70 vscale=70}
3029: \end{center}
3030: \caption{$\frac{P}{T^4}$ as a function of temperature for N=3 as obtained on a $16^3\times4$
3031: lattice using the differential method with a universal two-loop
3032: perturbative $\beta$ function \cite{Brown1988,Deng1988} and using the integral
3033: method (solid line) \cite{EngelsFingberg1990}. The figure is taken
3034: from \cite{EngelsFingberg1990}.
3035: \label{pressureLat}}
3036: \end{figure}
3037: %************************
3038: Notice that the pressure is negative
3039: in the electric phase close to $\lambda_{E,c}$ and
3040: even more so in the magnetic phase where the ground-state
3041: strongly dominates the thermodynamics.
3042:
3043: \noindent In the electric phase the energy density $\rho$ divided by $T^4$ is given as
3044: %**********
3045: %************
3046: \eqb
3047: \label{rhoEP}
3048: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_E^4}
3049: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left\{2\,(\mbox{N}-1)\bar{\rho}(0)+
3050: 3\sum_{k=1}^{\tiny\mbox{N(N-1)}}\bar{\rho}(a_k)\right\}+
3051: \frac{\lambda_E}{2}\left(\frac{\mbox{N}}{2}+1\right)\mbox{N}\right]\,,\ \ (\mbox{N\ \ even})\,,
3052: \eqe
3053: %**********
3054: where the function $\bar{\rho}(a)$ is defined as
3055: %**********
3056: \eqb
3057: \label{rhobar}
3058: \bar{\rho}(a)\equiv \int_{0}^{\infty}dx\,x^2 \frac{\sqrt{x^2+a^2}}{\exp(\sqrt{x^2+a^2})-1}\,.
3059: \eqe
3060: %**********
3061: In the magnetic phase we have
3062: %**********
3063: \eqb
3064: \label{rhoMP}
3065: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_M^4}
3066: \left[\frac{2\lambda_M^4}{(2\pi)^6}\left\{2\,(\frac{\mbox{N}}{2}-1)\bar{\rho}(0)+
3067: 3\sum_{l=1}^{\frac{\tiny\mbox{N}}{2}}\bar{\rho}(a_l)\right\}+
3068: \frac{\lambda_M}{16}\mbox{N}(\mbox{N}+2)\right]\,,\ \ (\mbox{N\ \ even})\,.
3069: \eqe
3070: %**********
3071: For N=3 we have in the electric and the magnetic phase, respectively:
3072: %**********
3073: \eab
3074: \label{rho3EPrho3MP}
3075: \frac{\rho}{T^4}&=&\frac{(2\pi)^4}{\lambda_E^4}
3076: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left\{4\bar{\rho}(0)+
3077: 3\left(4\,\bar{\rho}(a)+2\,\bar{\rho}(2a)\right)\right\}+
3078: 2\lambda_E\right]\,,\ \ (\mbox{N=3})\,,\nonumber\\
3079: \frac{\rho}{T^4}&=&\frac{(2\pi)^4}{\lambda_M^4}
3080: \left[\frac{12\lambda_M^4}{(2\pi)^6}
3081: \bar{\rho}(a)+\lambda_M\right]\,,\ \ (\mbox{N=3})\,.
3082: \eae
3083: %**********
3084: Our results for $\frac{\rho}{T^4}$ as
3085: a function of temperature in the electric and
3086: magnetic phase are shown in Fig.\,\ref{rho}. Slight discontinuities at $\lambda_{E,c}$ are seen.
3087: This is explained by the mismatch in the number of
3088: polarizations of fluctuating gauge bosons across the electric-magnetic transition -
3089: an approximation to scale-matching -
3090: and the fact that continuity in $P$ does not imply continuity in $\rho$. Again,
3091: the energy density is dominated by the
3092: ground-state contribution in the electric phase close to the electric
3093: magnetic transition and even more so in the magnetic
3094: phase.
3095: %***********************
3096: \begin{figure}
3097: \begin{center}
3098: \leavevmode
3099: %\epsfxsize=9.cm
3100: \leavevmode
3101: %\epsffile[80 25 534 344]{}
3102: \vspace{4.5cm}
3103: \special{psfile=Fig-11.ps angle=0 voffset=-150
3104: hoffset=-250 hscale=70 vscale=70}
3105: \end{center}
3106: \caption{$\frac{\rho}{T^4}$ as a function of temperature for N=2,3. The horizontal lines denote
3107: the respective asymptotic values. \label{rho}}
3108: \end{figure}
3109: %************************
3110: The entropy density $S$ is defined as the derivative of the
3111: pressure with respect to temperature:
3112: %*********
3113: \eqb
3114: \label{Sdef}
3115: S\equiv\frac{dP}{dT}\,.
3116: \eqe
3117: %*********
3118: Using the thermodynamical relation $\rho=T\frac{dP}{dT}-P$, we may write
3119: %********
3120: \eqb
3121: \label{sdef}
3122: \frac{S}{T^3}=\frac{1}{T^4}\left(\rho+P\right)\,.
3123: \eqe
3124: %********
3125: Our results for $S/T^3$ are shown in Fig.\,\ref{ST3}.
3126: The reasons for the slight discontinuities at $\lambda_{E,c}$ are the same as in the
3127: case $\frac{\rho}{T^4}$. The entropy density $S$ is a measure for the `mobility' of gauge
3128: boson excitations. That $S$ is zero at the critical
3129: temperature $\lambda_{M,c}$ for the center transition
3130: is explained by the fact that all dual gauge
3131: bosons acquire an infinite mass there and thus no fluctuating degrees
3132: of freedom are left in the thermodynamical balance.
3133:
3134:
3135: \subsection{Comparison with the lattice\label{complat}}
3136:
3137: An early lattice measurements of the energy density $\rho$
3138: and the interaction measure $\Delta\equiv \rho-3P$ in a pure SU(2) gauge theory
3139: were reported on in \cite{EnKaSaMo1982}. In that work the critical temperature $T_c$ for
3140: the deconfinement transition was determined from the critical behavior of the
3141: Polyakov-loop expectation and the peak position
3142: of the specific heat using a Wilson action. The function $\Delta(T)$ was extracted by multiplying the
3143: lattice $\beta$ function with the difference of plaquette expectations at finite and zero temperature.
3144: This assures that $\Delta$ vanishes as $T\to 0$. What is subtracted at finite $T$ is, however,
3145: {\sl not} the value $\Delta(T=0)$ since the associated plaquette expectation is
3146: multiplied with the value of the $\beta$ function at {\sl finite} $T$. Apart from this
3147: approximation, the use of a perturbative $\beta$ function was assumed for all
3148: temperatures.
3149: %***********************
3150: \begin{figure}
3151: \begin{center}
3152: \leavevmode
3153: %\epsfxsize=9.cm
3154: \leavevmode
3155: %\epsffile[80 25 534 344]{}
3156: \vspace{5.0cm}
3157: \special{psfile=Fig-12.ps angle=0 voffset=-150
3158: hoffset=-250 hscale=70 vscale=70}
3159: \end{center}
3160: \caption{$\frac{S}{T^3}$ as a function of temperature for N=2,3. The horizontal lines denote
3161: the respective asymptotic values. \label{ST3}}
3162: \end{figure}
3163: %************************
3164: The lattice results for $\rho$ differ drastically from
3165: our results for temperatures close the deconfinement, that is, the electric-magnetic
3166: transition. We obtain
3167: %*****
3168: \eqb
3169: \label{rhooverrhoSBc}
3170: \left.\frac{\rho}{\rho_{SB}}\right|_{T\sim T_{E,c}}\sim 1.49
3171: \eqe
3172: %*****
3173: where $\rho_B\equiv\frac{\pi^2}{5}T^4$ denotes the Stefan-Boltzmann limit
3174: (ideal gas of massless gluons with two polarizations). On
3175: the lattice, this ratio is measured to be
3176: smaller than unity. At $T\sim 5T_{E,c}$ we obtain
3177: %*****
3178: \eqb
3179: \label{rhooverrhoSB5}
3180: \left.\frac{\rho}{\rho_{SB}}\right|_{T\sim 5\,T_{E,c}}\sim 1.34\,
3181: \eqe
3182: %*****
3183: while the lattice measures a ratio of about 0.85. Our asymptotic
3184: value\footnote{The asymptotic temperature $\lambda_{E,as}=75$ is determined
3185: by the boundary condition $\lambda_E(0)=1000$ for solving the evolution equations
3186: (\ref{eeq}) and (\ref{eeq3}).} is
3187: %*****
3188: \eqb
3189: \label{rhooverrhoSBa}
3190: \left.\frac{\rho}{\rho_{SB}}\right|_{T\sim 6.4\,T_{E,c}}\sim 1.33\,.
3191: \eqe
3192: %*****
3193: Notice the latent heat released at the electric magnetic transition for N=3
3194: (Fig.\,\ref{rho}).
3195:
3196: Our result for the pressure $P$ indicates negative values for $T$
3197: close to $T_{E,c}$ (see Fig.\, \ref{pressure}) - much in
3198: contrast to the positive values obtained
3199: in \cite{EnKaSaMo1982}. At $T\sim 5 T_{E,c}$ we obtain
3200: %*****
3201: \eqb
3202: \label{PoverrhoSB5}
3203: \left.\frac{P}{P_{SB}}\right|_{T\sim 5\,T_{E,c}}\sim 1.30\,
3204: \eqe
3205: %*****
3206: while the lattice measures ($P$ is extracted
3207: from the measured values of
3208: $\Delta$ and $\rho$) a ratio of about 0.88. Our asymptotic
3209: value for $P$ is
3210: %*****
3211: \eqb
3212: \label{PoverrhoSBa}
3213: \left.\frac{P}{P_{SB}}\right|_{T\sim 6.4\,T_{E,c}}\sim 1.32\,.
3214: \eqe
3215: %*****
3216: The asymptotic values of Eqs.\,(\ref{rhooverrhoSBa}) and
3217: (\ref{PoverrhoSBa}) are very close to the ratio $R$ of the number of polarizations
3218: for massive TLH modes and massless TLM modes and the
3219: number of polarizations for massless gluon modes with two polarizations:
3220: %*************
3221: \eqb
3222: \label{R}
3223: R=\frac{2\times 3+1\times 2}{3\times 2}=\frac{4}{3}\sim 1.33\,.
3224: \eqe
3225: %************
3226: Indeed, at $\lambda_E\sim 75$ the mass
3227: parameter $a$, defined in Eq.\,(\ref{massP}), is
3228: $a\sim 2\pi\frac{5.5}{650}\sim 0.053$ and
3229: therefore the Boltzmann suppression of TLH modes is small.
3230: At extremely high temperatures a TLH mode `remembers' its massiveness
3231: at low temperatures in terms of an extra polarization coming from
3232: a tiny mass which, however, still solves the infrared problem
3233: of naive perturbation theory.
3234:
3235: For a comparison with N=3 lattice data we use the results obtained with a Wilson action in
3236: \cite{Bielfeld1996} on the lattice of the largest
3237: time extension, $N_\beta=8$. In the vicinity of the transition point $T_{E,c}$
3238: the situation for both $\rho$ and $P$ is similar as for N=2:
3239: drastic differences between the lattice measurements and
3240: our results occur. Again, $P$ comes out negative in our calculation,
3241: contradicting the positive values obtained in
3242: \cite{Bielfeld1996}. It should be remarked at this point that a
3243: lattice simulation of $P$, which did not rely on the integral method (see below)
3244: as it was used in \cite{Bielfeld1996}, has seen negative
3245: pressure for $T$ not far above $T_{c,e}$ \cite{Deng1988}. Moreover,
3246: the most negative value of $P/T^4\sim -0.5$ obtained
3247: in \cite{Deng1988} very close to the phase transition coincides
3248: with our result at the electric-magnetic transition,
3249: see Figs.\,\ref{pressure} and \ref{pressureLat}.
3250:
3251:
3252: \noindent At $T=5\,T_{E,c}$ we have
3253: %**********
3254: \eqb
3255: \label{rhoN3}
3256: \left.\frac{\rho}{\rho_{SB}}\right|_{T\sim 5\,T_{E,c}}\sim 1.38
3257: \eqe
3258: %**********
3259: while the lattice measures a ratio of about 0.93. Our asymptotic
3260: value for $\rho$ is
3261: %*****
3262: \eqb
3263: \label{rhooverrhoSB3a}
3264: \left.\frac{\rho}{\rho_{SB}}\right|_{T\sim 8.86\,T_{E,c}}\sim 1.37\,.
3265: \eqe
3266: %*****
3267: For the pressure $P$ we obtain at $T=5\,T_{E,c}$:
3268: %*****
3269: \eqb
3270: \label{PoverrhoSB35}
3271: \left.\frac{P}{P_{SB}}\right|_{T\sim 5\,T_{E,c}}\sim 1.34\,
3272: \eqe
3273: %*****
3274: while the lattice measures a ratio of about 0.97. Our asymptotic
3275: value for $P$ is
3276: %*****
3277: \eqb
3278: \label{PoverrhoSB3a}
3279: \left.\frac{P}{P_{SB}}\right|_{T\sim 8.86\,T_{E,c}}\sim 1.36\,.
3280: \eqe
3281: %*****
3282: Both asymptotic values in Eqs.\,(\ref{rhooverrhoSB3a}) and
3283: (\ref{PoverrhoSB3a}) are very close to the ratio of
3284: polarizations $R=\frac{11}{8}=1.375$.
3285:
3286: According to Fig.\,(\ref{ST3}) the entropy
3287: density $\frac{S}{T^3}$ vanishes at $T_{M,c}$. In our approach
3288: this reflects the fact that at $T_{M,c}$ all gauge-field are thermodynamically decoupled
3289: because of their infinite mass. As a consequence,
3290: thermodynamics is entirely determined by the ground state.
3291: This is not observed in the lattice simulations \cite{EngelsKarschScheideler1999}
3292: for both N=2,3 where a continuous behavior
3293: of $\frac{S}{T^3}$ across the deconfinement transition at $T_{E,c}$ was obtained.
3294:
3295: In \cite{Brown1988} a discontinuous behavior of $\frac{S}{T^3}$ was observed for N=3
3296: using a Wilson action and a perturbative beta function.
3297: There is an excellent agreement of their
3298: result with our result for temperatures ranging $T_{E,c}$ down to $T_{M,c}$,
3299: compare Figs.\,\ref{ST3} and \ref{ST3lat}.
3300: %***********************
3301: \begin{figure}
3302: \begin{center}
3303: \leavevmode
3304: %\epsfxsize=9.cm
3305: \leavevmode
3306: %\epsffile[80 25 534 344]{}
3307: \vspace{5.0cm}
3308: \special{psfile=Fig-13.ps angle=0 voffset=-150
3309: hoffset=-180 hscale=50 vscale=50}
3310: \end{center}
3311: \caption{$\frac{S}{T^3}$ as a function of $\beta$
3312: obtained in SU(3) lattice gauge theory using the differential method and a
3313: perturbative beta function. The simulations were performed on (a) $16^3\times 4$, (b)
3314: $24^3\times 4$, (c) $16^3\times 6$ (open circles) and $20^3\times 6$ (closed circles), and
3315: (d) $24^3\times 6$ lattices. Using the $24^3\times 6$ lattice, the critical
3316: value of $\beta$ is between 5.8875 and 5.90. \label{ST3lat}}
3317: \end{figure}
3318: %************************
3319: The almost discontinuous
3320: behavior in Fig.\,(\ref{ST3}) is due to the large rise
3321: of the magnetic gauge coupling with decreasing temperature.
3322: According to Fig.\,(\ref{gevol}) the `duration' $\delta T\equiv\frac{T_{E,c}-T_{M,c}}{T_{E,c}}$
3323: of the magnetic phase is only $\delta T\sim 0.1$. In a lattice simulation the resolution of
3324: such a small temperature interval depends very much on the choice of the beta function.
3325: In \cite{Brown1988} a universal perturbative beta function was used which may have lead to interprete
3326: the behavior of $\frac{S}{T^3}$ as discontinuous in
3327: dependence on temperature. Since $\frac{S}{T^3}$ is a
3328: quantity which is much less sensitive to the infrared physics
3329: than, say, $\frac{P}{T^4}$ (the direct
3330: contribution of the ground state is canceled out in $\frac{S}{T^3}$) the use of (the universal part of) a
3331: perturbative beta function
3332: in the lattice simulations \cite{Brown1988} may be justified. We stress at this point
3333: that lattice simulations based on lattice sizes of 1-3 times the inverse Yang-Mills scale
3334: are not capable of being sensitive to the infrared effects of the theory which
3335: have correlation lengths of the order of the gauge couplings $e$ and $g$ times the inverse Yang-Mills
3336: scale at decoupling (Standard Model physics in the electroweak sector
3337: suggest that $e_{\tiny\mbox{decoup}}\sim g_{\tiny\mbox{decoup}}\sim 10^6\,$!).
3338:
3339: The alert reader would
3340: object that the thermodynamical relation
3341: %**********
3342: \eqb
3343: \label{pdt}
3344: dP=S\,dT\,,
3345: \eqe
3346: %**********
3347: which implies that in a homogeneous thermal system the pressure has to be a
3348: monotonic function of temperature, is violated at the center transition ($T_{c,M}$), see
3349: Fig.\,\ref{pressure}, where with decreasing temperature
3350: the pressure quickly rises from a negative to a value close to zero
3351: \footnote{This effect should be the
3352: more pronounced the higher N is \cite{LuciniTeperWenger2003}.}. What is the resolution of
3353: this puzzle? There are two anwers. First, on a microscopic level, the
3354: homogeneity of the system at the point where center
3355: vortices start to condense is badly violated by the generation of
3356: (intersecting) center-vortex loops through discontinuous and local
3357: phase changes of the fields $\Phi_k(x)$
3358: in Eq.\,(\ref{Wilsonloopx}). The derivation of Eq.\,(\ref{pdt})
3359: from the partition function, however, relies on homogeneity.
3360: Second, assuming homgeneity, one can easily convinces oneself that
3361: the spectral density $\rho(E)$ in the center phase is
3362: exponentially increasing with energy $E$\footnote{The number
3363: of center vortex loops with $L$ intersections increases stronger than
3364: factorially with $L$ and the energy of a vortex state with $L$ intersections
3365: is $\propto L\Lambda_C$, for the SU(2) case see \cite{BenderWu1976} where a counting of the vacuum
3366: diagrams in a $\lambda\phi^4$ theory is carried out.}:
3367: %*******
3368: \eqb
3369: \label{specdens}
3370: \rho(E)\propto\exp[\frac{E}{T_{\tiny\mbox{H}}}]\,.
3371: \eqe
3372: %*******
3373: Thus the partition function diverges at $T=T_{\tiny\mbox{H}}\sim T_{c,M}$,
3374: the homogeneous system would need an infinite amount of energy per volume
3375: to increase the temperature beyond $T_{\tiny\mbox{H}}$. We conclude that
3376: homogeneity is violated at $T=T_{\tiny\mbox{H}}$. We conclude that the SU(N) YM dynamics
3377: indeed predicts a violation of the thermodynamical relation in Eq.\,(\ref{pdt})
3378: at $T=T_{\tiny\mbox{H}}\sim T_{c,M}$ \footnote{The author would like to thank
3379: D. T. Son for initiating this discussion.}.
3380:
3381: What are the possible reasons for the qualitative difference between the results obtained
3382: in \cite{Bielfeld1996,EngelsKarschScheideler1999} and \cite{Brown1988,Deng1988}? To avoid the use
3383: of derivatives of the bare coupling, which are multiplying the sum of spatial
3384: and time plaquette averages in the
3385: differential method for the computation of the pressure, the integral method was introduced in
3386: \cite{EngelsFingberg1990}. Using a perturbative beta function
3387: in the differential method, negative values for the pressure for $T$ close to
3388: $T_c$ and a rapid approach of the $\rho$ and
3389: $P$ to their respective ideal gas limits were obtained in \cite{Deng1988}.
3390: The rapid approach to the ideal gas limit in $\rho$ is also obtained in the present approach,
3391: compare Eqs.\,(\ref{rhooverrhoSB5}),
3392: (\ref{rhooverrhoSBa}) and Eqs.\,(\ref{rhoN3}),(\ref{rhooverrhoSB3a}).
3393: At the time when the results in
3394: \cite{Brown1988,Deng1988} appeared a negative pressure was regarded as
3395: unphysical and attributed to the use of a
3396: perturbative beta function. The integral method proposed in \cite{EngelsFingberg1990}
3397: solely generates positive pressure. This method {\sl assumes} the validity of the
3398: thermodynamical limit in the lattice simulation. Namely,
3399: for large spatial volume the thermodynamical relation
3400: %*********
3401: \eqb
3402: \label{ptd}
3403: P=T\frac{\pd\ln Z}{\pd V}
3404: \eqe
3405: %*********
3406: valid for a finite volume $V$ is approximated by
3407: %*********
3408: \eqb
3409: \label{ptdL}
3410: P=T\frac{\ln Z}{V}\,
3411: \eqe
3412: %*********
3413: so that the pressure equals minus the free energy density.
3414: In Eqs.\,(\ref{ptd}) and (\ref{ptdL}) $Z$ denotes the partition function.
3415: Based on Eq.\,(\ref{ptdL}) the derivative of the
3416: pressure with respect to the bare coupling $\beta=\frac{2\mbox{N}}{\bar{g}^2}$
3417: can be expressed as an expectation over the sum of spatial and time plaquettes
3418: without the beta-function prefactor. Thus the pressure can be obtained
3419: by an integral over $\beta$ up to an unknown integration constant. The latter
3420: is chosen in such a way that the pressure vanishes at a temperature well
3421: below $T_c$. Instead of only integrating over minus the sum of spatial and
3422: time plaquette expectations twice
3423: the plaquette expectation at $T=0$
3424: was added to the integrand in \cite{Bielfeld1996,EngelsKarschScheideler1999}
3425: to assure that the pressure vanishes at $T=0$. We would like to stress that
3426: this prescription does not follow from relation (\ref{ptdL}).
3427: Thus it seems to be natural that integral and differential methods
3428: generate qualitatively different results. The results for $P(T)$ obtained by the
3429: integral method show a rather large dependence on the cutoff and the time extent $N_\tau$ of the lattice
3430: \cite{Bielfeld1996}. We believe that this
3431: reflects the considerable deviation from the assumed
3432: thermodynamical limit for so-far available lattice sizes. The problem
3433: was addressed in \cite{EngelsKarschScheideler1999} where a
3434: correction factor $r$ was introduced to relate $P$
3435: obtained with the integral method to $P$ obtained
3436: with the differential method. For a given $N_\tau$
3437: the factor $r$ was determined from the pressure ratio at $\bar{g}=0$.
3438: Subsequently, this value of $r$ was used at finite coupling $\bar{g}$ to extract the spatial anisotropy
3439: coefficient $c_\sigma$ (essentially the beta function)
3440: by demanding the equality of the pressure obtained
3441: with the integral and the differential method.
3442: In doing so, twice the plaquette expectation at $T=0$ was, again, added to minus
3443: the sum of spatial and time plaquette expectations in the expression
3444: for the pressure using the differential method. It may be questioned
3445: whether a simple multiplicative correction $r$ does correctly
3446: account for finite-size and cutoff effects and, if yes,
3447: whether it is justified to determine $r$ in the limit
3448: of noninteracting gluons\footnote{The $c_\sigma$-values obtained in this way
3449: do not coincide with those obtained in \cite{Klassen1998}.}.
3450:
3451: To summarize, we observe a qualitative disagreement
3452: between the lattice results \cite{EnKaSaMo1982,Bielfeld1996} for $\rho$ and $P$, using the integral method,
3453: and the results of the present approach
3454: for temperatures close to $T_{E,c}$. At $T\sim 5\,T_{E,c}$ there is better agreement.
3455: Although for N=3 our negative values for $P$ in the vicinity of $T_{E,c}$
3456: disagree with those obtained with the differential method
3457: (we obtain a modulus at $T_{E,c}$ which {\sl coincides} with that of the minimal value obtained by using a
3458: perturbative beta-function differential method \cite{Deng1988}). On the other hand,
3459: the entropy density obtained in \cite{Brown1988} with the differential method
3460: agrees well with our results. This is expected since the entropy
3461: is an infrared insensitive quantity.
3462:
3463:
3464: \section{Conclusions and Outlook\label{CO}}
3465:
3466: We have developed a nonperturbative approach to SU(N) Yang-Mills thermodynamics
3467: which is based on the (self-consistent) assumption
3468: that the theory `condenses' SU(2) embedded, BPS saturated
3469: topological fluctuations of trivial holonomy
3470: at an asymptotically high temperature. In \cite{HerbstHofmann2004} we have shown
3471: for the SU(2) case the redundancy of this assumption.
3472: The concept for the construction of an effective theory based
3473: on the above assumption is similar to that applied to the construction of
3474: the macroscopic field theory for superconductivity \cite{GinzburgLandau1950,Abrikosov1957}.
3475: We stress that the effects on nontrivial topology die
3476: off in a power-like fashion in the effective theory
3477: (as a function of temperature). Thus the perturbatively
3478: derived suppression of topologically nontrivial field
3479: configurations does take place in our effective theory as well.
3480:
3481: We have constructed a (uniquely determined) potential for the thermodynamics of an
3482: energy- and pressure-free adjoint Higgs background $\phi$
3483: which macroscopically describes the collective effects due to noninteracting,
3484: trivial-holonomy calorons in the only deconfining phase of the
3485: theory which we call electric phase for obvious reasons. As a consequence,
3486: the ground state of the system is described by a
3487: BPS saturated solution to the field equation of the scalar
3488: sector and an associated pure-gauge configuration solving the macroscopic
3489: equation of motion for trivial-topology gauge-field fluctuations.
3490: The latter has macroscopic, nontrivial holonomy and
3491: thus describes the presence of isolated magnetic monopoles being generated
3492: from decaying nontrivial-holonomy calorons as a result of microscopic
3493: interactions between trivial-holonomy calorons. These interactions are mediated
3494: by the trivial-topology sector of the theory. The modulus of $\phi$ falls off as
3495: $\propto T^{-1/2}$ and the ground-state pressure is only linear in $T$ so that
3496: nonperturbative effects (apart from extra polarizations for excitations)
3497: are {\sl irrelevant} at asymptotically high temperatures.
3498:
3499: Some of the topologically trivial gauge-field fluctuations
3500: are massive on tree-level due to the adjoint Higgs mechanism, and
3501: the underlying SU(N) gauge symmetry is
3502: spontaneously broken accordingly. An evolution equation, describing the temperature
3503: dependence of the effective gauge coupling constant $e$, was obtained from the requirement of thermodynamical
3504: self-consistency of the effective theory. The two fixed points of this
3505: evolution were identified. These fixed points
3506: predict the existence of a highest and a lowest
3507: attainable temperature in the electric phase.
3508: Based on the evolution $e(T)$ a physical argument was given why caloron
3509: `condensation' in a grandly unifying theory must take place at a temperature close
3510: to the cutoff scale for validity of a local, four dimensional
3511: field-theory description. We have investigated some aspects of the loop expansion of
3512: thermodynamical potentials in the effective electric theory. Our conclusion
3513: is that the present nonperturbative approach resolves the infrared problems associated with the usual,
3514: perturbative loop expansion. We investigate the two-loop corrections
3515: to the pressure for N=2 in \cite{HerbstHofmannRohrer2004}. Two-loop contributions to the
3516: pressure are corrections
3517: to the one-loop contributions which
3518: range within the $0.1$\% level. Thus as far a bulk thermodynamical quantities are concerned
3519: the gauge-boson fluctuations in the electric fields are
3520: practically {\sl noninteracting} at the expense of
3521: some of them being thermal quasiparticles.
3522:
3523: The downward temperature evolution of the effective gauge coupling $e$ has an attractor and
3524: thus the IR-UV decoupling observed in the underlying theory due to
3525: renormalizability is recovered in the effective theory. The temperature evolution $e(T)$
3526: predicts a transition, driven by the
3527: condensation of magnetic monopoles, to a phase with less gauge symmetry
3528: (magnetic phase, confining heavy fundamental test charges). This transition
3529: is the deconfinement-confinement transition identified in lattice simulations. Due to the
3530: typical correlation length in the monopole condensate being of the order of $e_{\tiny\mbox{decoup}}\times
3531: \Lambda^{-1}$, where $\Lambda$ denotes the Yang-Mills scale and
3532: $e_{\tiny\mbox{decoup}}\gg 1$, it is very hard
3533: (in practice impossible) for lattice simulations performed within the magnetic phase to
3534: predict thermodynamical quantities that are
3535: infrared sensitive.
3536:
3537: The macroscopic ground-state structure of the
3538: magnetic phase is determined in analogy to the electric phase, and,
3539: as a consequence, some of the residual (dual)
3540: gauge-field fluctuations acquire mass by the Abelian Higgs mechanism. The evolution of the
3541: magnetic gauge coupling constant $g$, again being a consequence of
3542: thermodynamical self-consistency, predicts
3543: the existence of a highest and a lowest attainable temperature
3544: also for the magnetic phase. At the lowest temperature a transition to the center phase,
3545: where center-vortex loops are condensed into the ground state, takes place. Assuming maximal gauge-symmetry
3546: breaking in the electric phase, {\sl all} gauge-field modes decouple thermodynamically at the
3547: transition point, and at this point the equation of state is maximally negative, $P=-\rho$.
3548:
3549: The magnetic-center transiton is of
3550: the Hagedorn type and thus nonthermal. A remarkable feature of the center phase is that
3551: ground-state pressure and energy density are {\sl precisely} zero after a period of rapid
3552: reheating has taken place. This follows from the shape of the effective potential $V_C$ for the
3553: vortex-condensate fields $\Phi_k$ which implements the spontaneous
3554: breakdown of the local center symmetry $Z_{\mbox{N,mag}}$.
3555:
3556: In the limit N$\to\infty$ analytical access to the center-vortex dynamics is granted by the
3557: thermodynamical and quantum mechanical stability of the
3558: classical solutions to the BPS equations for the center-vortex fields.
3559: For finite N this stability is lost
3560: due to the presence of tachyonic modes if the center-vortex condensates $\Phi_k$
3561: are away from the minima of their potential. Once the
3562: minima are reached, no fluctuations in $\Phi_k$ exist for any $\mbox{N}\ge 2$,
3563: and thus the result of a vanishing
3564: ground-state energy and pressure, which is based on the
3565: classical analysis, is strictly reliable.
3566:
3567: For N=2,3 the present approach predicts a Stefan-Boltzmann
3568: like behavior (with additional polarizations) of the thermodynamical potentials pressure, energy density,
3569: and entropy density at temperatures of about $10\,T_{E,c}$.
3570: Throughout the magnetic phase we predict a {\sl negative}
3571: equation of state which is in contradiction to lattice results for N=2,3
3572: using the integral method. For the (infrared insensitive) entropy density at N=3
3573: we obtain excellent agreement with lattice data generated with the
3574: differential method and a perturbative beta function.
3575:
3576: There are many applications of the approach presented in this paper. In \cite{Hofmann20032} we have
3577: proposed that a strongly interacting gauge theory underlying the leptonic sector of the Standard Model
3578: should be based on the following gauge group
3579: %************
3580: \eqb
3581: \label{SIGUE}
3582: SU(2)_{\tiny\mbox{CMB}}\times SU(2)_{e}\times SU(2)_{\mu}\times SU(2)_{\tau}\cdots\,.
3583: \eqe
3584: %***********
3585: In addition, mixing angles for the gauge bosons of one factor with
3586: those of the other factors at temperatures much higher than the Yang-Mills
3587: scale of the last factor should be supplied. In Eq.\,(\ref{SIGUE}) the Yang-Mills scale of the first factor is
3588: roughly given (but can be precisely determined
3589: \cite{Hofmann20032}) by the temperature $T_{\tiny\mbox{CMB}}$
3590: of the cosmic microwave background $\sim 10^{-4}\,\mbox{eV}$. The
3591: Yang-Mills scales of the other factors are roughly given by the mass of the corresponding charged leptons.
3592: While the CMB-scale theory is in its magnetic phase
3593: very close to the magnetic-electric transition (only there is the dual gauge boson, the photon, massless)
3594: the other theories are in their center phases generating the stable leptons as solitons (neutrino ... single
3595: center-vortex loop, charged lepton ... center-vortex loop with one self-intersection.).
3596: The two polarization states of these solitons
3597: arise as a consequence of the spontaneously broken, local $Z_{2,\tiny\mbox{mag}}$ symmetry. The photon,
3598: which is the fluctuating degree of freedom in the magnetic phase of
3599: the CMB-scale theory, would couple to the electric charges and the
3600: magnetic moments of these leptons because the gauge dynamics subject to Eq.\,(\ref{SIGUE}) was
3601: embedded into the gauge dynamics subject to a higher gauge group SU(N) at temperatures
3602: larger than the mass of the heaviest charged lepton. The same reasoning goes through
3603: for the coupling of the photon to quarks if we allow for SU(3) factors in Eq.\,(\ref{SIGUE}).
3604: Given the mixing angles it is possible to compute the
3605: electric charge of each soliton from the plateau value of the gauge-coupling evolution in the
3606: electric phase in each factor theory. Since the CMB-scale theory is in its magnetic phase at the magnetic-electric transition
3607: (a very small magnetic gauge coupling constant $g$) the ground state of the universe at present
3608: is slightly superconducting: a possible explanation for
3609: intergalactic magnetic fields. The ground-state energy density due to
3610: the CMB-scale theory is about 1\% of the gravitationally observed value \cite{WMAP2003,Hofmann20032}, no
3611: contribution. Recall, that no ground-state energy density of pressure is generated by
3612: SU(N) Yang-Mills theories in their center phases. We believe that the missing part can be
3613: linked to a CP violating, additional
3614: term in the Yang-Mills action \cite{Wilczek2004}, see also \cite{AlexanderMoffat2004}
3615: for another intelligent (albeit incomplete) way of addressing
3616: the cosmological constant `problem'. That the temperature of the
3617: Universe is stabilized at $T=T_{\tiny\mbox{CMB}}$ follows from the behavior of the energy density $\rho$ at the
3618: electric-magnetic transition, see Fig.\,\ref{rho}. This is an extraordinary useful fact
3619: since it allows for our mere existence \footnote{Biology would be unthinkable
3620: with a massive photon.}. The decoupled $W^\pm$ bosons of
3621: SU(2)$_{\tiny\mbox{CMB}}$ are stable since they cannot decay into the matter that would
3622: arise if SU(2)$_{\tiny\mbox{CMB}}$ would go into its center phase (a disaster for
3623: entropy generating individuals). They are an extraordinarily viable candidate for clustering dark matter (the stuff
3624: responsible for the anomalous rotation curves of galaxies).
3625:
3626: The $Z_0$ and the $W_{\pm}$
3627: bosons of the Standard Model would be interpreted as the thermodynamically
3628: decoupled dual and TLH gauge-boson
3629: fluctuations of the $SU(2)_{e}$ factor in Eq.\,(\ref{SIGUE}). One would expect to
3630: see heavy gauge bosons $Z^\prime_0$ and $W^\prime_{\pm}$, arising
3631: from the factor $SU(2)_{\mu}$ in Eq.\,(\ref{SIGUE}), at about $\frac{m_\mu}{m_e}\sim 200$
3632: times the mass of the $Z_0$ and the $W_{\pm}$ bosons, respectively. There is no {\sl fluctuating}
3633: Higgs-field in this (stepwise) description of electroweak symmetry breaking. Obviously,
3634: a lot of the extraordinarily precisely checked features of
3635: the Standard Model can not be derived at the present stage of development,
3636: for example the absence of flavor-changing neutral currents
3637: on tree-level. Moreover, it is
3638: questionable that scattering processes like $e^+e^-\to e^+e^-$, say, at the $Z_0$ resonance
3639: can be well understood in a thermodynamical framework.
3640: It is, however, conspicuous that the total cross section
3641: of this process deviates substantially from the QED prediction for $\sqrt{s}\sim m_e, m_\mu$
3642: \cite{QEDtests}. The apparent structurelessness of a charged lepton as measured
3643: for momentum transfers away from its mass
3644: may be understood by the over-exponentially rising density of (instable)
3645: states in the center phase of the SU(2)$_e$, SU(2)$_\mu$, ... Yang-Mills theories in Eq.\,(\ref{SIGUE}).
3646:
3647: To make contact with ultra-relativistic heavy ion collision our approach to pure SU(3) Yang-Mills
3648: theory would have to be extended to include fundamentally charged fermions. The assumption of
3649: rapid thermalization, which underlies the (very successful) hydrodynamical approach to
3650: the early stages of an ultrarelativistic heavy-ion collision
3651: \cite{ShuryakTeaney2001}, would be explained
3652: by rigid correlations in the magnetic phase (magnetic monopole condensate)
3653: or the electric phase (caloron `condensate') if the ground-state structure
3654: of a pure SU(3) gauge theory would not
3655: significantly be altered by the presence of quarks.
3656:
3657: To describe thermalized Quantum Chromodynamics
3658: one would introduce quarks as fundamental fermionic fields
3659: $\psi_i$ where $i$ is a flavor index and the color index is implicit.
3660: Quarks may couple to the caloron `condensate' $\phi$
3661: in the electric phase via Yukawa terms
3662: %*********
3663: \eqb
3664: \label{Yuk}
3665: \kappa\sum_i \bar{\psi}_i\phi\psi_i
3666: \eqe
3667: %********
3668: and to topologically trivial gauge-field fluctuation $\delta a_\rho$
3669: via the usual covariant derivative. Due to
3670: Eq.\,(\ref{Yuk}) quarks acquire mass dynamically. The ground-state structure in this approach
3671: would be the same as the one of a pure SU(3)
3672: Yang-Mills theory. In addition to gauge-field loops there
3673: would be quark loops in the expansion of the thermodynamical potentials.
3674: Thermodynamical self-consistency would imply a system of two coupled
3675: evolution equations whose solutions would predict $e(T)$ and $\kappa(T)$ and can be used to compute
3676: the temperature dependence of the thermodynamical potentials. Presumably, the dynamical quark
3677: masses would become large close to the transition to the magnetic phase.
3678: This would mean that chiral
3679: symmetry is dynamically broken. As a consequence, quarks would decouple
3680: thermodynamically at the phase boundary and be replaced by (relatively strongly interacting)
3681: chiral Goldstone modes in the magnetic phase. The latter could overcompensate
3682: the negative pressure generated by the ground state of
3683: condensed magnetic monopoles. An extension of this approach to the case
3684: of finite quark chemical potential $\mu_q$ should be relatively straight forward.
3685: A more fundamental approach, were quarks arise as topological solitons in the
3686: center phases of various SU(3) Yang-Mills
3687: theories, would be much more difficult.
3688:
3689: Due to the dominance of the ground state in the magnetic phase a
3690: gauge theory for cosmological inflation based on SU(N) Yang-Mills
3691: thermodynamics would be a natural application. This would
3692: be a gauge-theory realization of warm inflation \cite{Berera}.
3693: Density perturbations generated in the magnetic phase would be dominated by
3694: thermal fluctuations. Along these lines an attempt to construct
3695: a gauge theory for warm inflation was made in \cite{HofmannKeil2002}.
3696:
3697: If the entire matter of the Universe would be described in terms of a `mother' SU(N) Yang-Mills
3698: theory, which break into SU(K) factors at $M_P$,
3699: then the energy-momentum tensor of the ground state would vanish
3700: identically for all those `daughter' theories that are in their center
3701: phases now. {\sl No cosmological constant is generated by the latter.}
3702:
3703:
3704: \section*{Acknowledgments}
3705: The author would like to thank B. Garbrecht, H. Gies, Th. Konstandin, T. Prokopec, H. Rothe, K. Rothe,
3706: M. Schmidt, I.-O. Stamatescu, and W. Wetzel for very helpful, continuing discussions.
3707: Important support for numerical calculations was provided by J. Rohrer and is
3708: thankfully acknowledged.
3709: Very useful discussions with P. van Baal, E. Gubankova, J. Moffat, J. Polonyi, D. Rischke, and F. Wilczek and
3710: illuminating conversations with D. B\"odeker, R. Brandenberger, G. Dunne, Ph. de Forcrand, A. Guth,
3711: F. Karsch, A. Kovner, M. Laine, H. Liu, C. Nunez, R. D. Pisarski, K. Rajagopal, K. Redlich, D. T. Son,
3712: A. Vainshtein, J. Verbaarschot, and F. Wilczek are gratefully acknowledged.
3713: The warm hospitality of the Center for
3714: Theoretical Physics at M.I.T, where part of this research was carried out
3715: (sponsored by Deutsche Forschungsgemeinschaft), is thankfully acknowledged.\\
3716: This paper is dedicated to my family.
3717:
3718:
3719: \bibliographystyle{prsty}
3720:
3721: \begin{thebibliography}{10}
3722:
3723: \bibitem{Higgs2000}
3724: G. Gomez-Ceballos {\sl et al.}, Int. J. Mod. Phys. A {\bf 16S1B}, 839 (2001).
3725:
3726: \bibitem{RHIC2003}
3727: U. W. Heinz and P. F. Kolb, Nucl. Phys. A {\bf 702}, 269 (2002).
3728:
3729: \bibitem{ShuryakTeaney2001}
3730: D. Teaney, J. Lauret, and E. V. Shuryak, nucl-th/0110037.
3731:
3732: \bibitem{Linde1982}
3733: A. D. Linde, Phys. Lett. B {\bf 108}, 389 (1982).
3734:
3735: \bibitem{Guth1982}
3736: A. H. Guth and S. Y. Pi, Phys. Rev. Lett. {\bf 49}, 1110 (1982).
3737:
3738: \bibitem{Starobinsky1982}
3739: A. A. Starobinsky, Phys. Lett. B {\bf 117}, 175 (1982).
3740:
3741: \bibitem{WMAP2003}
3742: D. N. Spergel {\sl et al.}, Astrophys. J. Suppl. {\bf 148}, 175 (2003).
3743:
3744: \bibitem{Dai2002}
3745: Z. G. Dai {\sl et al.}, Astrophys. J. {\bf 580}, L7 (2002).
3746:
3747: \bibitem{IMF}
3748: J. Bagchi {\sl et al.}, New Astron. {\bf 7}, 249 (2002).
3749:
3750: \bibitem{Hagedorn1965}
3751: R. Hagedorn, Nuovo Cim. Suppl. {\bf 3}, 147 (1965).
3752:
3753: \bibitem{GrossWilczek1973}
3754: D. J. Gross and F. Wilczek, Phys. Rev. Lett. {\bf 30}, 1343 (1973).\\
3755: D. J. Gross and F. Wilczek, Phys. Rev. D {\bf 8}, 3633 (1973).
3756:
3757: \bibitem{Politzer1973}
3758: H. D. Politzer, Phys. Rev. Lett. {\bf 30}, 1346 (1973).
3759:
3760: \bibitem{Linde1980}
3761: A. D. Linde, Phys. Lett. B {\bf 96}, 289 (1980).
3762:
3763: \bibitem{BraatenPisarski1990}
3764: J. C. Taylor and S. M. H. Wong, Nucl. Phys. B {\bf 346}, 115 (1990).\\
3765: E. Braaten and R. D. Pisarski, Nucl. Phys. B {\bf 337}, 569.
3766:
3767: \bibitem{Bodeker1998}
3768: D. B\"odeker, Phys. Lett. B {\bf 426}, 351 (1998).
3769:
3770: \bibitem{Bodapps1998}
3771: P. Arnold, D. T. Son, and L. G. Yaffe, Phys. Rev. D {\bf 59}, 105020 (1999).\\
3772: J.-P. Blaizot and E. Iancu, Nucl. Phys. B {\bf 557}, 183 (1999).\\
3773: D. F. Litim and C. Manuel, Phys. Rev. Lett. {\bf 82}, 4981 (1999).
3774:
3775: \bibitem{Kajantie2002}
3776: K. Kajantie, M. Laine, K. Rummukainen, and Y. Schroder , Phys. Rev. D {\bf 67}, 105008 (2003).\\
3777: K. Kajantie, M. Laine, K. Rummukainen, and Y. Schroder, Phys. Rev. Lett. {\bf 86}, 10 (2001).
3778:
3779: \bibitem{Blaizot2003}
3780: J.-P. Blaizot, E. Iancu, and A. Rebhan, hep-ph/0303185.
3781:
3782: \bibitem{Hofmann2003F}
3783: R. Hofmann, hep-ph/0312046.
3784:
3785: \bibitem{Hofmann2000t2003}
3786: R. Hofmann, Phys. Rev. D {\bf 62}, 065012 (2000).\\
3787: R. Hofmann, Phys. Rev. D {\bf 62}, 105021 (2000).\\
3788: R. Hofmann, Phys. Rev. D {\bf 65}, 125025 (2002).\\
3789: R. Hofmann, Phys. Rev. D {\bf 68}, 065015 (2003).
3790:
3791: \bibitem{GinzburgLandau1950}
3792: V. L. Ginzburg and L. D. Landau, JETP {\bf 20}, 1064 (1950).
3793:
3794: \bibitem{Abrikosov1957}
3795: A. A. Abrikosov, Sov. Phys. JETP {\bf 5}, 1174 (1957).
3796:
3797: \bibitem{Wetterich1996}
3798: C. Wetterich, Z. Phys.C {\bf 72}, 139 (1996).
3799:
3800: \bibitem{HarrigtonShepard1977}
3801: B. J. Harrington and H. K. Shepard, Phys. Rev. D {\bf 17}, 105007 (1978).
3802:
3803: \bibitem{PrasadSommerfield1975}
3804: M. K. Prasad and C. M. Sommerfield, Phys. Rev. Lett. {\bf 35 }, 760 (1975);
3805: E.B. Bogomolnyi, Sov. J. Nucl. Phys. {\bf 24}, 449 (1976).
3806:
3807: \bibitem{Nahm1984}
3808: W. Nahm, Lect. Notes in Physics. 201, eds. G. Denaro, e.a. (1084) p. 189.
3809:
3810: \bibitem{KraanVanBaalNPB1998}
3811: T. C. Kraan and P. van Baal, Nucl. Phys. B {\bf 533}, 627 (1998).
3812:
3813: \bibitem{vanBaalKraalPLB1998}
3814: T. C. Kraan and P. van Baal, Phys. Lett. B {\bf 435}, 389 (1998).
3815:
3816: \bibitem{Brower1998}
3817: R. C. Brower, D. Chen, J. Negele, K. Orginos, and C-I Tan, Nucl. Phys. Proc. Suppl. {\bf 73}, 557 (1999).
3818:
3819: \bibitem{GrossPisarskiYaffe1981}
3820: D. J. Gross, R. D. Pisarski, and L. G. Yaffe, Rev. Mod. Phys. {\bf 53}, 43 (1981).
3821:
3822: \bibitem{BPST}
3823: A. Belavin, A. Polyakov, A. Schwartz, and Yu. Tyupkin, Phys. Lett. {\bf 59}, 85 (1975).
3824:
3825: \bibitem{Diakonov}
3826: D. Diakonov, N. Gromov, V. Petrov, and S. Slizovskiy, hep-th/0404042.
3827:
3828: \bibitem{Kibble1976}
3829: T. W. B. Kibble, J. Phys. A {\bf 9}, 1387 (1976).
3830:
3831: \bibitem{HerbstHofmann2004}
3832: U. Herbst and R. Hofmann, hep-th/0411214.
3833:
3834: \bibitem{DiakonovPetrov2003}
3835: D. Diakonov and V. Petrov, Phys. Rev. D {\bf 67}, 105007 (2003).
3836:
3837: \bibitem{Hosotani1983}
3838: Y. Hosotani, Phys. Lett. B {\bf 126}, 309 (1983).
3839:
3840: \bibitem{HerbstHofmannRohrer2004}
3841: U. Herbst, R. Hofmann, and J. Rohrer, hep-th/0410178.
3842:
3843: \bibitem{Gorenstein1995}
3844: M. I. Gorenstein and S. N. Yang, Phys. Rev. D {\bf 52}, 5206 (1995).
3845:
3846: \bibitem{Biro1990}
3847: T. S. Biro, P. Levai and B. M\"uller, Phys. Rev. D {\bf 42}, 3078 (1990).
3848:
3849: \bibitem{Peshier1995}
3850: A. Peshier {\sl et al.}, Phys. Rev. D {\bf 54}, 2399 (1996).
3851:
3852: \bibitem{LevaiHeinz1998}
3853: P. Levai and U. Heinz, Phys. Rev. D {\bf 57}, 1879 (1998).
3854:
3855: \bibitem{KorthalsAltes}
3856: C. P. Korthals Altes, hep-ph/0406138,hep-ph/0408301.\\
3857: C. P. Korthals Altes, Acta Phys. Polon. B {\bf 34}, 5825 (2003).\\
3858: P. Giovannangeli and C. P. Korthals Altes, Nucl. Phys. B {\bf 608}, 203 (2001).\\
3859: C. Korthals-Altes, A. Kovner, and Misha A. Stephanov, Phys. Lett. B {\bf 469}, 205 (1999).
3860:
3861: \bibitem{HoelbingRebbiRubakov2001}
3862: Ch. Hoelbing, C. Rebbi, and V. A. Rubakov, Phys. Rev. D {\bf 63}, 034506 (2001).
3863:
3864: \bibitem{'tHooft1974}
3865: G. 't Hooft, Nucl. Phys. B {\bf 79}, 276 (1974).
3866:
3867: \bibitem{Polayakov1974}
3868: A. M. Polyakov, JETP Lett. {\bf 20}, 194 (1974).
3869:
3870: \bibitem{PrasadSommerfield1974}
3871: M. K. Prasad and C. M. Sommerfield, Phys. Rev. Lett. {\bf 35}, 760 (1975)
3872:
3873: \bibitem{'tHooft1976}
3874: G. 't Hooft, Phys. Rev. D {\bf 14}, 3432 (1976).
3875:
3876: \bibitem{LandsmanWeert1987}
3877: N. P Landsman and C. G. van Weert, Phys. Rep. {\bf 145}, 141 (1987)
3878:
3879: \bibitem{Hofmann2002}
3880: R. Hofmann, Phys. Rev. D {\bf 65}, 125025 (2002).
3881:
3882: \bibitem{ForgasVolkov2003}
3883: P. Forgas and M. Volkov, hep-th/0311062.
3884:
3885: \bibitem{Kaluza1921}
3886: T. Kaluza, Sitzungsber. Preuss. Akad. Wiss. Berlin (Math.Phys.) {\bf 1921}, 966 (1921).
3887:
3888: \bibitem{Klein1926}
3889: O. Klein, Z. Phys. {\bf 37}, 895 (1926).\\
3890: O. Klein, Surveys High Energ.Phys. {\bf 5}, 241 (1986).
3891:
3892: \bibitem{Olejnik1997}
3893: L. Del Debbio {\sl et al.}, proc.
3894: NATO Adv. Res. Workshop on Theor.
3895: Phys., Zakopane (1997) [hep-lat/9708023].
3896:
3897: \bibitem{NielsenOlesen1973}
3898: H. B. Nielsen and P. Olesen, Nucl. Phys. B {\bf 61}, 45 (1973).
3899:
3900: \bibitem{'tHooft1978}
3901: G. 't Hooft, Nucl. Phys. B {\bf 138}, 1 (1978).
3902:
3903: \bibitem{deForcrandEliaPepe2001}
3904: Ph. de Forcrand, M. D'Elia, and M. Pepe, Phys. Rev. Lett. {\bf 86}, 1438 (2001).
3905:
3906: \bibitem{SvetitskyYaffe1982-1}
3907: B. Svetitsky and L. G. Yaffe, Nucl. Phys. B {\bf 210}, 423 (1982).
3908:
3909: \bibitem{SvetitskyYaffe1982-2}
3910: B. Svetitsky and L. G. Yaffe, Phys. Rev. D {\bf 26}, 963 (1982).
3911:
3912: \bibitem{Hofmann20033}
3913: R. Hofmann, hep-ph/0312051.
3914:
3915: \bibitem{HongLiu2004}
3916: H. Liu, hep-th/0408001.
3917:
3918: \bibitem{nonequ}
3919: J.H. Traschen and R.H. Brandenberger, Phys. Rev. D {\bf 42}, 2491 (1990).\\
3920: L. Kofman, A.D. Linde and A.A. Starobinsky, Phys. Rev. Lett. {\bf 73}, 3195 (1994).\\
3921: J. Berges and J. Serreau, Phys. Rev. Lett. {\bf 91}, 111601 (2003).
3922:
3923: \bibitem{EnKaSaMo1982}
3924: J. Engels {\sl et} al., Nucl. Phys. B {\bf 205}[FS5], 545 (1982).
3925:
3926: \bibitem{Bielfeld1996}
3927: G. Boyd {\sl et} al., Phys. Rev. Lett. {\bf 75}, 4169 (1995).\\
3928: G. Boyd {\sl et} al., Nucl. Phys. {\bf B469}, 419 (1996).
3929:
3930: \bibitem{EngelsKarschScheideler1999}
3931: J. Engels, F. Karsch, T. Scheideler, Nucl. Phys. B {\bf 564}, 302 (1999).
3932:
3933: \bibitem{Brown1988}
3934: F. R. Brown {\sl et al.}, Phys. Rev. Lett. {\bf 61}, 2058 (1988).
3935:
3936: \bibitem{Deng1988}
3937: Y. Deng, in BATAVIA 1988, proc. LATTICE 88, 334.\\
3938: J. Engels {\sl et al.}, Phys. Lett. B {\bf 252}, 625 (1990).
3939:
3940: \bibitem{EngelsFingberg1990}
3941: J. Engels {\sl et al.}, Phys. Lett. B {\bf 252}, 625 (1990).
3942:
3943: \bibitem{LuciniTeperWenger2003}
3944: B. Lucini, M. Teper, and U. Wenger, JHEP {\bf 0401}, 061 (2004).
3945:
3946: \bibitem{BenderWu1976}
3947: C. M. Bender and T. T. Wu, Phys. Rev. D {\bf 7}, 1620 (1973).
3948:
3949: \bibitem{Klassen1998}
3950: T. R. Klassen, Nucl. Phys. B {\bf 533}, 557 (1998).
3951:
3952: \bibitem{Hofmann20032}
3953: R. Hofmann, hep-ph/0312048.
3954:
3955: \bibitem{QEDtests}
3956: A. Ashkin, L. A. Page, and W. M. Woodard, Phys. Rev. {\bf 94}, 357 (1953).\\
3957: W. C. Barber, G. E. Becker, and E. L. Chu, Phys. Rev. {\bf 89}, 950 (1953).\\
3958: M. B. Scott, A. O. Hanson, and E. M. Lyman, Phys. Rev. {\bf 84}, 638 (1951).\\
3959: W. C. Barber {\sl et al.}, Phys. Rev. Lett. {\bf 16}, 1127 (1966). \\
3960: W. C. Barber {\sl et al.}, Phys. Rev. {\bf D3}, 2796 (1971). \\
3961: R. B. Blumenthal {\sl et al.}, Phys. Rev. {144}, 1199 (1966).\\
3962: H. Alvensleben {\sl et al.}, Phys. Rev. Lett. {\bf 21}, 1501 (1968).\\
3963: J. G. Asbury {\sl et al.}, Phys. Rev. Lett. {\bf 18}, 65 (1967).\\
3964: J. E. Augustin {\sl et al.}, Phys. Lett. {\bf B31}, 673 (1970).\\
3965: V. E. Balakin {\sl et al.}, Phys. Lett. {\bf B34}, 99 (1971). \\
3966: V. E. Balakin {\sl et al.}, Phys. Lett. {\bf B37}, 435 (1971).\\
3967: C. Bernadini, in proc. 1971 intern.
3968: symposium on electron and photon interactions at high energies, Cornell
3969: University, Ithaca, N. Y., Aug. 23-27 1971.\\
3970: P. Dittmann and V. Hepp, Z. Phys. {\bf C10}, 283 (1981), Erratum-ibid.{\bf C11}, 271 (1981).
3971:
3972: \bibitem{Wilczek2004}
3973: F. Wilczek, hep-ph/0408167.
3974:
3975: \bibitem{AlexanderMoffat2004}
3976: S. Alexander, M. Mbonye, and J. Moffat, hep-th/0406202.
3977:
3978: \bibitem{Reinhardt}
3979: H. Reinhardt {\sl et al.}, Phys. Rev. D {\bf 66}, 085004 (2002).\\
3980: H. Reinhardt, Nucl. Phys. Proc. Suppl. {\bf 119}, 658 (2003).\\
3981: H. Reinhardt and T. Tok, Phys. Rev. D {\bf 68}, 065004 (2003).
3982:
3983: \bibitem{Berera}
3984: A. Berera, Phys. Rev. Lett. {\bf 75}, 3218 (1995).\\
3985: A. Berera, Phys. Rev. D {\bf 54}, 2519 (1996).
3986:
3987: \bibitem{HofmannKeil2002}
3988: R. Hofmann and M. T. Keil, Phys. Rev. D {\bf 66}, 023527 (2002).\\
3989: R. Hofmann and M. T. Keil, hep-ph/0204203, published in *Cape Town 2002,
3990: Dark matter in astro- and particle physics* 296-305.
3991:
3992: \baselineskip25pt
3993:
3994: \end{thebibliography}
3995:
3996:
3997:
3998:
3999:
4000:
4001:
4002: \end{document}
4003: