1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %% Trim Size: 9.75in x 6.5in
3: %% Text Area: 8in (include Runningheads) x 5in
4: %% ws-ijmpa.tex : 25-10-04
5: %% Tex file to use with ws-ijmpa.cls written in Latex2E.
6: %% The content, structure, format and layout of this style file is the
7: %% property of World Scientific Publishing Co. Pte. Ltd.
8: %% Copyright 1995, 2002 by World Scientific Publishing Co.
9: %% All rights are reserved.
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: %%
12:
13: %\documentclass[draft]{ws-ijmpa}
14: \documentclass{ws-ijmpa}
15: %\documentclass{report}
16:
17: \usepackage{bbm}
18:
19: \newcommand{\UM}{\mathbbm 1}
20: \newcommand{\R}{\mathbbm R}
21: \newcommand{\C}{\mathbbm C}
22: \newcommand{\Z}{\mathbbm Z}
23: \newcommand{\Pl}{\mathbbm P}
24: \newcommand{\eqb}{\begin{equation}}
25: \newcommand{\eqe}{\end{equation}}
26: \newcommand{\dmb}{\begin{displaymath}}
27: \newcommand{\dme}{\end{displaymath}}
28: \newcommand{\pd}{\partial}
29: \newcommand{\ep}{\varepsilon}
30: \newcommand{\eab}{\begin{eqnarray}}
31: \newcommand{\eae}{\end{eqnarray}}
32: \newcommand{\ra}{\right\rangle}
33: \newcommand{\la}{\left\langle}
34: \newcommand{\e}{\mbox{e}}
35: \newcommand{\be}{\begin{equation}}
36: \newcommand{\ee}{\end{equation}}
37: \newcommand{\sgn}{\text{sgn}\,}
38: \newcommand{\munu}{{\mu\nu}}
39: \newcommand{\ad}{{\dot{\alpha}}}
40: \newcommand{\bd}{{\dot{\beta}}}
41: \newcommand{\La}{\Lambda}
42:
43: \begin{document}
44:
45: \markboth{Ralf Hofmann}{Nonperturbative approach to
46: Yang-Mills thermodynamics}
47:
48: %%%%%%%%%%%%%%%%%%%%% Publisher's Area please ignore %%%%%%%%%%%%%%%
49: %
50: \catchline{}{}{}{}{}
51: %
52: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53:
54: \title{Nonperturbative approach to
55: Yang-Mills thermodynamics}
56:
57: \author{Ralf Hofmann}
58:
59: \address{Institut f\"ur Theoretische Physik\\
60: Johann Wolfgang Goethe -- Universit\"at\\
61: Max von Laue -- Str. 1\\
62: D-60438 Frankfurt am Main\\
63: Germany\\
64: r.hofmann@thphys.uni-heidelberg.de}
65:
66:
67:
68: \maketitle
69:
70: \begin{history}
71: \received{7 April 2005}
72: \revised{26 November 2006}
73: \end{history}
74:
75: \begin{abstract}
76: An analytical and nonperturbative approach to SU(2) and SU(3) Yang-Mills thermodynamics is
77: developed and applied. Each theory comes in
78: three phases: A deconfining, a preconfining, and a
79: confining one. We show how macroscopic and inert scalar fields emerge in each
80: phase and how they determine the ground-state physics and
81: the properties of the excitations. While the excitations in the deconfining and preconfining phase
82: are massless or massive gauge modes of spin 1 the excitations in the confining phase
83: are massless or massive spin-1/2 fermions. The nature of the two phase transitions
84: is investigated for each theory. We compute the temperature evolution of
85: thermodynamical quantities in the deconfining and
86: preconfining phase, and we estimate the density of
87: states in the confining phase. Some implications
88: for particle physics and cosmology are discussed.
89:
90:
91: \keywords{thermal gauge theories; holonomy; caloron; magnetic monopole; center-vortex loop; thermal ground state;
92: thermal quasiparticles; Bose condensation; preconfinement; complete confinement;
93: Polyakov-loop expectation; Hagedorn transition; Planck-scale axion; cosmic coincidence;
94: CP violation; electroweak symmetry breaking; fractional Quantum Hall effect}
95: \end{abstract}
96:
97: \ccode{PACS numbers: 11.15.Ex, 11.10.Wx, 11.15.Tk}
98:
99: \section{Introduction\label{intro}}
100:
101: The beauty, richness and usefulness of nonabelian gauge theories is
102: generally appreciated. Yet, in a perturbative approach to gauge theories
103: like the Standard Model of particle physics (SM) and its (non)supersymmetric extensions it is hard if not impossible
104: to convincingly address a number of recent and not so recent experimental and
105: observational results in particle physics and
106: cosmology: Nondetection of the Higgs particle at LEP \cite{Higgs2000},
107: indications for a rapid thermalization and strongly
108: collective behavior of the plasma that emerges in the early stage of an ultra-relativistic heavy-ion collision
109: \cite{RHIC2003,ShuryakTeaney2001} despite the fact that the ideal hydrodynamical expansion essentially obeys
110: a free-gas equation of state, dark energy and dark matter, a
111: strongly favored epoch of cosmological inflation \cite{Cobe,WMAP2003I,Perlmutter1998,Schmidt1998}
112: in the very early Universe \cite{Linde1982,Guth1982,Dymnikova} and today's
113: accelerated cosmological expansion, a
114: puzzling large-angle signal in the power spectrum of the
115: cosmic microwave background for the cross correlation of
116: electric-field polarization and temperature fluctuations \cite{WMAPPol},
117: the likely existence of intergalactic magnetic
118: fields of so far unclarified origin \cite{Dai2002,IMF},
119: and the departure of $\sim 3\times 10^{43}$ protons p. a. (and hardly any negative charges) \cite{Manuel2004}
120: from the sun's surface (solar wind) which clearly is contradicting the
121: charge conservation inherent in the SM. An analytical and
122: nonperturbative approach to strongly interacting gauge
123: theories may further our understanding of these phenomena.
124:
125: The objective of the present work
126: is the thermodynamics of SU(2) and SU(3) Yang-Mills theories
127: in four dimensions. It is difficult to gain
128: insights in the dynamics of a strongly interacting
129: four-dimensional gauge theory by analytical means if this dynamics is not severely
130: constrained by certain global symmetries.
131: We conjecture with Ref.\,\cite{Hagedorn1965} that a thermodynamical approach
132: is an appropriate starting point for such an endeavor.
133: On the one hand, this conjecture is reasonable since a strongly interacting
134: system, being in equilibrium, communicates distortions almost instantaneously by
135: rigid correlations, and thus a return to equilibrium takes place very rapidly.
136: On the other hand, it turns out that
137: the requirement of thermalization allows for an analytical and
138: nonperturbative derivation of macroscopic ground states and the properties of their (quasiparticle)
139: excitations in two of the three phases of each theory. A breakdown of equilibrium in a
140: transition to the third phase is unproblematic since the dynamics
141: then is uniquely determined by the remaining symmetry.
142:
143: Let us very briefly recall some aspects of the analytical
144: approaches to thermal SU($N$) Yang-Mills theory as they are
145: discussed in the literature. Because of asymptotic
146: freedom \cite{GrossWilczek1973,Politzer1973} one would naively
147: expect thermal perturbation theory to work well for temperatures
148: $T$ much larger than the Yang-Mills scale $\La$ since the gauge coupling
149: constant $\bar{g}(T)$ logarithmically approaches zero
150: for $\frac{T}{\La}\to\infty$.
151: It is known for a long time that this expectation is too optimistic since at
152: any temperature perturbation theory
153: is plagued by instabilities arising from the infrared sector (weakly screened,
154: soft magnetic modes \cite{Linde1980}). As a consequence, the pressure $P$ can be computed
155: perturbatively only up
156: to (and including) order $\bar{g}^5$. The effects of resummations of
157: one-loop diagrams (hard thermal loops), which rely on a
158: scale separation furnished by the small value of
159: the coupling constant $\bar{g}$, are summarized in terms of a
160: nonlocal effective theory for soft and
161: semi-hard modes \cite{BraatenPisarski1990}. In the computation of radiative corrections,
162: based on this effective theory, infrared effects due to soft magnetic modes still appear in an
163: uncontrolled manner. This has
164: led to the construction of an effective theory where soft modes are collectively described in
165: terms of classical fields whose dynamics is influenced by integrated
166: semi-hard and hard modes \cite{Bodeker1998,Bodapps1998}.
167: In Quantum Chromodynamics (QCD) a perturbative calculation of $P$ was pushed up to order
168: $\bar{g}^6\log\,\bar{g}$, and an additive `nonperturbative'
169: term at this order was fitted to lattice results \cite{Kajantie2002}. Within the
170: perturbative orders a poor convergence of the expansion is observed for
171: temperatures not much larger than the $\overline{\mbox{MS}}$ scale. While
172: the work in \cite{Kajantie2002} is a computational masterpiece it could, by definition,
173: not shed light on the missing, nonperturbative
174: dynamics of the infrared sector. Screened perturbation theory, which relies
175: on a split of the tree-level Yang-Mills action by the introduction of
176: variational parameters, is a very interesting idea. Unfortunately, this approach generates
177: temperature dependent ultraviolet divergences in its presently used
178: form, see \cite{Blaizot2003} for a recent review.
179:
180: The purpose of the present work is to report, in a detailed way, on the development of
181: a nonperturbative and analytical approach to the
182: thermodynamics of SU(2) and SU(3) Yang-Mills theory
183: (see \cite{Hofmann2000t2003} for intermediate stages). The reason why we consider only these
184: two gauge groups is that for SU($N$) with $N\ge 4$ it is likely that the
185: phase structure of the theory is not unique: In contrast to SU(2) and SU(3),
186: which possess one confining (center), one
187: preconfining (magnetic), and one deconfining (electric) phase, more
188: than three phases may exist for an SU($N$) Yang-Mills theory with $N\ge 4$.
189:
190: Our starting point is the derivation of a macroscopic ground state in
191: the deconfining phase. This ground state originates
192: from instantaneous, long-range correlations mediated by field configurations of
193: topological charge $\pm 1$. Technically, this situation
194: is described by a spatially homogeneous, quantum mechanically
195: and statistically inert scalar field $\phi$, which transforms
196: under the adjoint representation of the gauge group, and a
197: pure-gauge configuration of trivial topology solving
198: the Yang-Mills equations subject to a source term provided
199: by $\phi$. Both the field $\phi$ and the
200: pure-gauge configuration emerge after a spatial
201: coarse-graining over quantum fluctuations
202: down to a resolution corresponding to
203: the length scale $|\phi|^{-1}$.
204: While $\phi$ represents the spatial average over BPS saturated,
205: topological defects, that is, `large' quantum fluctuations the
206: pure-gauge configuration is a manifestation of averaged-over
207: plane-wave quantum fluctuations.
208:
209: Conceptually, our approach is similar to the
210: macroscopic Landau-Ginzburg-Abrikosov theory for
211: superconductivity in metals \cite{GinzburgLandau1950,Abrikosov1957}.
212: Recall that this theory describes
213: the existence of a condensate of Cooper
214: pairs in terms of a nonvanishing expectation for a
215: complex scalar field $\varphi$ (local order parameter),
216: which is charged under the electromagnetic gauge group U(1), and in terms of a
217: pure-gauge configuration. A nonzero value of $\varphi$ is enforced by a
218: phenomenologically introduced potential $V$.
219: As a consequence, coarse-grained U(1) gauge-field modes $\delta a_\mu$ (photons),
220: which are deprived of the
221: microscopic gauge-field fluctuations contributing to
222: the formation of Cooper pairs and their subsequent condensation,
223: acquire a mass. Microscopically, the generation of a photon mass can be
224: visualized as a large sequence of elastic scattering processes off the
225: electrons residing within individual Cooper pairs in the condensate. At a
226: given photon momentum this slows down the effective velocity of propagation in comparison
227: to a propagation without a Cooper-pairs condensate, see Fig.\,\ref{Fig0}.
228: %***********************
229: \begin{figure}
230: \begin{center}
231: \leavevmode
232: %\epsfxsize=9.cm
233: \leavevmode
234: %\epsffile[80 25 534 344]{}
235: \vspace{4.3cm}
236: \special{psfile=Fig-1.ps angle=0 voffset=-140
237: hoffset=-105 hscale=55 vscale=75}
238: \end{center}
239: \caption{\protect{Photon propagation without (a) and within (b) a
240: Cooper-pair condensate.\label{Fig0}}}
241: \end{figure}
242: %************************
243: In the superconducting phase the U(1) gauge symmetry is
244: spontaneously broken by the Cooper-pair condensate,
245: and physical phenomena associated with this breakdown
246: can be analysed in dependence of the parameters entering an
247: effective action and in dependence of external conditions such as a
248: magnetic field and/or temperature.
249:
250: When pursuing the idea of a dynamically generated, macroscopic
251: ground state in each of the phases of an SU(2) or SU(3) Yang-Mills
252: theory it turns out that a situation similar to superconductivity holds. Moreover, in a
253: thermalized SU(2) or SU(3) Yang-Mills
254: theory one is in the comfortable position of being able to {\sl derive} the dynamics
255: of macroscopic scalar fields from first principles. Thus the
256: stabilizing potentials for each scalar field are uniquely determined (up to an undetermined mass parameter --
257: the scale of the Yang-Mills theory). Each (gauge invariant)
258: potential is specified by a unique microscopic
259: definition of the scalar field's phase (in a suitably chosen gauge)
260: and by the assumption that a dynamically generated, constant mass scale exists.
261: This assumption is supported by
262: perturbation theory \cite{GrossWilczek1973,Politzer1973}
263: where the specification of the running of the
264: gauge coupling requires such a boundary condition. The microscopic
265: definition of the scalar field's phase is an average
266: over an (nonlocal) operator saturated by noninteracting Bogomolnyi-Prasad-Sommerfield (BPS)\cite{PrasadSommerfield1975}
267: saturated configurations of
268: topological charge $\pm 1$. If a particular phase supports
269: propagating gauge modes then, in a second step,
270: interactions between the topological defects and microscopic
271: radiative corrections are taken into account macroscopically by pure-gauge
272: solutions to the equations of motion for gauge fields residing in the
273: trivial-topology sector. The source term for these equations of motion is provided by
274: the (inert) scalar fields.
275:
276: More specifically, we have shown in \cite{HerbstHofmann2004}
277: that at large temperatures upon spatial coarse-graining
278: an adjoint scalar field $\phi$ emerges due to the spatial correlations mediated by
279: trivial-holonomy calorons\cite{HarrigtonShepard1977}. (By large temperature
280: we mean large as compared to the dynamically generated scale.)
281: We discuss in Sec.\,\ref{eveffgc} why the critical
282: temperature $T_P$ for the onset of $\phi$'s existence should be comparable to
283: the cutoff-scale for a field theoretic description in four dimensions.
284: Trivial-holonomy calorons are BPS saturated (or selfdual) solutions \cite{PrasadSommerfield1975}
285: to the Euclidean Yang-Mills equations at finite temperature. (The time coordinate $\tau$ is
286: compactified on a circle, $0\le\tau\le\frac{1}{T}$.) The topological charge of these
287: configurations is integer. (Whenever we speak of a
288: topological soliton in this section its antisoliton is also meant.)
289: It will turn out that only calorons with topological charge
290: one contribute to the moduli-space average defining $\phi$'s phase. The reason is that
291: for higher-charge calorons the larger number of dimensionful moduli does not admit the
292: definition of a dimensionless entity without the introduction of
293: an explicit temperature dependence \cite{HerbstHofmann2004}. The latter, however, ought
294: to be absent because of a temperature independent weight on the
295: classical level. To understand macroscopic results generated by microscopic
296: interactions between calorons and between their constituents an investigation
297: of the properties of quantum corrected nontrivial-holonomy calorons
298: is necessary \cite{Diakonov2004}.
299:
300: For a given SU($N$) gauge-field configuration nontrivial holonomy refers to the
301: following property: Evaluate the Polyakov-loop on this configuration
302: at spatial infinity and observe that the result is not an element
303: %*********
304: \eqb
305: \label{noho}
306: \UM\,\exp\left[\frac{2\pi i}{N}k\right]\,\ \ \ (k=0,\cdots,N-1)
307: \eqe
308: %*********
309: of the center $Z_N$ of the gauge group. If this is the case then
310: a mass scale exists in the configuration which determines the behavior
311: $A_4(|\vec{x}|\to\infty)$. For a classical solution at finite temperature this
312: mass scale must be temperature itself. The quantity $u\equiv T\int_0^\beta d\tau\,A_4(\tau,|\vec{x}|\to\infty)$ defines
313: the holonomy of the configuration ($\beta\equiv\frac{1}{T}$).
314: Due to a reflection symmetry $u\to -u$ one can restrict the values of $u$ as
315: $0\le u\le 2\pi T$ for SU(2).
316:
317: Calorons with nontrivial holonomy are selfdual configurations that possess BPS magnetic monopole constituents
318: \cite{Nahm1984,KraanVanBaalNPB1998,vanBaalKraalPLB1998,LeeLu1998,Brower1998}:
319: For an SU($N$) caloron with no net magnetic charge
320: there are $N$ constituent monopoles whose magnetic
321: charges add up to zero. The masses of the monopoles are
322: determined by the holonomy $u$ and thus are $\propto T$.
323: By a recent heroic calculation the one-loop quantum weight for an SU(2) caloron
324: with nontrivial holonomy was derived in \cite{Diakonov2004}.
325:
326: Since the one-loop effective action of a nontrivial-holonomy caloron scales
327: with the spatial volume of the
328: system one is tempted to conclude that these configurations do not contribute
329: to the partition function of the theory in the
330: thermodynamical limit and thus are irrelevant
331: \cite{GrossPisarskiYaffe1981}. This conclusion,
332: however, is not valid since one can show that
333: nontrivial-holonomy calorons are {\sl instable} under
334: quantum corrections\cite{Diakonov2004}. Moreover, if, on spatial average, interacting calorons are shown to be described by a
335: quantum mechanically and statistically inert,
336: {\sl macroscopic} adjoint scalar field $\phi$ and
337: a pure-gauge configuration $a_\mu^{bg}$ \cite{HerbstHofmann2004}
338: then the exponent of minus the associated effective action $S_{cl}$ can be factored out in the
339: partition function and thus cancels in any physical average.
340: This situation holds even if $S_{cl}$ scales with the spatial
341: volume of the system and thus is infinite in the thermodynamical limit. On the microscopic level, the generation
342: of (instable) nontrivial holonomy is due to interactions between
343: trivial-holonomy calorons mediated by long-wavelength fields that
344: reside in the topologically trivial sector of the theory.
345:
346: For the SU(2) case it was shown in \cite{Diakonov2004} that the one-loop fluctuations around
347: calorons with a small holonomy generate an
348: {\sl attractive} potential between the two BPS monopole
349: constituents. This implies that monopole and antimonopole
350: annihilate shortly after they have been created.
351: Thus the likelihood of such a process roughly is determined by the finite
352: one-loop quantum weight of a trivial-holonomy caloron. Up to an additive correction, which depends on
353: temperature and the caloron radius $\rho$ and which is finite,
354: the effective action equals the classical action $S=\frac{8\pi^2}{g^2}$.
355: For $g$ not too small the likelihood of generating a
356: small holonomy is sizable. In the opposite case
357: of a large holonomy a {\sl repulsive} potential between the
358: monopole constituents arises due to one-loop fluctuation
359: \cite{Diakonov2004}. Thus monopole and antimonopole
360: separate back-to-back, and the caloron dissociates into a pair of an isolated monopole and an isolated
361: antimonopole which are screened by intermediate caloron fluctuations
362: \cite{KorthalsAltes,HoelbingRebbiRubakov2001}. Before
363: screening, that is, on the level of the
364: classical solution the mass of both monopole and
365: antimonopole is much larger than
366: temperature \cite{LeeLu1998}. We conclude that the generation and subsequent
367: dissociation of a large-holonomy caloron is a very
368: rare process due to an extreme Boltzmann suppression. Thus attraction between a monopole and its
369: antimonopole is the dominating situation in the ground-state physics of an SU(2) or SU(3) Yang-Mills
370: theory being in the electric phase.
371: The macroscopic manifestation of monopole-antimonopole attraction, their subsequent
372: annhihilation, and recreation is a {\sl negative}
373: ground-state pressure. Equating the exponent in
374: the Boltzmann distribution of the monopole-antimonopole system before screening
375: with the exponent in the one-loop quantum weight of the caloron allows for an estimate of the typical distance
376: between a monopole and an antimonopole at
377: a given temperature. We will see that, on the scale of
378: the inverse temperature, isolated and screened
379: monopoles are very dilute. The case of SU(2) has a
380: straight-forward generalization to SU(3): No qualitative changes
381: take place in the above discussion when going from SU(2) to SU(3).
382:
383: From the selfduality or BPS saturation of the caloron it follows
384: that its energy-momentum tensor vanishes identically. Since the
385: macroscopic field $\phi$ originates from the spatial correlations of noninteracting
386: trivial-holonomy calorons of topological charge
387: one \cite{HerbstHofmann2004} $\phi$'s
388: macroscopic energy-momentum tensor vanishes in the absence of a
389: coupling to the topologically trivial sector
390: of the theory. This is a derived condition which needs to be
391: imposed to fix some of the ambiguities which emerge when calculating $\phi$'s phase. Namely, one insists
392: on a BPS saturation of the $\tau$ dependence
393: of this phase: A linear
394: second-order equation of motion for $\phi$'s phase can be derived
395: from a microscopic definition involving a moduli-space average over a
396: two-point correlator, and the requirement of BPS saturation determines the
397: solution up to an irrelevant global gauge choice and an
398: irrelevant constant phase shift \cite{HerbstHofmann2004}.
399:
400: Subsequently, a potential
401: $V_E$ for the canonically normalized field $\phi$ is derived by appealing to the derived
402: information on $\phi$'s phase and to the assumptions
403: that a dynamically generated scale $\La_E$ exists and that
404: the right-hand side of $\phi$'s BPS equation
405: is analytic in $\phi$. (Here the
406: subscript $E$ refers to the electric phase.)
407: As a consequence, $\phi$'s modulus is $\propto T^{-1/2}$ and $V_E\propto T$.
408: Thus the caloron sector contributes to thermal quantities
409: in a power-suppressed way at large temperatures. Moreover,
410: we will show that the ground state in the electric phase is degenerate with
411: respect to a global electric $Z_2$ (SU(2)) and a global electric $Z_3$ (SU(3))
412: symmetry. Therefore, the electric phase is {\sl deconfining}.
413:
414: In the coarse-grained theory a useful decomposition of
415: field configurations $a_\mu$ with trivial topology (only those
416: ones appear as explicit gauge fields) is
417: %***********
418: \eqb
419: \label{fludec}
420: a_\mu=a_\mu^{bg}+\delta a_\mu\,
421: \eqe
422: %***********
423: where $a_\mu^{bg}$ denotes a pure-gauge configuration
424: belonging to the ground state, and $\delta a_\mu$
425: is a finite-curvature fluctuation. In unitary gauge, where $a_\mu^{bg}=0$ and
426: thus no coupling between $\delta a_\mu$ and $a_\mu^{bg}$
427: exists, a fluctuation $\delta a_\mu$ acquires a mass by the adjoint Higgs mechanism
428: if $[\phi,\delta a_\mu]\not=0$. Since the field $\phi$ dynamically breaks
429: the gauge symmetries SU(2)$\to$U(1)and SU(3)$\to$U(1)$^2$, two and six directions in the three and
430: eight dimensional adjoint color space acquire mass, respectively.
431:
432: We shall discuss
433: why the scale $\La_E$, which measures the strength
434: of apparent gauge-symmetry breaking by calorons at a given temperature, is physically
435: set at a temperature scale $T_P$ where any four-dimensional
436: gauge theory fails to describe reality. Common belief is that $T_P$ is comparable to
437: the Planck mass $M_P\sim 10^{19}\,$GeV.
438:
439: How does the existence of the temperature dependent
440: scale $|\phi|$ influence the
441: propagation of gauge modes in the
442: infrared and ultraviolet? Calorons induce
443: quasiparticle masses on tree level
444: in the effective theory which are
445: of the order $e|\phi|$. Here $e$ denotes
446: the {\sl effective} gauge coupling. This coupling measures the
447: interaction strength between the topologically trivial
448: off-Cartan fluctuations (in unitary gauge) and the coarse-grained
449: manifestation of nontrivial topology. Furthermore, it is a measure for
450: the screening of the magnetic charge of a monopole. (As far as thermodynamical quantities are concerned, essentially
451: all excitations are free (quasi)particles \cite{HerbstHofmannRohrer2004}.).
452: The evolution of $e$ with temperature follows from the
453: requirement of thermodynamical selfconsistency
454: of this interaction. Except for a small range in temperature
455: to the right of the electric-magnetic transition
456: at $T_{c,E}$, where
457: %********
458: \eqb
459: \label{coupldiv}
460: e(T)\sim -\log(T-T_{c,E})\,,
461: \eqe
462: %********
463: the coupling $e$ is constant: A manifestation of the existence
464: of screened\footnote{The screening of the monopoles is due to surrounding
465: caloron fluctuations of small holonomy and not due to Cartan {\sl excitations}
466: since the latter are not capable of screening
467: static magnetic fields \cite{Linde1980}.}, isolated and conserved magnetic charges generated
468: by dissociating large-holonomy calorons.
469:
470: Infrared cutoffs $\sim e|\phi|$ arise
471: from a reduction of propagation speed for off-Cartan fluctuations by their
472: interactions with calorons, compare with the analogous situation for a superconducting material in
473: Fig.\,\ref{Fig0}. Due to the existence of these cutoffs in the loop expansions of
474: thermodynamical quantities the problem of a
475: magnetic instability, as encountered in perturbation theory \cite{Linde1980}, is resolved
476: \cite{Linde1980,HerbstHofmannRohrer2004}.
477: In the ultraviolet, $|\phi|$ acts as a compositeness constraint by setting a maximal
478: scale for the off-shellness of quantum fluctuations. This is a
479: consistent requirement since {\sl all} gauge modes, on-shell or off-shell,
480: originate from the nontrivial ground state and thus ought
481: not be capable of destroying it. Moreover, plane-wave quantum fluctuations of an off-shellness
482: larger than $|\phi|^2$ are contained, in a coarse-grained form,
483: in the pure-gauge configuration $a_\mu^{bg}$. Compositeness constraints are
484: implemented in a physical gauge with respect to the unbroken
485: subgroups U(1) or U(1)$^2$. The usual renormalization
486: program, needed to make sense of ultraviolet
487: divergences in thermal perturbation theory, is
488: abandoned in the effective theory: The ground state itself provides
489: for a {\sl physical} ultraviolet cutoff.
490:
491: As a consequence of the simultaneous existence of both an
492: ultraviolet and an infrared cutoff the contributions of higher loops in the
493: expansion of thermodynamical quantities are very small.
494: (Technically, an evaluation of two-loop corrections to the pressure is
495: quite involved \cite{HerbstHofmannRohrer2004}.) Obviously, the situation outlined so far differs
496: substantially from the idea of a Wilsonian flow for
497: nonabelian gauge theories in its usual
498: implementation: One derives effective dynamics in dependence
499: of an externally set scale\footnote{This scale either is continuous,
500: see \cite{LitimPawlowski1999} for a review on gauge theories, or it reflects a scale
501: hierarchy originating from the assumed smallness of the
502: coupling constant $\bar{g}$ at a large temperature $T$, for example $k=\bar{g}T$
503: \cite{BraatenPisarski1990}.} $k$
504: by integrating plane-wave modes harder than $k$ into couplings
505: that appear in an ansatz for an effective action. This effective
506: action describes the dynamics of gauge
507: modes of maximal hardness $k$. The dynamics of theses modes is,
508: however, not only influenced by integrated-out
509: high-momentum fluctuations but also by (spatially) small-scale
510: fluctuations of nontrivial topology. Recall that the later can
511: not be expanded in terms of the former because of an
512: essential singularity in their weight at a vanishing value of
513: the fundamental gauge coupling. Thus we propose an approach which is the
514: converse of the usual picture: Integrate the
515: topological sector first and determine subsequently, that is, after spatial coarse-graining,
516: what its average effect on the trivial-topology fluctuations is.
517:
518: We have already mentioned that this approach to SU(2) or SU(3) Yang-Mills thermodynamics
519: implies the existence of three rather than two phases.
520: In the magnetic phase, where the isolated and
521: screened magnetic monopoles of the electric phase are massless and thus
522: condensed and where off-Cartan modes are thermodynamically
523: decoupled, the dual gauge symmetries U(1)$_D$ or U(1)$^2_D$ are broken
524: dynamically, and the global electric center
525: symmetries $Z_2$ or $Z_3$ are restored in the ground state. We will show that
526: the ground state in the magnetic phase is a Bose condensate
527: of monopole-antimonopole systems. Each condensate is described in terms of a macroscopic and inert
528: complex scalar field and a pure-gauge configuration.
529:
530: A monopole condensate confines fundamentally
531: charged, fermionic and static test charges. At the same
532: time, dual and massive gauge modes {\sl propagate} in the magnetic phase.
533: Thus it is appropriate to refer to the magnetic phase as a preconfining phase.
534: The magnetic coupling $g$,
535: which measures the interaction strength between dual gauge modes
536: and condensed magnetic monopoles on the one hand and the screening
537: of center-vortex loops on the other hand, is zero at $T_{c,E}$ and
538: rises rapidly into a logarithmic divergence of the same
539: form as in Eq.\,(\ref{coupldiv}) at a
540: temperature $T_{c,M}$. For SU(2) and SU(3) we
541: have $T_{c,M}=0.835\times\,T_{c,E}$ and
542: $T_{c,M}=0.877\times\,T_{c,E}$, respectively. Thus the magnetic
543: phase occupies only a small region in the phase diagram of either theory. This and the fact
544: that the monopole condensates possess infinite correlation
545: lengths $\sim \mbox{(monopole mass)}^{-1}$ are the reasons
546: why the magnetic phase has escaped its direct
547: detection by simulations on finite-size lattices. We will
548: show though that it is possible to observe the existence of the
549: magnetic phase in a lattice simulations of the infrared
550: insensitive entropy density when using the
551: so-called differential method. The latter provides the
552: best-controlled circumvention of the infrared problem in simulations on finite-size
553: lattices \cite{Brown1988}.
554:
555: At $T_{c,M}$, where $g$ diverges logarithmically,
556: another phase transition takes place. All dual gauge modes
557: decouple and thus the monopole condensates, macroscopically described by
558: nonfluctuating, BPS saturated complex scalar fields $\varphi$ (SU(2)) and
559: $\varphi_1,\varphi_2$ (SU(3)) and pure gauges, dominate the
560: thermodynamics. As a consequence, the
561: entropy density vanishes at $T_{c,M}$ and the
562: equation of state is
563: %********
564: \eqb
565: \label{eosintro}
566: \rho=-P\,.
567: \eqe
568: %*******
569: Just like magnetic monopoles are isolated defects
570: in the electric phase there are isolated and closed
571: magnetic flux lines in the magnetic phase\footnote{The fact that
572: only closed loops occur is explained by the absence of isolated
573: magnetic charges in a monopole condensate.} which,
574: however, collapse as soon as they are created if the magnetic coupling $g$ is finite:
575: Abrikosov-Nielsen-Olesen (ANO) vortex loops \cite{NielsenOlesen1973}.
576: Only one unit of flux with respect to U(1)$_D$ (SU(2)) or either factor
577: in U(1)$^2_D$ (SU(3)) is carried by a
578: given vortex loop since in the electric phase
579: only charge-one calorons dissociate into magnetic monopoles
580: with one unit of magnetic charge\footnote{The core of a vortex line, where U(1)$_D$ (SU(2)) or one factor in
581: U(1)$^2_D$ (SU(3)) is restored, can be pictured as a directed motion of
582: magnetic monopoles (to the right) and antimonopoles (to the left)
583: in the rest frame of the heatbath\cite{Olejnik1997}, see Fig.\,\ref{Fig0b}. The magnetic flux,
584: which penetrates a spatial
585: hyperplane perpendicular to
586: the direction of monopole or
587: antimonopole motion, is by Stoke's theorem
588: measured by a line integral $g\oint_{{\cal C}} dz_\mu A^D_{\mu}$ along a circular
589: curve ${\cal C}$ with infinite radius lying in this plane. Here $A^D_\mu$
590: denotes the gauge field with respect to the dual gauge group U(1) (SU(2)) or either factor
591: in U(1)$^2$ (SU(3)) generated by the moving chains of
592: monopoles and antimonopoles. If we choose to evaluate the
593: line integral in a covariant gauge then the contribution to $dz_\mu A^D_\mu$
594: of each moving monopole or antimonopole is that of a
595: static monopole or antimonopole since the perpendicular part
596: of the gauge field is invariant under boosts
597: along the vortex axis. Thus the state of
598: motion of each monopole and antimonopole
599: is irrelevant for its effect on the total magnetic flux
600: carried by the vortex as long as the
601: net motion of all monopoles and antimonopoles in a given segment
602: defines the direction of the vortex axis: The only thing that determines
603: the magnetic flux of the vortex line is the {\sl charge} of a monopole.}.
604: %***********************
605: \begin{figure}
606: \begin{center}
607: \leavevmode
608: %\epsfxsize=9.cm
609: \leavevmode
610: %\epsffile[80 25 534 344]{}
611: \vspace{4.3cm}
612: \special{psfile=Fig-2.ps angle=0 voffset=-100
613: hoffset=-105 hscale=55 vscale=75}
614: \end{center}
615: \caption{Microscopics of the core of a center vortex. Monopoles (M) and antimonopoles (A) move in
616: opposite directions.\label{Fig0b}}
617: \end{figure}
618: %************************
619: This allows for an interpretation of ANO vortex loops as center-vortex loops. For SU($N$)
620: the magnetic flux of the latter is determined by the differences
621: in phase modulo $N$ of two center elements, see Eq.\,(\ref{noho}). There is one unit of center
622: flux for SU(2), and there are two separate units of center flux for SU(3).
623:
624: The core-size of a center vortex is determined
625: by the length $l_g\sim (g|\varphi|)^{-1}$ (SU(2)) or
626: $l_{g,1}\sim (g|\varphi_1|)^{-1},l_{g,2}\sim (g|\varphi_2|)^{-1}$ (SU(3)) of penetration of the vortex' gauge
627: field into a direction perpendicular to the vortex. While center vortices are
628: thick close to $T_{c,E}$ their core size vanishes at $T_{c,M}$. Since the energy of a typical
629: center-vortex loop\cite{NielsenOlesen1973} is $\propto g^{-1}$ spin-0 systems composed of a vortex loop and its
630: antivortex loop condense at $T_{c,M}$. We will show that there is a parameter with discrete values
631: characterizing the possible values of a macroscopic,
632: complex scalar field $\Phi$ which describes the vortex condensate.
633:
634: The transition to the center phase is
635: of the Hagedorn type and thus nonthermal. An order parameter for this transition is the
636: expectation of the 't Hooft loop whose modulus measures the
637: strength of center-vortex condensation. If the 't Hooft
638: loop does not vanish then the {\sl magnetic} $Z_2$ (SU(2)) or $Z_3$ (SU(3))
639: symmetry is dynamically broken. These center symmetries are
640: {\sl local} in four-dimensional spacetime. Under large U(1)$_D$ (SU(2))
641: or large U(1)$^2_D$ (SU(3)) gauge transformations
642: the macroscopic field $\Phi$ transforms by multiplications
643: with center elements.
644:
645: In the course of the Hagedorn transition the ground state of the magnetic
646: phase decays through creation of single and
647: self-intersecting center-vortex loops. In contrast to the magnetic phase,
648: center-vortex loops are stable in the center phase, thus are
649: particle-like, and possess a density of states which is
650: over-exponentially rising with energy.
651: There are precisely two polarization states for each self-intersecting or single
652: loop (spin-1/2 fermions). After the decay of the monopole condensate is completed
653: the new ground state is a Cooper-pair condensate of systems composed of a
654: massless single vortex and antivortex
655: loop. The energy-momentum tensor on this ground state {\sl vanishes identically}. This result
656: is protected against radiative corrections.
657:
658: The center phase is truly confining: A pair of electric, static, oppositely and fundamentally charged, and
659: fermionic test charges forms a confining electric flux tube because of
660: the presence of condensed electric dipoles
661: (single center-vortex loops) in the
662: ground state, and all gauge modes are infinitely
663: heavy. Notice that the absence of propagating gauge modes implies that
664: only contact interactions are possible between (thin) center fluxes and the
665: self-intersection points of their vortex loops.
666:
667: Thermal lattice simulations fail to produce physical results for infrared sensitive quantities
668: at temperatures shortly below $T_{c,E}$: The center
669: phase as well as the magnetic phase possess infinite correlation lengths. In the center phase
670: this correlation length is given by the inverse (vanishing) mass of a single center-vortex loop. Thus the fermionic
671: nature of excitations and their over-exponentially rising density of states in the
672: center phase escapes simulations performed on finite-size lattices.
673:
674: It is self-evident that what was said above has implications
675: for particle physics and cosmology. We only would like to
676: mention a few questions that are likely to
677: be answered by SU(2) Yang-Mills (thermo)dynamics alone:
678: Electroweak symmetry breaking, namely,
679: the origin of the masses of $Z_0$ and $W^\pm$; the mass hierarchy
680: between the two members of a lepton family; the question of whether
681: the neutrino is Dirac or Majorana; the smallness of the
682: cosmological constant on particle physics scales; the nature of
683: cosmological and clustering dark matter; cosmic coincidence; baryon and lepton asymmetries;
684: intergalactic magnetic fields; and large-angle
685: anomalies in some of the power spectra of fluctuations in the cosmic
686: microwave background.
687:
688: The article is organized as follows. In the first part of Sec.\,\ref{EP}
689: we provide prerequisites on caloron physics. The classical solutions of trivial and nontrivial holonomy
690: and their behavior under quantum corrections are discussed for SU(2). As an aside we
691: estimate the separation of isolated and screened monopoles in terms of
692: the inverse temperature in the electric phase. The second part of Sec.\,\ref{EP}
693: is devoted to the derivation of the macroscopic, adjoint
694: field $\phi$ in terms of a moduli-space and $S_3$ integral over the spatial two-point
695: correlations in a caloron-anticaloron system. Subsequently, we discuss the full ground-state physics
696: in the electric phase. The existence of a dynamically generated scale $\La_E$ needs to be assumed
697: to determine $\phi$'s modulus at a given temperature. We show that a degeneracy
698: with respect to a global electric
699: $Z_2$ (SU(2)) and $Z_3$ (SU(3)) symmetry occurs which proves that
700: this phase is deconfining. The last part of Sec.\,\ref{EP} addresses the full
701: thermodynamics of the electric phase, including excitations. An evolution equation
702: for the effective gauge coupling $e$ is derived and solved numerically. It is observed
703: that an ultraviolet-infrared decoupling is manifest in this evolution.
704: We present analytical results for the electric screening mass, associated
705: with the massless mode in the SU(2) case, and for the two-loop
706: correction to the pressure.
707:
708: In Sec.\,\ref{MP} we investigate the magnetic phase. Prerequisites on
709: the BPS monopole are given. Considering the average magnetic flux, generated by a
710: screened monopole-antimonopole system in a thermal environment,
711: through an $S_2$ with infinite radius, where the monopole-antimonopole system is
712: located outside of this $S_2$, a continuous parameter (proportional to the Euclidean time)
713: is derived. This parameter governs
714: the temporal winding of a spatially homogeneous, complex, and
715: inert scalar field $\varphi$ in the limit where monopole mass and spatial
716: momentum of the system vanish (Bose condensation). For SU(3) two such
717: fields exist, each for every independent monopole species. Assuming the existence
718: of a dynamically generated scale $\La_M$, the modulus of $\varphi$ is derived. We then discuss
719: the ground-state physics in the magnetic phase. It is shown that the electric $Z_2$ (SU(2)) and electric
720: $Z_3$ (SU(3)) degeneracy of the ground state, as observed in the electric phase, gives
721: way to a unique ground state in the magnetic phase. Thus we derive test-charge
722: confinement in the magnetic phase. Evolution equations for the effective
723: magnetic coupling $g$ are derived and solved numerically. A discussion of the full
724: Polyakov-loop expectation (including excitations) and an analysis of the critical behavior at the
725: electric-magnetic phase boundary are presented. We find that this transition is
726: second order with mean-field exponents both for SU(2) and SU(3) with the difference that the
727: peak in the specific heat is about three times smaller in the former as compared to the latter case.
728:
729: In Sec.\,\ref{CVCM} we investigate the center phase. We start by providing prerequisites
730: on the ANO vortex. In particular, we emphasize the fact that an ANO vortex generates
731: negative pressure at finite magnetic coupling $g$. While a vortex-loop is a particle-like
732: excitation at $g=\infty$ it collapses as soon as it is created for $g<\infty$. Collapsing ANO or
733: center-vortex loops dominate the (negative) pressure inside the
734: magnetic phase where $g<\infty$. In analogy to the monopole condensate we determine a parameter (mean center
735: flux through an $S_1$ of infinite radius) which governs the expectation of a macroscopic,
736: complex scalar field $\Phi$ describing the Bose condensate
737: of massless vortex-antivortex-loop pairs. The values of this parameter are discrete: Two possible values
738: for SU(2) and three possible values for SU(3). At $T_{c,M}$, where $g$ diverges in a logarithmic
739: way and where dual gauge modes acquire an infinite mass, the center-vortex
740: condensate starts to form under (spin-1/2)
741: particle creation. We construct an effective potential $V_C$ for $\Phi$ involving a scale $\La_C$. We
742: check $V_C$'s uniqueness, and
743: discuss how vortex-loop creation takes place by center jumps of $\Phi$'s phase. An estimate for the
744: density of fermion states, created by $\Phi$'s relaxation to zero energy density and
745: pressure, is provided. As a result, we analytically establish
746: that the center-magnetic transition is of the Hagedorn type.
747:
748: In Sec.\,\ref{MCC} we discuss how the scales $\La_E$ and $\La_M$ are related
749: by the continuity of the pressure across the electric-magnetic phase boundary. We also provide
750: an approximate relation between $\La_M$ and $\La_C$.
751:
752: In Sec.\,\ref{PEEN} we present numerical results for the temperature dependence
753: of thermodynamical quantities thoughout
754: the electric and the magnetic phase. The following quantities are
755: discussed: Pressure, energy density, interaction measure,
756: specific heat per volume, and entropy density. While the former quantities are very
757: sensitive to the ground-state physics at low temperatures, which is determined by
758: very large spatial correlations and thus is inaccessible to finite-size lattices,
759: the entropy density is only
760: sensitive to the excitations. This fact makes a comparison of our results for
761: the entropy density with those obtained on lattices
762: (employing the differential method) useful, all other quantities exhibit quantitative
763: disagreements with their lattice-obtained values at low temperatures.
764:
765: In Sec.\,\ref{Apps} we discuss implications of our work
766: for particle physics and cosmology. We start by addressing the
767: cosmic-coincidence and the old cosmological-constant
768: problem in view of a Planck-scale axion, originating from dynamically generated and subsequently
769: integrated spin-1/2 fermions at the Planck scale, and in view of an SU(2) gauge theory of
770: Yang-Mills scale close to the temperature of the cosmic microwave background (CMB)
771: (SU(2)$_{\tiny\mbox{CMB}}$). This theory is at the electric-magnetic
772: phase boundary, and its only massless and unscreened excitation is the photon. Throughout cosmological
773: evolution the axion mass is provided by the axial anomaly involving nonconfining
774: SU(N) gauge theories of Yang-Mills scales lower than the Planck scale. The presence of
775: a Planck-scale axion may explain the particle-number asymmetries and
776: CP violation in the weak interactions. Some of the phenomenology of the
777: electroweak sector of the SM is addressed in view of leptons being
778: stable solitons in the center phase of various
779: SU(2) Yang-Mills theories. These solitons are embedded into instable higher-charge excitations with
780: an over-exponentially rising density of states which protect their apparent
781: structurelessness seen in scattering experiments
782: (the photon couples to the lepton because of mixing)
783: up to center-of-mass energies comparable to the mass
784: of the $Z$ boson (with exceptions at momenta comparable to the lepton masses). We postdict the mass ratio
785: $\frac{m_{\nu_e}}{m_e}$ in terms of the mass ratio $\frac{m_{e}}{m_Z}$.
786: Finally, we present some ideas on how fractionally charged light
787: quarks and their confinement may arise in Quantum Chromodynamics (QCD) in terms
788: of electric-magnetically dual SU(3) gauge dynamics and the
789: fractional Quantum Hall effect.
790:
791: The last section of the present work briefly summarizes our results.
792:
793:
794:
795: \section{The electric phase\label{EP}}
796:
797: \subsection{Prerequisites}
798:
799: \subsubsection{The Harrington-Shepard solution (trivial holonomy)\label{HSS}}
800:
801: Calorons of trivial holonomy are the field configurations which
802: enter the definition of the phase of the macroscopic adjoint scalar
803: field $\phi$. We only need to consider the SU(2)
804: case since the SU(3) ground-state thermodynamics can be
805: derived from a 'democratic' embedding of SU(2) calorons. We use the nonperturbative
806: definition of the gauge field where the coupling
807: constant is absorbed into the field.
808:
809: (Anti)Calorons are (anti)selfdual, that is, their
810: field strength is up to a sign equal to
811: their dual field strength
812: %*********
813: \eqb
814: \label{BPSCal}
815: F_{\mu\nu}[A^{(C,A)}]=\pm\tilde{F}_{\mu\nu}[A^{(C,A)}]\,
816: \eqe
817: %********
818: where the superscript $C(A)$ refers to caloron (anticaloron).
819: Only calorons of topological charge one (minus one)
820: enter the definition of $\phi$'s phase and
821: thus we will focus on this case only.
822:
823: The Harrington-Shepard solutions \cite{HarrigtonShepard1977} are given as
824: %*********
825: \eab
826: \label{HS}
827: A^C_\mu(\tau,\vec{x})&=&\bar{\eta}_{a\mu\nu}\frac{\lambda^a}{2}\pd_{\nu}\ln \Pi(\tau,\vec{x})\,\ \ \ \mbox{or}\nonumber\\
828: A^A_\mu(\tau,\vec{x})&=&\eta_{a\mu\nu}\frac{\lambda^a}{2}
829: \pd_{\nu}\ln \Pi(\tau,\vec{x})\,
830: \eae
831: %********
832: where the 't Hooft symbols $\eta_{a\mu\nu}$ and $\bar{\eta}_{a\mu\nu}$
833: are defined by
834: %********
835: \eab
836: \label{tHooftsym}
837: \eta_{a\mu\nu} &=& \epsilon_{a\mu\nu} + \delta_{a\mu}\delta_{\nu4} - \delta_{a\nu}\delta_{\mu4} \nonumber\\
838: \bar\eta_{a\mu\nu} &=& \epsilon_{a\mu\nu} - \delta_{a\mu}\delta_{\nu4} + \delta_{a\nu}\delta_{\mu4}\,.
839: \eae
840: %**********
841: In Eq.\,(\ref{HS}) $\lambda^a$, ($a=1,2,3$),
842: denote the Pauli matrices. The periodic solutions in Eq.\,(\ref{HS}), $A^{C,A}_\mu(0,\vec{x})=
843: A^{C,A}_\mu(1/T,\vec{x})\,,$
844: are generated by a temporal mirror sum of
845: the 'pre'potential
846: %*********
847: \eqb
848: \label{PsI}
849: \Pi_0=1+\frac{\rho^2}{x^2}
850: \eqe
851: %*********
852: of a single BPST (anti)instanton of scale $\rho$ \cite{BPST}
853: in singular gauge \cite{Atiyah1978}.
854: Here $x^2\equiv\tau^2+\vec{x}^2$. The scalar function $\Pi(\tau,\vec{x})$
855: in Eq.\,(\ref{HS}) is given as
856: %***********
857: \eab
858: \label{Pi}
859: \Pi(\tau,\vec{x})&=&\sum_{n=-\infty}^{\infty}\frac{\rho^2}{(\tau-n\beta,\vec{x})^2}\nonumber\\
860: &=&\bar{\Pi}(\tau,r)\equiv1+\frac{\pi\rho^2}{\beta r}
861: \frac{\sinh\left(\frac{2\pi r}{\beta}\right)}{\cosh\left(\frac{2\pi r}{\beta}\right)-
862: \cos\left(\frac{2\pi\tau}{\beta}\right)}\,
863: \eae
864: %*********
865: where $r\equiv|\vec{x}|$ and $\beta\equiv 1/T $. Evaluating the integral
866: of the Chern-Simons current over a
867: small three-sphere $S_3$, centered at the singular point
868: $(\tau=0,\vec{x}=0)$, one obtains plus (or minus) one unit of
869: topological charge. For a given value of
870: $\rho$ the solutions in Eq.\,(\ref{HS}) can be generalized by
871: shifting the center from $z=0$ to $z=(\tau_z,\vec{z})$ by the (quasi)
872: translational invariance of the classical action. (The temporal shift $\tau_z$ is
873: restricted to $0\le\tau_z\le\beta$ because of periodicity.) In addition, the color orientation of each
874: solution can be rotated by global gauge transformations.
875:
876: Computing the Polyakov loop at spatial infinity on either of
877: the configurations $A^C_\mu(\tau,\vec{x})$ and $A^A_\mu(\tau,\vec{x})$ yields
878: the following result
879: %*********
880: \eqb
881: \label{Polyainf}
882: {\bf P}(|\vec x|\to\infty)=
883: {\cal P}\exp\left[i\int_0^\beta d\tau A^{C,A}_4(\tau,|\vec x|\to\infty)
884: \right]=\UM\,.
885: \eqe
886: %*********
887: Thus the Harrington-Shepard solutions possess trivial holonomy.
888:
889: \subsubsection{The Lee-Lu-Kraan-van Baal
890: solution (nontrivial holonomy)}
891:
892: For a discussion of (anti)selfdual SU(2) configurations
893: with nontrivial holonomy and topological charge one (minus one)
894: we use the conventions and closely follow the presentation
895: of \cite{LeeLu1998} which to our taste makes the
896: magnetic monopole content most explicit. (In \cite{Diakonov2004} the constituents
897: of nontrivial-holonomy calorons
898: are referred to as dyons because
899: the nonabelian magnetic and electric field
900: of each constituent is equal and Coulomb-like for large
901: distances away from a given monopole core. This property, however, follows
902: from the selfduality of the caloron
903: configuration. With respect to the unbroken
904: U(1) the charge of a constituent monopole
905: is purely magnetic (and not dyonic as in \cite{JuliaZee1975})
906: since the $A_4$ field serves as a Higgs field and not as
907: the gauge potential for the electric field.) The existence of these solutions was shown
908: by Nahm \cite{Nahm1984}. Explicit analytical constructions
909: were independently performed by Lee and Lu
910: \cite{LeeLu1998} and Kraan and
911: van Baal \cite{KraanVanBaalNPB1998,vanBaalKraalPLB1998}.
912:
913: Lee and Lu use antihermitian generators and parametrize
914: the holonomy $u$ as
915: %***********
916: \eqb
917: \label{holonomy}
918: A^C_4(\tau,|\vec x|\to\infty)=-i\frac{u}{2}\lambda_3\,
919: \eqe
920: %***********
921: where $0\le u\le \frac{2\pi}{\beta}$. Using the
922: Nahm data for a monopole coexisting with an antimonopole as an
923: input to the Atiyah-Drinfeld-Hitchin-Manin-Nahm (ADHMN)
924: equations (subject to a normalization condition), a
925: selfdual field configuration with monopole-antimonopole constituents
926: was constructed in \cite{LeeLu1998}. It reads
927: %***********
928: \eab
929: \label{caloronnth}
930: A_\mu(\vec{x},\tau)&=&C_1^\dagger V_\mu(\vec{y}_1;u)C_1+
931: C_2^\dagger V_\mu(\vec{y}_2;\frac{2\pi}{\beta}-u)C_2+\nonumber\\
932: &&C^\dagger_1\pd_\mu C_1+C^\dagger_1\pd_\mu C_1+
933: C^\dagger_2\pd_\mu C_2+S^\dagger\pd_\mu S\,
934: \eae
935: %**********%***********************
936: \begin{figure}
937: \begin{center}
938: \leavevmode
939: %\epsfxsize=9.cm
940: \leavevmode
941: %\epsffile[80 25 534 344]{}
942: \vspace{5.0cm}
943: \special{psfile=Fig-3.ps angle=0 voffset=-130
944: hoffset=-130 hscale=50 vscale=50}
945: \end{center}
946: \caption{Meaning of the spatial arguments $\vec{y}_1,\vec{y}_2$ entering the
947: solution in Eq.\,(\protect\ref{caloronnth}). The points $\vec{x}_1,\vec{x}_2$ are the
948: core positions of the monopole and the antimonopole. At the
949: point $\tau=0,\vec{x}_{\tiny\mbox{cm}}$ the solution is singular.\label{Fig-1b}}
950: \end{figure}
951: %************************
952: where
953: %********
954: \eab
955: \label{Vmu}
956: V_4(\vec{x};u)&=&\frac{\lambda_a}{2i}\hat{x}_a\left(\frac{1}{|\vec{x}|}-
957: \frac{u}{\coth(u|\vec{x}|)}\right)\,,\nonumber\\
958: V_i(\vec{x};u)&=&\frac{\lambda_a}{2i}\epsilon_{aij}\hat{x}_j\left(\frac{1}{|\vec{x}|}-
959: \frac{u}{\sinh(u|\vec{x}|)}\right)\,.
960: \eae
961: %*********
962: Interpreting $V_4(\vec{x};u)$ as an adjoint Higgs field, Eqs.\,(\ref{Vmu}) represent
963: the BPS magnetic monopole \cite{PrasadSommerfield1974}. The matrices $C_1,C_2$
964: in Eq.\,(\ref{caloronnth}) are given as
965: %*********
966: \eab
967: \label{C1C2}
968: C_1&=&\sqrt{\frac{2D N_1}{{\cal N}}}\frac{B_1^\dagger}{{\cal M}}\left[\exp\left(-\frac{\vec{\lambda}}{2}
969: \cdot \vec{s}_2\right)Q_++\exp\left(\frac{\vec{\lambda}}{2}
970: \cdot \vec{s}_2\right)Q_-\right]\exp\left(-i\frac{\pi}{\beta}\tau\lambda_3\right)\,,\nonumber\\
971: C_2&=&\sqrt{\frac{2D N_2}{{\cal N}}}\frac{B_2^\dagger}{{\cal M}}\left[\exp\left(\frac{\vec{\lambda}}{2}
972: \cdot \vec{s}_1\right)Q_++\exp\left(-\frac{\vec{\lambda}}{2}
973: \cdot \vec{s}_2\right)Q_-\right]
974: \eae
975: %**********
976: where $Q_\pm=\frac{1}{2}(1\pm\lambda_3)$ are projection operators. The
977: matrices $B_1,B_2$ are
978: %**********
979: \eab
980: \label{B1B2}
981: B_1&=&\exp\left[i\frac{\pi}{\beta}\tau\right]\exp\left[-\frac{\vec{\lambda}}{2}
982: \cdot\vec{s_1}\right]\exp\left[-\frac{\vec{\lambda}}{2}
983: \cdot\vec{s_2}\right]-\nonumber\\
984: &&\exp\left[-i\frac{\pi}{\beta}\tau\right]\exp\left[\frac{\vec{\lambda}}{2}
985: \cdot\vec{s_1}\right]\exp\left[\frac{\vec{\lambda}}{2}
986: \cdot\vec{s_2}\right]\,,\nonumber\\
987: B_2&=&\exp\left[i\frac{\pi}{\beta}\tau\right]\exp\left[-\frac{\vec{\lambda}}{2}
988: \cdot\vec{s_2}\right]\exp\left[-\frac{\vec{\lambda}}{2}
989: \cdot\vec{s_1}\right]-\nonumber\\
990: &&\exp\left[-i\frac{\pi}{\beta}\tau\right]\exp\left[\frac{\vec{\lambda}}{2}
991: \cdot\vec{s_2}\right]\exp\left[\frac{\vec{\lambda}}{2}
992: \cdot\vec{s_1}\right]\,
993: \eae
994: %**********
995: and the scalar ${\cal M}$ is defined as
996: %********
997: \eqb
998: \label{M}
999: {\cal M}=2\left(\cosh s_1\cosh s_2+\hat{y}_1\cdot\hat{y}_2\sinh s_1\sinh
1000: s_2-\cos\left[\frac{2\pi}{\beta}\tau\right]\right)\,.
1001: \eqe
1002: %*********
1003: In addition, one defines
1004: %********
1005: \eqb
1006: \label{Ni}
1007: N_i=\frac{1}{y_i}\sinh s_i\,,\ \ \ \ \ (i=1,2)\,,
1008: \eqe
1009: %********
1010: and
1011: %******
1012: \eqb
1013: \label{N}
1014: {\cal N}=1+\frac{2D}{{\cal M}}\left(N_1(\cosh s_2-(\hat{y}_2)_3\sinh s_2)+
1015: N_2(\cosh s_1+(\hat{y}_1)_3\sinh s_1)\right)\,,
1016: \eqe
1017: %********
1018: and
1019: %********
1020: \eqb
1021: \label{S}
1022: S=\frac{1}{\sqrt{\cal N}}\,\exp\left[-i\frac{u}{2}\tau\lambda_3\right]\,.
1023: \eqe
1024: %********
1025: The spatial arguments of the configuration in Eq.\,(\ref{caloronnth}),
1026: compare with Fig.\,\ref{Fig-1b}, are
1027: defined as
1028: %*********
1029: \eab
1030: \label{geom}
1031: \vec{y}_i&=&\vec{x}-\vec{x}_i\,,\ \ \ \ \ y_i=|\vec{y}_i|\,,\ \ \ \ \ \ \ \ s_i=|\vec{s}_i|\,\ \ \ \ \ \ \ (i=1,2)\,, \nonumber\\
1032: \vec{s}_1&=&u\vec{y}_1\,,\ \ \ \ \ \ \ \ \vec{s}_2=\left(\frac{2\pi}{\beta}-u\right)\vec{y}_2\,,\ \ \ \
1033: D=|\vec{x}_2-\vec{x}_1|\,.
1034: \eae
1035: %*********
1036: (A $\hat{\mbox{}}$ -sign indicates a unit vector.) Kraan and van Baal show that the distance $D$
1037: between the two BPS monopoles can be expressed by the scale $\rho$
1038: of a trivial-holonomy caloron which is deformed
1039: to nontrivial holonomy \cite{vanBaalKraalPLB1998}. One has
1040: %******
1041: \eqb
1042: \label{relrhoD}
1043: D=\frac{\pi}{\beta}\rho^2\,.
1044: \eqe
1045: %******
1046: %***********************
1047: \begin{figure}
1048: \begin{center}
1049: \leavevmode
1050: %\epsfxsize=9.cm
1051: \leavevmode
1052: %\epsffile[80 25 534 344]{}
1053: \vspace{4.3cm}
1054: \special{psfile=Fig-4.ps angle=0 voffset=-140
1055: hoffset=-205 hscale=55 vscale=75}
1056: \end{center}
1057: \caption{Action density of an SU(2) caloron with nontrivial holonomy plotted on a two-dimensional spatial slice.
1058: The caloron radius $\rho$ and therefore the separation $D$ (Eq.\,(\protect\ref{relrhoD}))
1059: increases from left to right while temperature and holonomy are fixed. Figures are taken
1060: from a paper by Kraan and van Baal. The peaks of the action density
1061: coincide with the core positions of the constituent BPS monopoles.\label{instrad}}
1062: \end{figure}
1063: %************************
1064: It was shown in \cite{LeeLu1998}
1065: that for $y_1\ll D$
1066: %*******
1067: \eqb
1068: \label{asy}
1069: C_2,S\sim \frac{1}{\sqrt{D}}
1070: \eqe
1071: %*******
1072: and
1073: %******
1074: \eqb
1075: \label{C1asy}
1076: C_1=\frac{\lambda_3\cosh\frac{s_1}{2}-\vec{\lambda}\cdot\hat{y}_1\sinh\frac{s_1}{2}}{\sqrt{\cosh s_1-
1077: (\hat{y}_1)_3\sinh s_1}}+{\cal O}(1/D)\,.
1078: \eqe
1079: %*******
1080: Thus $C_1$ is a single-valued unitary matrix, and for
1081: $y_1\ll D$ the configuration in Eq.\,(\ref{caloronnth}) is an approximate
1082: gauge transform of a BPS monopole. Similarily, for $y_2\ll D$ $C_2$
1083: is a unitary matrix. The difference as compared with $C_1$ for $y_1\ll D$
1084: is that for $y_2\ll D$ the matrix $C_2$ induces a large gauge rotation due to an extra
1085: factor $\exp\left[-i\frac{\pi}{\beta}\tau\lambda_3\right]$. This gauge transformation
1086: inverts the charge of the BPS monopole at $\vec{x}_2$ as compared to the charge of
1087: the BPS monopole at $\vec{x}_1$. There is a singularity of the
1088: solution at the point $(\tau=0,\vec{x}_{\tiny\mbox{cm}})$ where
1089: %********
1090: \eqb
1091: \label{cmpoint}
1092: \vec{x}_{\tiny\mbox{cm}}=\frac{\beta u}{2\pi}\vec{x}_1+
1093: \left(1-\frac{\beta u}{2\pi}\right)\vec{x}_2\,.
1094: \eqe
1095: %*********
1096: This point carries one unit of topological
1097: charge. One can show this by expanding the solution about
1098: $(\tau=0,\vec{x}_{\tiny\mbox{cm}})$ and by
1099: performing the integral of
1100: the Chern-Simons current over a
1101: small $S_3$ centered at this point. A plot of the action
1102: density of a nontrivial-holonomy caloron with varying radius $\rho$ at a fixed temperature and a fixed
1103: holonomy is presented in Fig.\,\ref{instrad}.
1104:
1105: On the classical level the masses of the monopoles
1106: at $\vec{x}_1$ and $\vec{x}_2$ are given as
1107: %***********
1108: \eqb
1109: \label{massesMonLL}
1110: m_1=4\pi u\,\ \ \ \ \ \ m_2=4\pi \left(\frac{2\pi}{\beta}-u\right)\,,
1111: \eqe
1112: %***********
1113: respectively. For a large holonomy, that is $u\sim\frac{\pi}{\beta}$,
1114: we have $m_1\sim m_2\sim 4\pi^2\,T\sim 40\,T$. Thus large holonomy is extremely Boltzmann
1115: suppressed.
1116:
1117:
1118: \subsubsection{One-loop quantum weights}
1119:
1120: In this section we first present the results for the one-loop
1121: effective action of a trivial-holonomy caloron,
1122: which was obtained by Gross, Pisarski, and Yaffe
1123: \cite{GrossPisarskiYaffe1981} by appealing to the
1124: results obtained in \cite{BrownCarlitzLee1977,BrownCarlitzCreamerLee1978,BrownCreamer1978} and the
1125: pioneering work of 't Hooft \cite{Hooft1976}.
1126: Subsequently, we sketch the results obtained recently by
1127: Diakonov, Gromov, Petrov, and Slizovskiy \cite{Diakonov2004} for the
1128: one-loop quantum weight of a caloron with nontrivial holonomy.
1129: Both results are important for a grasp of
1130: the microcopics of the ground-state
1131: physics in the electric phase.\vspace{0.1cm}\\
1132:
1133: \noindent\underline{Harrington-Shepard solution:}\vspace{0.1cm}\\
1134:
1135: The functional determinant around a
1136: Harrington-Shepard caloron was
1137: calculated in \cite{GrossPisarskiYaffe1981}. The result for the
1138: quantum weight $\exp[-S_{\tiny\mbox{eff}}]$
1139: of this configuration is given in terms of the effective action as
1140: %***********
1141: \eqb
1142: \label{effactthcal}
1143: S_{\tiny\mbox{eff}}=\frac{8\pi^2}{\bar{g}^2}+\frac{4}{3}\sigma^2+16\,A(\sigma)\,
1144: \eqe
1145: %***********
1146: where the dimensionless quantity $\sigma$ is defined as
1147: $\sigma\equiv \pi\frac{\rho}{\beta}$ and
1148: %********
1149: \eqb
1150: \label{A(alpha)}
1151: A(\sigma)\equiv\frac{1}{12}\left[\int_0^\beta d\tau \frac{d^3x}{16\pi^2}
1152: \left(\frac{(\pd_\mu \Pi)^2}{\Pi^2}\right)^2-\int \frac{d^4x}{16\pi^2}
1153: \left(\frac{(\pd_\mu \Pi_0)^2}{\Pi_0^2}\right)^2\right]\,.
1154: \eqe
1155: %*********
1156: The weight $\exp[-S_{\tiny\mbox{eff}}]$ is relevant for
1157: the integration over the classical moduli space in the presence of one-loop quantum fluctuations.
1158: (The nonflat metric is the same as in the zero-temperature situation.)
1159:
1160: In Eq.\,(\ref{A(alpha)}) the scalar quantities $\Pi$ and $\Pi_0$ are
1161: defined in Eqs.\,(\ref{Pi}) and (\ref{PsI}), respectively. The first
1162: integral in Eq.\,(\ref{A(alpha)}) is over $S_1\times \R^3$ while the second integral
1163: is over $\R^4$. It is worth mentioning how $A(\sigma)$ behaves in the
1164: high- and low-temperature limits $\sigma\to\infty$ and $\sigma\to 0$:
1165: %********
1166: \eqb
1167: \label{hlTL}
1168: A(\sigma)\to -\frac{1}{6}\log\sigma\,,\ \ \ (\sigma\to\infty)\,,\ \ \ \
1169: A(\sigma)\to -\frac{\sigma^2}{36}\,,\ \ \ (\sigma\to 0)\,.
1170: \eqe
1171: %********
1172: At a given caloron radius $\rho$ the correction to the classical action
1173: $\frac{8\pi^2}{\bar{g}^2}$ thus is large in the
1174: high-temperature regime, indicating that the contribution of
1175: trivial-holonomy calorons to the partition function is
1176: suppressed, while it is small at low temperatures,
1177: implying the increasing importance of calorons as the
1178: temperature of the system drops. The distinction between high and
1179: low temperatures is made by a dynamically generated scale $\Lambda_E$ which also
1180: determines the $\rho$ dependence of the
1181: coupling constant $\bar{g}$ in Eq.\,(\ref{effactthcal}). The latter, by zero-temperature
1182: one-loop renormalization-group running \cite{GrossWilczek1973,Politzer1973},
1183: estimatedly becomes larger than unity for $\rho^{-1}\sim\Lambda_E$ and is
1184: logarithmically small for $\rho^{-1}\gg\Lambda_E$. \vspace{0.1cm}\\
1185:
1186: \noindent\underline{Lee-Lu-Kraan-van Baal
1187: solution}\vspace{0.1cm}\\
1188:
1189: The calculation of the one-loop quantum weight for a caloron of
1190: nontrivial holonomy is much harder than for the trivial case. This explains why this result only
1191: appeared in the literature \cite{Diakonov2004} more than six years
1192: after the analytical form of the nontrivial-holonomy solution was published in
1193: \cite{LeeLu1998,vanBaalKraalPLB1998}. The expressions
1194: are so involved that the contribution ${\cal Z}_{\tiny\mbox{n.h.}}$
1195: of an isolated, quantum-blurred caloron to the total
1196: partition function ${\cal Z}$ of the theory has so far only been stated in closed analytical form
1197: in the limit
1198: %*********
1199: \eqb
1200: \label{limitDiakonov}
1201: \frac{D}{\beta}=\pi\left(\frac{\rho}{\beta}\right)^2\gg 1\,.
1202: \eqe
1203: %*********
1204: This, however, is
1205: the relevant physical situation, see
1206: Sec.\,\ref{EC} where is is shown
1207: that (trivial-holonomy) calorons with $\rho\gg \beta$
1208: dominate the phase of the macroscopic adjoint
1209: scalar field $\phi$. As we shall see, the
1210: ground-state physics in the electric phase is dominated by small-holonomy deformations of
1211: the trivial case.
1212:
1213: Apart from the restriction in Eq.\,(\ref{limitDiakonov})
1214: the result obtained in \cite{Diakonov2004} for ${\cal Z}_{\tiny\mbox{n.h.}}$ is valid
1215: for any value of the holonomy, $0\le u\le \frac{2\pi}{\beta}$. After the (trivial)
1216: integrations of the overall color orientation
1217: and time translations are performed one
1218: obtains \cite{Diakonov2004}
1219: %**********
1220: \eab
1221: \label{Zcalonh}
1222: {\cal Z}_{\tiny\mbox{n.h.}}&=C\beta^{-6}&\int d^3 x_1\int d^3 x_2\,
1223: \left(\frac{8\pi^2}{\bar{g}^2}\right)^4\left(\frac{\La\e^{\gamma_E}\beta}{4\pi}\right)^{22/3}
1224: \left(\frac{\beta}{D}\right)^{5/3}\times\nonumber\\
1225: &&(2\pi+\beta u\bar{u} D)(uD+1)^{\frac{4}{3\pi}u\beta-1}(\bar{u}D+1)^{\frac{4}{3\pi}\bar{u}\beta-1}\times\nonumber\\
1226: &&\exp[-V\,P(u)-2\pi DP^{\prime\prime}(u)]\,,\ \ \ \ \ (D\gg\beta).
1227: \eae
1228: %***********
1229: (The number of independent integration
1230: variables in Eq.\,(\ref{Zcalonh}) is four because $\int d^3 x_1\int d^3 x_2=4\pi\int d^3x\int dD D^2$
1231: where $\vec{x}=\frac{1}{2}\left(\vec{x}_1+\vec{x}_2\right)$).
1232: In Eq.\,(\ref{Zcalonh}) $C\sim 1.0314$, $\vec{x}_1$ and $\vec{x}_2$ are the core positions
1233: of the monopoles in the classical solution
1234: (compare with Fig.\,\ref{Fig-1b}), $\bar{u}\equiv\frac{2\pi}{\beta}-u$, $\gamma_E$ is the Euler constant,
1235: $V$ denotes the typical spatial volume belonging to the one-caloron system,
1236: and $\La$ is a scale which is a one-loop
1237: renormalization group invariant (dimensional transmutation). The functions $P(u)$
1238: and $P^{\prime\prime}(u)$ are given as
1239: %*********
1240: \eab
1241: \label{PPdP}
1242: P(u)&=&\frac{\beta}{12\pi^2}u^2\bar{u}^2\,,\nonumber\\
1243: P^{\prime\prime}(u)&=&\frac{\beta}{\pi^2}\left[\frac{\pi}{\beta}\left(1-\frac{1}{\sqrt{3}}\right)-u\right]
1244: \left[\bar{u}-\frac{\pi}{\beta}\left(1-\frac{1}{\sqrt{3}}\right)\right]\,.
1245: \eae
1246: %**********
1247: The function $P(u)$ is always positive for $u\not=0,\frac{2\pi}{\beta}$.
1248: The occurrence of the spatial volume $V$ in the exponent
1249: in Eq.\,(\ref{Zcalonh}) would mean total suppression of nontrivial holonomy
1250: in the naive thermodynamical limit $V\to\infty$. This, however, is not a valid conclusion since
1251: nontrivial-holonomy calorons are unstable: They either dissociate into a pair of BPS
1252: monopoles (large holonomy) or they collapse back onto trivial holonomy by
1253: an annihilation of their BPS monopole constituents. This can be checked by investigating
1254: the second contribution to the exponent in Eq.\,(\ref{Zcalonh}).
1255:
1256: In Fig.\,\ref{Fig1c} a plot of $-\frac{\beta}{\pi}P^{\prime\prime}(\hat{u})$
1257: is shown. For $0\le u\le \frac{\pi}{\beta}(1-\frac{1}{\sqrt{3}})$ and
1258: for $\frac{\pi}{\beta}(1+\frac{1}{\sqrt{3}})\le u\le 2\,\frac{\pi}{\beta}$ the quantity
1259: $P^{\prime\prime}(u)$ is positive (small holonomy) while it is negative in the complementary range (large holonomy).
1260: According to Eq.\,(\ref{Zcalonh}) this means that in the former (latter)
1261: case the BPS monopoles experience a linear attractive (repulsive) potential.
1262: %***********************
1263: \begin{figure}
1264: \begin{center}
1265: \leavevmode
1266: %\epsfxsize=9.cm
1267: \leavevmode
1268: %\epsffile[80 25 534 344]{}
1269: \vspace{4.3cm}
1270: \special{psfile=Fig-5.ps angle=0 voffset=-120
1271: hoffset=-160 hscale=60 vscale=30}
1272: \end{center}
1273: \caption{The quantity $-\frac{\beta}{\pi}P^{\prime\prime}(\hat{u})$, compare with Eq.\,(\protect\ref{Zcalonh}),
1274: as a function of the dimensionless holonomy $\hat{u}\equiv\frac{u}{\pi T}$.\label{Fig1c}}
1275: \end{figure}
1276: %************************
1277: Let us now make an estimate of the typical size of an equilateral
1278: tetrahedron whose corners are the positions of
1279: (screened) magnetic monopoles, see Fig.\,\ref{Fig1d}, which are generated by
1280: the dissociation of a caloron and an anticaloron whose large holonomy
1281: was created by their interaction.
1282: %***********************
1283: \begin{figure}
1284: \begin{center}
1285: \leavevmode
1286: %\epsfxsize=9.cm
1287: \leavevmode
1288: %\epsffile[80 25 534 344]{}
1289: \vspace{4.3cm}
1290: \special{psfile=Fig-6.ps angle=0 voffset=-120
1291: hoffset=-105 hscale=60 vscale=50}
1292: \end{center}
1293: \caption{The typical volume spanned by two pairs of BPS monopoles created
1294: by the dissociation of two calorons whose large holonomy was generated
1295: by the interaction of trivial-holonomy calorons.\label{Fig1d}}
1296: \end{figure}
1297: %************************
1298: The edge length $R$ of the tetrahedron is the
1299: typical maximal distance between two BPS
1300: monopoles generated by a caloron with a
1301: holonomy close to maximally nontrivial,
1302: $u_{\tiny\mbox{max}}=\frac{\pi}{\beta}$. Once a large holonomy
1303: has been created the dissociation of the caloron generates the
1304: stabilized distance $R$ with probability one. (Once a monopole is at rest with
1305: respect to the heat bath there is no screening of its magnetic charge
1306: by Cartan fluctuations \cite{Linde1980} but only by small-holonomy calorons
1307: in its surroundings. Thus the screening of magnetic
1308: charge is not described by Eq.\,(\ref{Zcalonh}).) Thus it is
1309: appropriate to equate the probability for reaching the
1310: distance $R$, where monopoles are sufficiently screened to be at rest, governed
1311: by Eq.\,(\ref{Zcalonh}), with the thermal
1312: probability for exciting the monopoles in a caloron of large holonomy
1313: to start with. Since monopoles also are at rest shortly after being created the latter probability is
1314: roughly given as
1315: %*********
1316: \eqb
1317: \label{Boltzmann}
1318: \exp[-\beta(m_1+m_2)]\,,\ \ \ \ \ m_1\sim m_2\sim \frac{4\pi^2}{\beta}\,,
1319: \eqe
1320: %***********
1321: see Eq.\,(\ref{massesMonLL}). Taking only the exponentially sensitive part of the
1322: caloron weight into account and substituting for $V$ the volume of the
1323: tetrahedron, $V=\frac{1}{6\sqrt{2}}\,R^3$, this
1324: translates into the following condition:
1325: %*********
1326: \eqb
1327: \label{conddoms}
1328: -\frac{\pi^2}{72\sqrt{2}}\left(\frac{R}{\beta}\right)^3+\frac{2}{3}\pi\frac{R}{\beta}+8\pi^2=0\,.
1329: \eqe
1330: %*********
1331: There exists only a single real and positive solution to this equation. Numerically, we
1332: obtain $R\sim 10.1\,\beta$. So on the scale of
1333: the inverse temperature the gas of screened magnetic monopoles
1334: is dilute. This fits nicely with the lattice results
1335: obtained in \cite{KorthalsAltes,HoelbingRebbiRubakov2001}.
1336:
1337: While the (extremely small)
1338: likelihood for the generation
1339: of large-holonomy calorons depends on the
1340: value of the holonomy only (and not on the distance $D$) this is
1341: not true for a caloron with holonomy close to trivial. Since the latter configuration
1342: always collapses back onto trivial holonomy the likelihood for its generation
1343: is determined by the caloron weight $\exp[-S_{\tiny\mbox{eff}}]$ with $S_{\tiny\mbox{eff}}$
1344: given in Eq.\,(\ref{effactthcal}). A strong dependence of $S_{\tiny\mbox{eff}}$
1345: on $D$ (or $\rho$) at a given temperature exists. In contrast to $\exp[-\beta(m_1+m_2)]\sim \exp[-8\pi^2]$
1346: the weight $\exp[-S_{\tiny\mbox{eff}}]$ is sizable at
1347: $S_{\tiny\mbox{eff}}$'s minimum $\sigma_{\tiny\mbox{min}}$.
1348:
1349: We conclude that {\sl attraction} between a BPS monopole and its antimonopole
1350: (small holonomy), which are in equilibrium with respect of their
1351: creation and annihilation, by far dominates the ground-state physics as compared
1352: to the case where monopole and antimonopole {\sl repulse} one another
1353: (large holonomy). Macroscopically, this situation expresses itself by a {\sl negative} pressure
1354: of the ground state. We shall compute the temperature dependence of this
1355: pressure in Sec.\,\ref{pomod}.
1356:
1357:
1358: \subsubsection{Microscopic definition for the phase
1359: of an adjoint and macroscopic scalar field $\phi$\label{MDms}}
1360:
1361: The results that were discussed in the last two subsections are important
1362: for an understanding of the infrared physics
1363: in the electric phase. The detailed microscopic
1364: dynamics is very complicated and, as it seems, it
1365: is impossible to derive macroscopic quantities such as the pressure or the
1366: energy density or the mass of thermal quasiparticles
1367: by performing literal ensemble averages on the microscopic level.
1368: What turns out
1369: to be feasible and thermodynamically exhaustive
1370: is to compute the spatial average (spatial coarse-graining)
1371: over the physics generated by the
1372: topologically nontrivial sector. This procedure introduces the concept of a
1373: macroscopic, thermal ground state. As far as thermodynamics is
1374: concerned one still obtains
1375: exact results this way. The advantage of such an approach is that the
1376: complications of a microscopic calculation are avoided. Once the ground-state physics
1377: is understood and quantitatively described its effect on the propagation of
1378: trivial-topology modes can be investigated.
1379:
1380: If the ground state is to be characterized by a macroscopic field other than a pure-gauge
1381: configuration then, by spatial isotropy, this macroscopic field must be a Lorentz
1382: scalar $\phi$. Moreover, in a pure Yang-Mills theory, where all local fields
1383: transform under the adjoint representation of the gauge group, the composite
1384: field $\phi$ needs to transform in an adjoint way under the
1385: remnants of a microscopic, spacetime dependent gauge transformations. Since space
1386: dependent gauge transformations are constant on the macroscopic level
1387: (due to the spatial average) no space dependence of
1388: $\phi$ occurs in any chosen gauge. Apart from its modulus, which is governed
1389: by a dynamically emerging scale $\Lambda_E$ and temperature,
1390: the only nontrivial information on $\phi$ is the $\tau$ dependence
1391: of its color orientation in a given gauge. In the following we will refer
1392: to $\phi$'s color orientation as $\phi$'s phase.
1393:
1394: Let us imagine a (hypothetical) Yang-Mills world where the
1395: only field configurations allowed to contribute to the partition function
1396: are classical and noninteracting caloron configurations of trivial holonomy.
1397: We will show in Sec\,\ref{EC} that it is consistent
1398: to adopt this point of view in the derivation of the {\sl macroscopic}
1399: ground-state physics. Since the Yang-Mills
1400: scale $\Lambda_E$ can not be computed we focus on the computation of
1401: $\phi$'s phase first. Because $\phi$'s phase is a ratio of the field and
1402: its modulus and hence dimensionless the associated measure for the $\rho$-average is flat.
1403:
1404: We closely follow the presentation
1405: in \cite{HerbstHofmann2004} for the remainder of this section
1406: and for Sec.\,\ref{EC}. Due to the selfduality of calorons
1407: any local definition of $\phi's$ phase
1408: yields the trivial result zero. Thus we start by defining:
1409: %*********
1410: \eab
1411: \label{defphi}
1412: \frac{\phi^a}{|\phi|}(\tau)&\sim &\mbox{tr}\Bigg[\nonumber\\
1413: &&\beta^0 1!\int d^3x\,\int d\rho\, \nonumber\\
1414: &&\frac{\lambda^a}{2} F_{\mu\nu}[A_\alpha(\rho,\beta)]\left((\tau,0)\right)\,
1415: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha(\rho,\beta)]\times
1416: \nonumber\\
1417: &&F_{\mu\nu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{x})\right)\,
1418: \left\{(\tau,\vec{x}),(\tau,0)\right\}[A_\alpha(\rho,\beta)]+
1419: \nonumber\\
1420: &&\beta^{-1} 2!\int d^3x\int d^3y\,\int d\rho\,
1421: \nonumber\\
1422: &&\frac{\lambda^a}{2} F_{\mu\lambda}[A_\alpha(\rho,\beta)]\left((\tau,0)\right)\,
1423: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha(\rho,\beta)]\times
1424: \nonumber\\
1425: &&\,F_{\lambda\nu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{x})\right)\,
1426: \left\{(\tau,\vec{x}),(\tau,\vec{y})\right\}[A_\alpha(\rho,\beta)]\times
1427: \nonumber\\
1428: &&F_{\nu\mu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{y})\right) \left\{(\tau,\vec{y}),(\tau,0)\right\}
1429: [A_\alpha(\rho,\beta)]+
1430: \nonumber\\
1431: &&\beta^{-2} 3!\int d^3x\,\int d^3y\,\int d^3u\,\int d\rho\,
1432: \nonumber\\
1433: &&\frac{\lambda^a}{2} F_{\mu\lambda}[A_\alpha(\rho,\beta)]\left((\tau,0)\right)\,
1434: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha(\rho,\beta)]\times
1435: \nonumber\\
1436: &&\,F_{\lambda\nu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{x})\right)\,
1437: \left\{(\tau,\vec{x}),(\tau,\vec{y})\right\}[A_\alpha(\rho,\beta)]\times
1438: \nonumber\\
1439: &&F_{\nu\kappa}[A_\alpha(\rho,\beta)]\left((\tau,\vec{y})\right)
1440: \left\{(\tau,\vec{y}),(\tau,\vec{u})\right\}[A_\alpha(\rho,\beta)]
1441: F_{\kappa\mu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{u})\right)\times
1442: \nonumber\\
1443: &&\left\{(\tau,\vec{u}),(\tau,0)\right\}[A_\alpha(\rho,\beta)]+\cdots\Bigg]\,.
1444: \eae
1445: %*********
1446: A number of comments are in order:
1447: The dots in (\ref{defphi}) stand for the contributions of higher $n$-point functions and for reducible,
1448: that is, factorizable contributions with respect to the spatial integrations. The factors $(n-1)!$ in front of the $n$-point contribution measures
1449: the multiplicity of the corresponding integral. Factors $\beta^{n-2}$ are needed
1450: to make the contribution dimensionless. The argument $A_\alpha(\rho,\beta)$
1451: (spacetime dependence suppressed)
1452: refers to either a caloron or an anticaloron configuration, the Harrington-Shepard
1453: solutions of Sec.\,\ref{HSS}. Moreover, the following definitions apply:
1454: %*********
1455: \eab
1456: \label{defdefphi}
1457: |\phi|&\equiv&\frac{1}{2}\,\mbox{tr}\,\phi^2\,
1458: ,\nonumber\\
1459: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha]&\equiv& {\cal P}\,
1460: \exp\left[i\int_{(\tau,0)}^{(\tau,\vec{x})}dy_\beta\,A_\beta(y,\rho)\right]
1461: \,,\nonumber\\
1462: \left\{(\tau,\vec{x}),(\tau,0)\right\}[A_\alpha]&\equiv& {\cal P}\,
1463: \exp\left[-i\int_{(\tau,0)}^{(\tau,\vec{x})}dy_\beta\,A_\beta(y,\rho)\right]\,
1464: \eae
1465: %*********
1466: where ${\cal P}$ is the path-ordering symbol.
1467: \noindent Under a microscopic gauge transformation $\Omega(y)$ we have:
1468: %*********
1469: \eab
1470: \label{micrOm}
1471: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha]&\rightarrow & \Omega^\dagger ((\tau,0))\,
1472: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha]\,
1473: \Omega((\tau,\vec{x}))\,,\nonumber\\
1474: \left\{(\tau,\vec{x}),(\tau,0)\right\}[A_\alpha]&\rightarrow & \Omega^\dagger
1475: ((\tau,\vec{x}))\,\left\{(\tau,\vec{x}),(\tau,0)\right\}[A_\alpha]\,
1476: \Omega((\tau,0))\,,\nonumber\\
1477: F_{\mu\nu}[A_\alpha]\left((\tau,\vec{x})\right)&\rightarrow &
1478: \Omega^\dagger((\tau,\vec{x}))\,F_{\mu\nu}[A_\alpha]((\tau,\vec{x}))\,\Omega((\tau,\vec{x}))\,,\nonumber\\
1479: F_{\mu\nu}[A_\alpha]\left((\tau,0)\right)&\rightarrow &
1480: \Omega^\dagger((\tau,0))\,F_{\mu\nu}[A_\alpha]((\tau,0))\,\Omega((\tau,0))\,.
1481: \eae
1482: %*********
1483: As a consequence of Eq.\,(\ref{micrOm}) the right-hand side of (\ref{defphi}) transforms as
1484: %*********
1485: \eqb
1486: \label{phitrans}
1487: \frac{\phi^a}{|\phi|}(\tau)\rightarrow R_{ab}(\tau)\,\frac{\phi^b}{|\phi|}(\tau)
1488: \eqe
1489: %********
1490: where the SO(3) matrix $R_{ab}(\tau)$ is defined as
1491: %*******
1492: \eqb
1493: \label{Rdef}
1494: R^{ab}(\tau)\lambda^b=\Omega((\tau,0))\,\lambda^a\,\Omega^\dagger((\tau,0))\,.
1495: \eqe
1496: %******
1497: Thus we have defined an adjointly transforming
1498: scalar in (\ref{defphi}). In addition, we have just
1499: shown that only the time-dependent part of a microscopic gauge transformation
1500: survives on the macroscopic level. (Shifting the spatial part of the argument
1501: $(\tau,0)\to(\tau,\vec{z})$ in (\ref{defphi})
1502: introduces a finite {\sl parameter} $\vec{z}$ to the gauge rotation
1503: $R^{ab}$: $R^{ab}(\tau)\to R^{ab}(\tau,\vec{z})$. Such a shift, however,
1504: introduces an arbitrary but finite mass scale $|\vec{z}|^{-1}$
1505: into the definition of $\phi$'s phase
1506: which, on the classical level, is absent. Also, a finite value of $|\vec{z}|$ would introduce an explicit breaking
1507: or rotational symmetry into the definition (\ref{defphi}). Thus we have $\vec{z}=0$. Moreover, the integration
1508: path connecting the points $(\tau,0)$ with $(\tau,\vec{x})$
1509: in Eq.\,(\ref{defdefphi}) ought to be a straight line since a spatial
1510: curvature would imply the existence of a mass scale other than
1511: temperature on the classical level.) Integrations over shifts $\tau\to\tau+\tau_s$ ($0\le\tau_s\le\beta$) project a nontrivial (periodic)
1512: $\tau$ dependence of $\phi$'s phase onto zero and thus are forbidden. Integrations
1513: over global gauge rotations are forbidden for the same reason.
1514: Spatial shifts $\vec{x}\to\vec{x}+\vec{x}_s,\vec{y}\to\vec{y}+\vec{x}_s,\cdots$
1515: leave the integrals in (\ref{defphi}) invariant. These averages are
1516: already performed. Thus the only admissible integration over moduli-space parameters is over
1517: $\rho$ with a flat measure.
1518:
1519: In (\ref{defphi}) the $\sim$ sign indicates that both left- and right-hand sides satisfy the same
1520: homogeneous evolution equation in $\tau$
1521: %**********
1522: \eqb
1523: \label{deffequationhomo}
1524: {\cal D}\left[\frac{\phi}{|\phi|}\right]=0\,.
1525: \eqe
1526: %*********
1527: Here ${\cal D}$ is
1528: a differential operator such that Eq.\,(\ref{deffequationhomo}) represents a
1529: homogeneous differential equation.
1530: As it will turn out, Eq.\,(\ref{deffequationhomo}) is a {\sl linear}
1531: second-order equation which, up to global gauge rotations and a choice of winding sense,
1532: determines the first-order or BPS equation whose
1533: solution $\phi$'s phase is. (The ambiguities in the evaluation
1534: of the right-hand side span the solution space of Eq.\,(\ref{deffequationhomo}), and thus ${\cal D}$
1535: is uniquely determined by (\ref{defphi}).)
1536:
1537: We now discuss why $n$-point functions with $n>2$ do not contribute to the
1538: right-hand side of (\ref{defphi}). Since the classical (anti)caloron action
1539: $S=\frac{8\pi^2}{g^2}$ and the classical moduli-space metric are
1540: independent of temperature we conclude that
1541: no {\sl explicit} dependence on $\beta$ may occur in the definition of $\phi$'s phase.
1542: For $n>2$, however, explicit $\beta$ dependences do occur, see (\ref{defphi}). We conclude that
1543: these contributions to $\phi$'s phase do not exist.
1544:
1545: What about calorons of higher topological charge? Some of these solutions have been
1546: constructed, see for example \cite{Actor1983,Chakrabarti1987}.
1547: The essential difference to the charge-one case is that
1548: more dimensionful moduli occur than just the parameter
1549: $\rho$. For example, for charge-two configurations
1550: there is a spatial core separation between the two seed
1551: instantons and an additional instanton
1552: radius $\rho^\prime$. The reader may now
1553: convince himself that along the lines of (\ref{defphi})
1554: a nonlocal definition of $\phi$'s {\sl dimensionless} phase, which would also
1555: have to include integrations over the additional moduli of dimension
1556: length, is impossible for higher-charge calorons. (This is certainly
1557: true for the integral over two-point functions. For every increment in
1558: $n$ there is an increase in power of length scale by one unit arising from an additional
1559: $d^3x\times F_{\mu\nu}$. This makes the situation even worse in comparison
1560: to the two-point case.)
1561:
1562:
1563: \subsubsection{Essentials of the calculation\label{EC}}
1564:
1565: Before we dive into the essential parts of the
1566: calculation, which will lead to the unique determination
1567: of the operator ${\cal D}$ in
1568: Eq.\,(\ref{deffequationhomo}), we would like
1569: to discuss a condition which severely constrains
1570: the possible solutions to this equation.
1571:
1572: By (anti)selfduality the energy-momentum tensor vanishes
1573: identically on a caloron or an
1574: anticaloron,
1575: %********
1576: \eqb
1577: \label{thetacal}
1578: \theta_{\mu\nu}[A^{(C,A)}_\alpha]\equiv 0\,.
1579: \eqe
1580: %*********
1581: Since $\phi$'s phase is
1582: obtained by an average over (the admissible part of)
1583: the moduli space of a caloron-anticaloron system (no interactions) the macroscopic energy-momentum tensor
1584: $\bar{\theta}_{\mu\nu}[\phi]$ should vanish identically as well,
1585: %********
1586: \eqb
1587: \label{thetabar}
1588: \bar{\theta}_{\mu\nu}[\phi]\equiv 0\,.
1589: \eqe
1590: %*********
1591: In a thermal equilibrium situation, described by Euclidean dynamics,
1592: this is true if and only if the $\tau$ dependence of $\phi$ (or $\phi$'s phase)
1593: is BPS saturated. Thus $\phi$ solves the first-order equation
1594: %********
1595: \eqb
1596: \label{BPSphi}
1597: \pd_\tau\phi=V_E^{(1/2)}\,
1598: \eqe
1599: %*********
1600: where $V_E^{(1/2)}$ denotes the 'square-root' of a suitable
1601: potential. (The fact that an ordinary and not a
1602: covariant derivative appears in Eq.\,(\ref{BPSphi}) is, of course, tied to our specific gauge
1603: choice. If we were to leave the (singular)
1604: gauge for the seed (anti)instanton, in which the solutions of Eq.\,(\ref{HS}) are constructed,
1605: by a time-dependent gauge rotation $\bar{\Omega}
1606: (\tau)$ then a pure-gauge configuration
1607: $A^{p.g.}_\mu(\tau)=i\delta_{\mu4}\bar{\Omega}^\dagger\pd_{\tau}\bar{\Omega}$
1608: would appear in a {\sl covariant} derivative on the left-hand side
1609: of Eq.\,(\ref{BPSphi}). Also recall the fact
1610: that the heat bath breaks boost invariance.
1611: This is encoded in the noninvariance of Eq.\,(\ref{BPSphi})
1612: under O(4) rotations.) $V_E\equiv\mbox{tr}\,\left(V_E^{(1/2)}\right)^\dagger\,V_E^{(1/2)}$.
1613: As we will see below, the right-hand side of Eq.\,(\ref{BPSphi}) is
1614: determined only up to a global gauge rotation and a choice of winding sense.
1615:
1616: Let us now discuss essential details of the
1617: calculation of the right-hand side of (\ref{defphi}) which,
1618: after what was said in Sec.\,\ref{MDms}, reduces to
1619: %******
1620: \eab
1621: \label{reddefphi}
1622: \frac{\phi^a}{|\phi|}(\tau)&\sim &\mbox{tr}\int d^3x\,\int d\rho\, \frac{\lambda^a}{2}\,F_{\mu\nu}[A_\alpha(\rho,\beta)]\left((\tau,0)\right)\,
1623: \left\{(\tau,0),(\tau,\vec{x})\right\}[A_\alpha(\rho,\beta)]\times\nonumber\\
1624: &&F_{\mu\nu}[A_\alpha(\rho,\beta)]\left((\tau,\vec{x})\right)\,
1625: \left\{(\tau,\vec{x}),(\tau,0)\right\}[A_\alpha(\rho,\beta)]\,
1626: \eae
1627: %*******
1628: where a sum over the contributions of a trivial-holonomy caloron and an anticalorons
1629: is to be performed.
1630:
1631: Since the integrand in the exponent of the Wilson line $\left\{(\tau,0),(\tau,\vec{x})\right\}
1632: [A_\alpha^{C,A}]$ is a hedgehog the path-ordering
1633: prescription can be omitted. For the caloron contribution one obtains
1634: %***********
1635: \eab
1636: \label{3DUP}
1637: \left.\frac{\phi^a}{|\phi|}\right|_{\tiny{C}}&\sim&
1638: i\int d\rho \int d^3x\,\frac{x^a}{r}\times\nonumber\\
1639: &&\left[\frac{\left(\partial_4\Pi(\tau+\tau_C,0)\right)^2}{\Pi^2(\tau+\tau_C,0)}-\frac23
1640: \frac{\partial^2_4\Pi(\tau+\tau_C,0)}{\Pi(\tau+\tau_C,0)}\right]\Big\{4\cos(2g(\tau+\tau_C,r))\times
1641: \nonumber\\
1642: &&\left.\left[\frac{\pd_r\pd_4\Pi(\tau+\tau_C,r)}{\Pi(\tau+\tau_C,r)}-
1643: 2\frac{\left(\pd_r\Pi(\tau+\tau_C,r)\right)\left(\pd_4\Pi(\tau+\tau_C,r)\right)}
1644: {\Pi^2(\tau+\tau_C,r)}\right]+\right.\nonumber\\
1645: &&\left.\sin(2g(\tau+\tau_C,r))\left[4\,\frac{\left(\pd_4\Pi(\tau+\tau_C,r)\right)^2-
1646: \left(\pd_r\Pi(\tau+\tau_C,r)\right)^2}{\Pi^2(\tau+\tau_C,r)}+\right.\right.\nonumber\\
1647: &&\left.\left. 2\,\frac{\pd^2_r\Pi(\tau+\tau_C,r)-\pd^2_4\Pi(\tau+\tau_C,r)}{\Pi(\tau+\tau_C,r)}
1648: \right]\right\}\,
1649: \eae
1650: %**********
1651: where
1652: %********
1653: \eqb
1654: \label{grtau}
1655: g(\tau+\tau_C,r)\equiv \int_0^1 ds\, \frac{r}{2} \, \pd_4 \ln \Pi(\tau+\tau_C,sr)\,,
1656: \eqe
1657: %********
1658: the function $\Pi(\tau,r)$ is defined in Eq.\,(\ref{Pi}), and $\tau_C$ refers
1659: to a constant but arbitrary temporal shift of the caloron center ($0\le\tau_C\le\beta$). It is worth mentioning
1660: that the integrand in Eq.\,(\ref{grtau}) is proportional to $\delta(s)$
1661: for $r\gg\beta$. The dependences on $\rho$ and $\beta$ are suppressed in the
1662: integrands of (\ref{3DUP}) and Eq.\,(\ref{grtau}).
1663:
1664: As compared to the
1665: contribution of the caloron there are ambiguities in the contribution
1666: of the anticaloron: First, the $\tau$ dependence of the contribution of the anticaloron may be shifted by $\tau_A$
1667: ($0\le\tau_A\le\beta$). Second, the color orientation of caloron and anticaloron
1668: contributions may be different. Third, the normalization of the two
1669: contributions may be different.
1670: To see that this is true, we need to investigate the convergence properties of the radial
1671: integration in (\ref{3DUP}). It is easily checked that
1672: all terms give rise to a converging $r$ integration
1673: except for the following one:
1674: %***********
1675: \eqb
1676: \label{nonconvr}
1677: 2\,\frac{x^a}{r}\,\sin(2g(\tau+\tau_C,r))\,\frac{\pd^2_r\Pi(\tau+\tau_C,r)}{\Pi(\tau+\tau_C,r)}\,.
1678: \eqe
1679: %***********
1680: Namely, for $r>R\gg\beta$ (\ref{nonconvr}) goes over in
1681: %***********
1682: \eqb
1683: \label{nonconvrlarger}
1684: 4\,\frac{x^a}{r}\frac{\pi\rho^2\sin(2g(\tau+\tau_C,r))}{\beta r^3}\,.
1685: \eqe
1686: %***********
1687: Thus the $r$-integral of the term in (\ref{nonconvr}) is
1688: logarithmically divergent in the infrared: (The integral converges for $r\to 0$.)
1689: %***********
1690: \eqb
1691: \label{nocI}
1692: 4\,\frac{\pi\rho^2}{\beta}\int_R^\infty\frac{dr}{r}\frac{x^a}{r}\,\sin(2g(\tau+\tau_C,r))\,\,.
1693: \eqe
1694: %***********
1695: Recall, that $g(\tau+\tau_C,r)$ behaves like a constant
1696: in $r$ for $r>R$. The angular
1697: integration, on the other hand, would
1698: yield zero if the radial integration was regular.
1699: Thus a logarithmic divergence can be cancelled
1700: by the angular integral to yield some
1701: finite and real answer. To investigate this in more detail, both angular and radial
1702: integration need to regularized.
1703:
1704: One introduces a regularization, conveniently we have chosen
1705: dimensional regularization in \cite{HerbstHofmann2004}
1706: with a dimensionless regularization parameter $\eta_C>0$,
1707: for the $r$-integral
1708: in Eq.\,(\ref{nocI}) while the angular integration can be
1709: regularized by introducing defect (or surplus) angles
1710: $2\eta^\prime_C$ in the $\theta$ integration
1711: (azimuthal angle in the $x_1x_2$ plane). Any other plane for the azimuthal
1712: angular integration could have been chosen. Moreover, the value of $\alpha_C$ is
1713: determined by a (physically irrelevant)
1714: initial condition, as we will show below,
1715: see Fig.\,\ref{Fig1e}. Together, the choice of the
1716: regularization plane and of the angle $\alpha_C$ amount
1717: to a global choice of gauge: an apparent breaking of rotational symmetry by the
1718: angular regularization is nothing but a gauge choice.)
1719: %***********************
1720: \begin{figure}
1721: \begin{center}
1722: \leavevmode
1723: %\epsfxsize=9.cm
1724: \leavevmode
1725: %\epsffile[80 25 534 344]{}
1726: \vspace{4.8cm}
1727: \special{psfile=Fig-7.ps angle=0 voffset=-150
1728: hoffset=-105 hscale=60 vscale=50}
1729: \end{center}
1730: \caption{The axis for the regularized azimuthal integration.\label{Fig1e}}
1731: \end{figure}
1732: %************************
1733: Without restriction of generality (global gauge choice)
1734: we may also for the contribution of the anticaloron
1735: use an axis for the angular regularization which lies in the
1736: $x_1x_2$ plane, but with a different angle $\alpha_A$. Then we have
1737: %*********
1738: \eab
1739: \label{totalc}
1740: \frac{\phi^a}{|\phi|}
1741: &=&\left.\frac{\phi^a}{|\phi|}\right|_{\tiny{C}}+
1742: \left.\frac{\phi^a}{|\phi|}\right|_{\tiny{A}}\nonumber\\
1743: &=&
1744: \pm\Xi_C\,\left(\delta_{a1}\cos\alpha_C+\delta_{a2}\sin\alpha_C\right) {\cal A}
1745: \left(\frac{2\pi(\tau+\tau_C)}{\beta}\right)
1746: \nonumber\\
1747: &&\pm\Xi_A\,\left(\delta_{a1}\cos\alpha_A+\delta_{a2}\sin\alpha_A\right) {\cal A}
1748: \left(\frac{2\pi(\tau+\tau_A)}{\beta}\right)
1749: \nonumber\\
1750: &\neq&0\,,
1751: \eae
1752: %*********
1753: where $\Xi_C$ , $\Xi_A$, $\tau_C$, $\tau_A$, $\alpha_C$, and $\alpha_A$
1754: are undetermined. ($\Xi_C$, $\Xi_A$ are the ratios of $\eta_{C,A}$ and $\eta^\prime_{C,A}$, respectively.)
1755: The function ${\cal A}\left(\frac{2\pi(\tau)}{\beta}\right)$
1756: is defined as
1757: %*********
1758: \eab
1759: \label{calAdef}
1760: {\cal A}
1761: \left(\frac{2\pi\tau}{\beta}\right)&\equiv&
1762: \frac{32}{3}\,\frac{\pi^7}{\beta^3}\,
1763: \int d\rho\,\left[\lim_{r\to\infty}\sin(2g(\tau,r))\right]\times\nonumber\\
1764: &&\rho^4\frac{\pi^2 \rho^2 + \beta^2 \left(2 + \cos \left(\frac{2\pi\tau}{\beta}\right)\right)}
1765: {\left[2\pi^2 \rho^2 + \beta^2 \left(1-\cos \left(\frac{2\pi\tau}{\beta}\right)\right)\right]^2}\,.
1766: \eae
1767: %**********
1768: Eq.\,(\ref{totalc}) and Eq.\,(\ref{calAdef}) provide the basis for
1769: fixing the operator $\cal D$ in Eq.\,(\ref{deffequationhomo}).
1770: To evaluate the function ${\cal A}\left(\frac{2\pi\tau}{\beta}\right)$
1771: in Eq.\,(\ref{calAdef}) numerically, we introduce the same cutoff for the $\rho$ integration
1772: in the caloron and anticaloron case
1773: as follows:
1774: %********
1775: \eqb
1776: \label{cutoffrho}
1777: \int d\rho\to \int_0^{\zeta \beta} d\rho\,,\ \ \ \ \ \ (\zeta>0)\,.
1778: \eqe
1779: %*********
1780: This introduces an additional dependence of ${\cal A}$ on $\zeta$.
1781: In Fig.\,\ref{Fig1f} the $\tau$ dependence of ${\cal A}$ for
1782: various values of $\zeta$ is depicted.
1783: %***********************
1784: \begin{figure}
1785: \begin{center}
1786: %\leavevmode
1787: %%\epsfxsize=9.cm
1788: %\leavevmode
1789: %%\epsffile[80 25 534 344]{}
1790: \vspace{5.3cm}
1791: \special{psfile=Fig-8.ps angle=0 voffset=-10
1792: hoffset=-20 hscale=50 vscale=70}
1793: \end{center}
1794: \caption{${\cal A}$ as a function of $\frac{2\pi}{\beta}\tau$ for $\zeta=1,2,10$. For each case
1795: the dashed line is a plot of $\mbox{max}\,{\cal A}\times\sin\left(\frac{2\pi}{\beta}\tau\right)$.
1796: We have fitted the asymptotic dependence on $\zeta$ of the amplitude of ${\cal A}$ as
1797: ${\cal A}\left(\frac{2\pi}{\beta}\tau=\frac{\pi}{2},\zeta\right)=272\,\zeta^3,\,(\zeta>10)$.
1798: The fit is stable under variations of the
1799: fitting interval. For the case $\zeta=10$ the difference
1800: between the two curves can not be resolved anymore.\label{Fig1f}}
1801: \end{figure}
1802: %************************
1803: Therefore we have
1804: %*********
1805: \eab
1806: \label{finalres}
1807: \frac{\phi^a}{|\phi|} &\sim& 272\,\zeta^3\,
1808: \Bigg(
1809: \Xi_C
1810: \left(\delta_{a1}\cos\alpha_C+\delta_{a2}\sin\alpha_C\right)\,
1811: \sin\left(\frac{2\pi}{\beta}(\tau+\tau_C)\right)
1812: \nonumber \\ &&
1813: +
1814: \Xi_A
1815: \left(\delta_{a1}\cos\alpha_A+\delta_{a2}\sin\alpha_A\right)\,
1816: \sin\left(\frac{2\pi}{\beta}(\tau+\tau_A)\right)
1817: \Bigg)
1818: \nonumber\\
1819: &\equiv& \hat \phi^a
1820: \,.
1821: \eae
1822: %**********
1823: Just like the numbers $\Xi_C$ and $\Xi_A$ are
1824: undetermined on the classical level due to the invariance of the classical action under
1825: spatial scale transformations so is the number $\zeta$. It is clear, however,
1826: from Eq.\,(\ref{finalres}) that the integral in ${\cal A}$
1827: is strongly dominated by $\rho$-values close to the upper integration limit.
1828: Let us now discuss the physical content of (\ref{finalres}). For fixed values of
1829: the parameters $\zeta^3\,\Xi_C$, $\zeta^3\,\Xi_A$, $\frac{\tau_C}{\beta}$ and $\frac{\tau_A}{\beta}$
1830: the right-hand side of Eq.\,(\ref{finalres}) resembles an elliptic polarization
1831: in the $x_1x_2$ plane of adjoint color space.
1832: For a given polarization plane the two independent numbers (normalization and phase-shift)
1833: for each of the two oscillations parametrize the solution space of the second-order linear differential equation
1834: %************
1835: \eqb\label{2order}
1836: {\cal D} \hat\phi=0\,.
1837: \eqe
1838: From (\ref{finalres}) we observe that the operator $\cal D$ is
1839: \eqb
1840: {\cal D} = \partial_\tau^2 + \left( \frac{2\pi}{\beta} \right)^2\,.
1841: \eqe
1842: %*************
1843: The ambiguities in Eq.\,(\ref{finalres}) parameterize the
1844: solution space of Eq.\,(\ref{2order}) for a
1845: given polarization plane which depends on a global choice of gauge. Thus the
1846: differential operator ${\cal D}$ is {\sl uniquely}
1847: determined by Eq.\,(\ref{finalres}). What is needed to assure the
1848: validity of Eq.\,(\ref{thetabar}) is a BPS
1849: saturation of the solution to the linear Eq.\,(\ref{2order}) since
1850: the modulus of $\phi$ may not depend on $\tau$ in thermal equilibrium.
1851:
1852: Thus we need to find first-order equations whose solutions are traceless, hermitian and
1853: solve the second-order equation (\ref{2order}).
1854: The relevant two first-order equations are
1855: %***********
1856: \eqb\label{1orderbps}
1857: \pd_\tau \hat{\phi}=\pm\frac{2\pi i}{\beta}\lambda_3\,\hat{\phi}\,.
1858: \eqe
1859: %**********
1860: Obviously, the right-hand sides of Eqs.\,(\ref{1orderbps}) are subject to a
1861: global gauge ambiguity (associated with the choice of polarization
1862: plane in which the regularization of the azimuthal angular
1863: integration is carried out) and a choice of sign: Any normalized generator other
1864: than $\pm\lambda_3$ could have appeared. Moreover, a solution to either of the two equations (\ref{1orderbps})
1865: also solves Eq.\,(\ref{2order}) for a given polarization plane,
1866: %************
1867: \eqb
1868: \partial_\tau^2 \hat\phi=\pm \frac{2\pi i}{\beta}\lambda_3\, \partial_\tau \hat \phi=
1869: \frac{2\pi i}{\beta}\lambda_3\,\frac{2\pi i}{\beta}\lambda_3\,\hat\phi=-
1870: \left( \frac{2\pi}{\beta} \right)^2\hat\phi\,.
1871: \eqe
1872: %************
1873: Solutions to Eqs.\,(\ref{1orderbps}) are given as
1874: %***********
1875: \eqb\label{solutiona}
1876: \hat{\phi} = C \, \lambda_1 \, \exp\left(\mp \frac{2\pi i}{\beta} \lambda_3 (\tau-\tau_0) \right)
1877: \eqe
1878: %************
1879: where $C$ and $\tau_0$ denote real integration constants which
1880: both are undetermined. We set $\tau_0=0$ in what follows.
1881: The solutions in Eq.\,(\ref{solutiona}) represent a circular polarization
1882: in the $x_1x_2$ plane of adjoint color space
1883: and thus indicate that the field $\phi$ winds along an $S_1$
1884: on the group manifold $S_3$ of SU(2). Both winding senses appear but
1885: can not be distinguished physically: A change in winding
1886: sense does not affect the potential nor does it affect
1887: the admissibility of the transformation to
1888: unitary gauge, see Sec.\,\ref{gtug}.
1889:
1890: \subsubsection{$\phi$'s modulus and potential\label{pomod}}
1891:
1892: \noindent\underline{SU(2) case:}
1893: \vspace{0.1cm}\\
1894: The information in Eq.\,(\ref{solutiona}) on $\phi$'s
1895: phase can be used to infer its modulus once the
1896: existence of an externally given mass scale $\Lambda_E$ is
1897: assumed. (The scale $\Lambda_E$ determines the typical distance between caloron centers at a
1898: given temperature.) As long as no interactions between trivial-holonomy calorons
1899: are allowed for this is consistent since the BPS saturation of $\phi$
1900: forbids the occurrence of (gravitationally)
1901: measurable effects: The macroscopic energy-momentum
1902: tensor $\bar{\theta}_{\mu\nu}$ vanishes identically, and thus assuming the existence of the
1903: scale $\Lambda_E$ does not yet influence the ground-state physics. We have
1904: %************
1905: \eqb
1906: \label{tafel}
1907: \phi = |\phi|(\beta,\Lambda_E)\,\hat{\phi}\left(\frac{\tau}{\beta}\right)\,.
1908: \eqe
1909: %********
1910: In order to reproduce the phase in Eq.\,(\ref{solutiona}) a {\sl linear} dependence on $\phi$
1911: must appear on the right-hand side of the BPS equation (\ref{BPSphi}). Moreover,
1912: this right-hand side ought not depend on $\beta$ explicitly and must be
1913: analytic in $\phi$. The former requirement derives from the fact that $\phi$ and its potential $V$ are
1914: obtained by functionally integrating over the (admissible part of the) moduli space of a
1915: caloron-anticaloron system with no interactions. The associated part of the partition function
1916: does not exhibit an explicit $\beta$ dependence since
1917: the action and thus the weight are $\beta$ independent on the moduli space.
1918: Thus a $\beta$ dependence of $V$ or $V^{(1/2)}$ can only be generated
1919: via the periodicity of $\phi$ itself. The latter requirement
1920: derives from the demand that the thermodynamics at temperature $T + \delta T$ to any given accuracy
1921: must be smoothly derivable from the thermodynamics at temperature $T$ for $\delta T$ sufficiently
1922: small provided no phase transition occurs at $T$. This is done
1923: by expanding the right-hand side of the
1924: BPS equation (finite radius of convergence)
1925: which, in turn, is the starting point for a
1926: perturbative treatment with expansion parameter $\frac{\delta T}{T}$.
1927:
1928: Linearity, analyticity, and no explicit
1929: dependence of $\beta$ only allow the BPS equation for $\phi$ to be one
1930: the two following possibilities:
1931: %*******
1932: \eqb\label{bps13}
1933: \pd_\tau\phi=\pm i\,\Lambda_E\,\lambda_3\,\phi
1934: \eqe
1935: %**********
1936: or
1937: %*******
1938: \eqb\label{bps14}
1939: \pd_\tau\phi=\pm i\,\Lambda_E^3\,\lambda_3\,\phi^{-1}
1940: \eqe
1941: %**********
1942: where $\phi^{-1}\equiv \frac{\phi}{|\phi|^2}$. Recall that
1943: %***********
1944: \eqb
1945: \phi^{-1} = \phi_0^{-1} \sum_{n=0}^{\infty} (-1)^n \phi_0^{-n} \left(\phi-\phi_0\right)^n
1946: \eqe
1947: %*********
1948: has a finite radius of convergence. According to Eqs.\,(\ref{tafel}) and
1949: (\ref{solutiona}) we may write
1950: %*********
1951: \eqb
1952: \label{ansatzBPS}
1953: \phi = |\phi|(\beta,\Lambda_E)\, \times\,\lambda_1\, \exp\left( \mp \frac{2\pi i}{\beta} \lambda_3 \tau\right) \,.
1954: \eqe
1955: %***********
1956: Substituting Eq.\,(\ref{ansatzBPS}) into
1957: Eq.\,(\ref{bps13}) yields
1958: %*******
1959: \eqb
1960: \label{contraBPS}
1961: \Lambda_E=\frac{2\pi}{\beta}
1962: \eqe
1963: %********
1964: which is unacceptable since $\Lambda_E$ is a constant scale. For the other possibility
1965: Eq.\,(\ref{bps14}), we obtain
1966: %*******
1967: \eqb
1968: \label{nocontraBPS}
1969: |\phi|(\beta,\Lambda_E)=\sqrt{\frac{\beta\Lambda_E^3}{2\pi}}=\sqrt{\frac{\Lambda_E^3}{2\pi\,T}}\,
1970: \eqe
1971: %********
1972: when substituting Eq.\,(\ref{ansatzBPS}) into Eq.\,(\ref{bps14}).
1973: This is acceptable and indicates
1974: that at $T\gg \Lambda_E$ $\phi$'s modulus is small.
1975: The right-hand side of Eq.\,(\ref{bps14}) defines the 'square-root' $V^{(1/2)}$ of a potential
1976: $V(|\phi|)\equiv\mbox{tr}\,\left(V^{(1/2)}\right)^\dagger\,V^{(1/2)}=\Lambda_E^6 \, \mbox{tr} \, \phi^{-2}$.
1977: The equation of motion (\ref{bps14})
1978: can be derived from the following action:
1979: \eqb \label{actionphi}
1980: S_{\phi} = \mbox{tr} \, \int_0^\beta d\tau \int d^3x
1981: \left( \partial_\tau \phi \partial_\tau \phi + \Lambda_E^6 \phi^{-2} \right) \,.
1982: \eqe
1983: Notice that due to BPS saturation it is not possible to
1984: add a constant to the potential in Eq.\,(\ref{actionphi}) without
1985: changing the ground-state physics. (In fact, adding a constant, the modified
1986: BPS equation would not admit periodic solutions anymore.) The scale $|\phi|$ must be
1987: interpreted as the maximal resolution that remains after the spatial coarse-graining over
1988: calorons and anticalorons is performed. As we shall show later, a
1989: critical temperature $2\pi T_{c,E}=13.867\,\Lambda_E$ exists.
1990: Thus, expressing the critical cutoff $|\phi|^{-1}=\sqrt{\frac{2\pi}{\Lambda_E^3\beta_{c,E}}}$ in units of $\beta_{c,E}$, yields
1991: $8.22$; for $T>T_{c,E}$ this number grows as $(T/T_{c,E})^{3/2}$. But cutting off the $\rho$- and $r$-integration at $>8.22\,\beta$ perfectly
1992: represents the infinite-volume limit in Eq.\,(\ref{calAdef})!
1993:
1994: The ratios of the mass-squared of $\phi$-field fluctuations, $\pd^2_{|\phi|}\,V(|\phi|)$, and
1995: the compositeness scale $|\phi|$ squared or $T^2$ are given as
1996: %*********
1997: \eqb
1998: \label{ratiosCC}
1999: \frac{\pd^2_{|\phi|}V_E}{|\phi|^2}=12\,\lambda_E^3\,,\ \ \ \ \ \ \ \
2000: \frac{\pd^2_{|\tilde{\phi}_l|}V_E}{T^2}=48\,\pi^2\,\,,
2001: \,
2002: \eqe
2003: %*********
2004: where $\lambda_E\equiv\frac{2\pi T}{\La_E}$. We will show in
2005: Sec.\,\ref{eveffgc} that $\lambda_E\ge 13.867$ in the
2006: electric phase. Thus both ratios in
2007: Eq.\,(\ref{ratiosCC}) are much larger than unity: The field
2008: $\phi$ is quantum mechanically and statistically inert.
2009: It represents a {\sl background} for the dynamics of
2010: the topologically trivial sector after spatial coarse-graining. As a consequence, our
2011: assumption that only noninteracting calorons of trivial
2012: holonomy contribute to the average in Eq.\,(\ref{reddefphi})
2013: is consistent.\vspace{0.1cm}\\
2014: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2015: For SU(3) we write three sets of SU(2) generators as
2016: %*********
2017: \eqb
2018: \label{lanot}
2019: {\lambda}_1=\left(\begin{array}{ccc}0&1&0\\
2020: 1&0&0\\
2021: 0&0&0\end{array}\right)\,,\ \ \
2022: {\lambda}_2=\left(\begin{array}{ccc}0&-i&0\\
2023: i&0&0\\
2024: 0&0&0\end{array}\right)\,,\ \ \
2025: {\lambda}_3=\left(\begin{array}{ccc}1&0&0\\
2026: 0&-1&0\\
2027: 0&0&0\end{array}\right)\,,
2028: \eqe
2029: %*********
2030: and
2031: %*********
2032: \eqb
2033: \label{barla}
2034: \bar{\lambda}_1=\left(\begin{array}{ccc}0&0&1\\
2035: 0&0&0\\
2036: 1&0&0\end{array}\right)\,,\ \ \
2037: \bar{\lambda}_2=\left(\begin{array}{ccc}0&0&-i\\
2038: 0&0&0\\
2039: i&0&0\end{array}\right)\,,\ \ \
2040: \bar{\lambda}_3=\left(\begin{array}{ccc}1&0&0\\
2041: 0&0&0\\
2042: 0&0&-1\end{array}\right)\,,
2043: \eqe
2044: %*********
2045: and
2046: %*********
2047: \eqb
2048: \label{tildela}
2049: \tilde{\lambda}_1=\left(\begin{array}{ccc}0&0&0\\
2050: 0&0&1\\
2051: 0&1&0\end{array}\right)\,,\ \ \
2052: \tilde{\lambda}_2=\left(\begin{array}{ccc}0&0&0\\
2053: 0&0&-i\\
2054: 0&i&0\end{array}\right)\,,\ \ \
2055: \tilde{\lambda}_3=\left(\begin{array}{ccc}0&0&0\\
2056: 0&1&0\\
2057: 0&0&-1\end{array}\right)\,.
2058: \eqe
2059: %*********
2060: One generator is dependent. This just reflects the fact that the group manifold
2061: of SU(3) locally is not $S_3\times S_3\times S_3$ but $S_3\times S_5$
2062: \cite{Aguilar1999,Steenrod1951}. A set of independent generators is obtained by
2063: replacing the two matrices
2064: $\bar{\lambda}_3$ and $\tilde{\lambda}_3$ by the single matrix
2065: %********
2066: \eqb
2067: \label{GellMann}
2068: \lambda_8=\frac{1}{\sqrt{3}}\left(\bar{\lambda}_3+\tilde{\lambda}_3\right)=
2069: \frac{1}{\sqrt{3}}\left(\begin{array}{ccc}1&0&0\\
2070: 0&1&0\\
2071: 0&0&-2\end{array}\right)\,
2072: \eqe
2073: %******
2074: and by keeping the other matrices. The result is the familiar set
2075: of Gell-Mann matrices generating the group SU(3).
2076:
2077: For the case of SU(3) the field $\phi$ may
2078: wind in each of the above SU(2) algebras. Except for the
2079: points $\tau=0,\frac{\beta}{3},\frac{2\beta}{3}$, where it jumps into a
2080: new algebra, a solution to the
2081: BPS equation
2082: %********
2083: \eqb
2084: \label{BPSSU(3)}
2085: \pd_\tau\phi=\pm i\,\Lambda_E^3\,\left\{\begin{array}{c}\lambda_3\,\frac{\phi}{|\phi|^2}\,,
2086: \ \ \ \ (0\le\tau<\frac{\beta}{3})\nonumber\\
2087: \bar{\lambda}_3\,\frac{\phi}{|\phi|^2}\,,\ \ \ \ (\frac{\beta}{3}\le\tau<\frac{2\beta}{3})\nonumber\\
2088: \tilde{\lambda}_3\,\frac{\phi}{|\phi|^2}\,,\ \ \ \ (\frac{2\beta}{3}\le\tau<\beta)\,\end{array}\right.
2089: \eqe
2090: %**********
2091: is given as
2092: %********
2093: \eqb
2094: \label{solBPSSU(3)}
2095: \phi(\tau)=\sqrt{\frac{\La_E^3}{2\pi T}}\,\left\{
2096: \begin{array}{c}
2097: \lambda_1\, \exp\left( \mp \frac{2\pi i}{\beta} \lambda_3 \tau\right)\,, \,\ \ \ \ \ \ \ \ \ \ \ \ (0\le\tau<\frac{\beta}{3})\nonumber\\
2098: \bar{\lambda}_1\, \exp\left( \mp \frac{2\pi i}{\beta} \bar{\lambda}_3 (\tau-\frac{\beta}{3})\right)\,,\ \ \ \ (\frac{\beta}{3}\le\tau<\frac{2\beta}{3})\nonumber\\
2099: \tilde{\lambda}_1\, \exp\left( \mp \frac{2\pi i}{\beta} \tilde{\lambda}_3 (\tau-\frac{2\beta}{3})
2100: \right)\,,\ \ \ \ (\frac{2\beta}{3}\le\tau<\beta)\,.
2101: \end{array}\right.\,
2102: \eqe
2103: %********
2104: Notice that the potential $V_E=2\frac{\Lambda_E^6}{|\phi|^2}$ is the same
2105: on the configuration $\phi(\tau)$ in Eq.\,(\ref{solBPSSU(3)}) as for the SU(2) case and that by the
2106: same calculation one shows its quantum mechanical and statistical
2107: inertness.
2108:
2109:
2110:
2111: \subsection{A macroscopic ground state\label{macgs}}
2112:
2113: The action Eq.\,(\ref{actionphi}) governs the dynamics of
2114: $\phi$. We have not yet included caloron interactions, mediated by the topologically trivial sector,
2115: which change the holonomy of calorons and induce interactions between
2116: their (BPS monopole) constituents. This is the objective of the present section.
2117:
2118: \subsubsection{Pure-gauge configuration\label{PGC}}
2119:
2120: The action Eq.\,(\ref{actionphi}) can be extended to include
2121: topologically trivial configurations $a_\mu$. This is accomplished by a minimal
2122: coupling $\partial_\tau \phi \to \partial_\mu \phi + ie[\phi,a_\mu] \equiv D_\mu \phi$
2123: and by adding a kinetic term for these configurations. Here $e$ denotes the {\sl effective} gauge coupling.
2124: The total Yang-Mills action $S$, governing the electric phase, can thus be rewritten as
2125: %************
2126: \eqb \label{actiontotal}
2127: S = \mbox{tr}\, \int_0^\beta d\tau \int d^3x \left( \frac12 G_{\mu\nu}G_{\mu\nu} +
2128: D_\mu \phi D_\mu \phi + \Lambda_E^6 \phi^{-2} \right)
2129: \,,
2130: \eqe
2131: %*************
2132: where $G_{\mu\nu}=G^a_{\mu\nu} \frac{\lambda^a}{2}$ and
2133: $G^a_{\mu\nu}=\pd_\mu a^a_\nu-\pd_\nu a^a_\mu+e\,f^{abc}a^b_\mu a^c_\nu$.
2134: The classical equation of motion for $a_\mu$,
2135: derived from the action (\ref{actiontotal}), reads
2136: %*********
2137: \eqb
2138: \label{gfeom}
2139: D_\mu G_{\mu\nu}=ie[\phi,D_\nu \phi] \,
2140: \eqe
2141: %**********
2142: There is a pure-gauge solution $a_\mu^{bg}$ to Eq.\,(\ref{gfeom}) with $D_\nu \phi=0$.
2143: Thus the total action density of the ground state $(\phi,a_\mu^{bg})$
2144: is the potential $V_E=4\pi\,\La_E^3\,T$, corresponding to an energy-momentum tensor
2145: $\bar{\theta}_{\mu\nu}=V_E\,\delta_{\mu\nu}$ or $P^{gs}=-\rho^{gs}=-4\pi\,\La_E^3\,T$ ($P^{gs},\rho^{gs}$
2146: the ground-state pressure and energy density, respectively): The so-far hidden scale
2147: $\Lambda_E$ becomes (gravitationally) measurable by coarse-grained interactions between calorons.\vspace{0.1cm}\\
2148: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
2149: In the background
2150: %********
2151: \eqb
2152: \label{backsu2}
2153: \phi(\tau)=\sqrt{\frac{\Lambda_E^3}{2\pi\,T}}\,
2154: \lambda_1\, \exp\left( \mp \frac{2\pi i}{\beta}\lambda_3 \tau\right)
2155: \eqe
2156: %*******
2157: we have
2158: %********
2159: \eqb
2160: \label{puregaugesu2}
2161: a_\mu^{bg}=\pm\delta_{\mu 4}\frac{\pi}{e\beta}\,\lambda_3\,.
2162: \eqe
2163: %********
2164: $\mbox{}$\vspace{0.1cm}\\
2165: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2166: In the background $\phi$ of Eq.\,(\ref{solBPSSU(3)}) the pure-gauge
2167: solution to Eq.\,(\ref{gfeom}) with $D_\nu \phi=0$ reads
2168: %********
2169: \eqb
2170: \label{puregaugesu3}
2171: a_\mu^{bg}=\pm\delta_{\mu 4}\frac{\pi}{e\beta}\,\left\{
2172: \begin{array}{c}\lambda_3\,,\ \ \ \ (0\le\tau<\frac{\beta}{3})\nonumber\\
2173: \ \bar{\lambda}_3\,,\ \ \ \ \ (\frac{\beta}{3}\le\tau<\frac{2\beta}{3})\nonumber\\
2174: \ \tilde{\lambda}_3\,,\ \ \ \ \ (\frac{2\beta}{3}\le\tau<\beta)\,.
2175: \end{array}\right.
2176: \eqe
2177: %********
2178:
2179:
2180: \subsubsection{Polyakov loop and rotation to unitary gauge\label{gtug}}
2181:
2182: Here we would like to investigate whether the
2183: ground state, described by $(\phi,a_\mu^{bg})$, is degenerate
2184: with respect to the global electric $Z_2$ (SU(2)) or
2185: $Z_3$ (SU(3)) symmetry under which the
2186: Polyakov loop ${\bf P}$ transforms as ${\bf P}\to Z {\bf P}$
2187: where $Z\in Z_2$ (SU(2)) or $Z\in Z_3$ (SU(3)). We will refer to
2188: the gauge, where $\phi$'s phase is $\tau$ dependent
2189: as in Eq.\,(\ref{ansatzBPS}) or in Eq.\,(\ref{solBPSSU(3)}),
2190: as winding gauge. (Microscopically, this is the
2191: singular gauge for an instanton in which the Harrington-Shepard solution is
2192: constructed.) \vspace{0.1cm}\\
2193:
2194: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
2195: Evaluating the Polyakov loop on the configuration $a^{bg}_\mu$ of Eq.\,(\ref{puregaugesu2}),
2196: we have
2197: %*********
2198: \eqb
2199: \label{Polsu2}
2200: {\bf P}[a^{bg}_\mu]=\exp\left[\pm i\pi\lambda_3\right]=-\UM_2\,.
2201: \eqe
2202: %*********
2203: We are searching a gauge transformation $\tilde{\Omega}\in$SU(2) to the unitary
2204: gauge $\phi=|\phi|\lambda_3$ and $a_\mu^{bg}=0$. A periodic but not
2205: differentiable gauge transformation $\tilde{\Omega}$ doing this
2206: can be obtained from a nonperiodic but smooth gauge transformation
2207: $\Omega$ by multiplication with a local center transformation
2208: $Z$ and by multiplication with a global gauge transformation $\Omega_{gl}$:
2209: %***********
2210: \eqb
2211: \label{gatugsu2}
2212: \tilde{\Omega}=\Omega(\tau) Z(\tau)\Omega_{gl}\,,
2213: \eqe
2214: %************
2215: where $\Omega(\tau)\equiv\exp[\mp i\pi\frac{\tau}{\beta}\lambda_3]$,
2216: $Z(\tau)=\left(2\Theta(\tau-\frac{\beta}{2})-1\right)\UM_2$, and
2217: $\Omega_{gl}=\exp[-i\frac{\pi}{4}\lambda_2]$. $\Theta$ denotes the Heavyside step function:
2218: %*********
2219: \eqb
2220: \label{Hevayside}
2221: \Theta(x)=\left\{\begin{array}{c}
2222: 0\,,\ \ \ \ (x<0)\,,\\
2223: \frac{1}{2}\,,\ \ \ \ (x=0)\,,\\
2224: 1\,,\ \ \ \ \ (x>0)\,.
2225: \end{array}\right.\,.
2226: \eqe
2227: %***********
2228: Applying $\tilde\Omega$
2229: to $a_\mu=a^{bg}_\mu+\delta a_\mu$,
2230: where $\delta a_\mu$ is a {\sl periodic} fluctuation in winding gauge, we have
2231: %**********
2232: \eab
2233: \label{trafiwtoug}
2234: a_\mu&\rightarrow &\tilde{\Omega}^\dagger(a^{bg}_\mu+\delta a_\mu)\tilde{\Omega}+
2235: \frac{i}{e}\pd_\mu \tilde{\Omega}^\dagger\tilde{\Omega}\nonumber\\
2236: &=&\Omega_{gl}^\dagger\left(\Omega^\dagger(a^{bg}_\mu+\delta a_\mu)\Omega+
2237: \frac{i}{e}\left((\pd_\mu\Omega^\dagger)\Omega+(\pd_\mu Z)Z\right)\right)\Omega_{gl}\nonumber\\
2238: &=&\Omega_{gl}^\dagger\left(\Omega^\dagger\delta a_\mu\Omega+
2239: \frac{2i}{e}\delta(\tau-\frac{\beta}{2})Z\right)\Omega_{gl}\nonumber\\
2240: &=&(\Omega\Omega_{gl})^\dagger\delta a_\mu\Omega\Omega_{gl}\,.
2241: \eae
2242: %***********
2243: Since $\Omega\Omega_{gl}(0)=-\Omega\Omega_{gl}(\beta)$
2244: we conclude that if the fluctuation $\delta a_\mu$
2245: is periodic in winding gauge then it is also periodic in unitary
2246: gauge. It thus is irrelevant whether we integrate out the
2247: fluctuations $\delta a_\mu$ in winding or unitary gauge in a loop expansion
2248: of thermodynamical quantities: Hosotani's mechanism \cite{Hosotani1983}
2249: does not take place. What changes though under
2250: the gauge transformation $\tilde{\Omega}$ is the Polyakov loop evaluated on
2251: the background configuration $a^{bg}_\mu$:
2252: %*********
2253: \eqb
2254: \label{Polch}
2255: {\bf P}[a^{bg}_\mu]=-\UM_2 \rightarrow {\bf P}[a^{bg}_\mu=0]=\UM_2\,.
2256: \eqe
2257: %***********
2258: We conclude that the theory has two equivalent ground states and
2259: that the global electric $Z_2$ symmetry is dynamically broken. Thus we have shown that
2260: the elecric phase is {\sl deconfining}.\vspace{0.1cm}\\
2261:
2262: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2263: Let us now compute the Polyakov loop on the configuration
2264: $a^{bg}_\mu$ of Eq.\,(\ref{puregaugesu3}). We have
2265: %*********
2266: \eab
2267: \label{Polsu3}
2268: {\bf P}[a^{bg}_\mu]&=&\exp\left[\pm i\frac{\pi}{3}\tilde{\lambda}_3\right]
2269: \exp\left[\pm i\frac{\pi}{3}\bar{\lambda}_3\right]\exp\left[\pm i\frac{\pi}{3}\lambda_3\right]\nonumber\\
2270: &=&\exp\left[i\frac{\pi}{3}(\pm\tilde{\lambda}_3\pm\bar{\lambda}_3\pm\lambda_3)\right]\,.
2271: \eae
2272: %*********
2273: The $+$ or $-$ sign can be chosen independently for
2274: each SU(2) algebra. The following combinations are possible:
2275: %******
2276: \eab
2277: \label{insu2com}
2278: \pm\left(+\tilde{\lambda}_3+\bar{\lambda}_3+\lambda_3\right)&=&\pm 2\,\bar{\lambda}_3\,,\nonumber\\
2279: \pm\left(-\tilde{\lambda}_3-\bar{\lambda}_3+\lambda_3\right)&=&\pm 2\,\tilde{\lambda}_3\,,\nonumber\\
2280: \pm\left(-\tilde{\lambda}_3+\bar{\lambda}_3+\lambda_3\right)&=&\pm 2\,\lambda_3\,,\nonumber\\
2281: \pm\left(+\tilde{\lambda}_3-\bar{\lambda}_3+\lambda_3\right)&=&0\,.
2282: \eae
2283: %*******
2284: The corresponding values of the Polyakov loop are
2285: %*******
2286: \eab
2287: \label{Polcosu3}
2288: {\bf P}^\pm_1&=&\left(\begin{array}{ccc}\exp[\pm\frac{2\pi i}{3}]&0&0\\
2289: 0&1&0\\
2290: 0&0&\exp[\mp\frac{2\pi i}{3}]\end{array}\right)\,,\ \ \ \ \
2291: {\bf P}^\pm_2=\left(\begin{array}{ccc}1&0&0\\
2292: 0&\exp[\pm\frac{2\pi i}{3}]&0\\
2293: 0&0&\exp[\mp\frac{2\pi i}{3}]\end{array}\right)\,,\nonumber\\
2294: {\bf P}^\pm_3&=&\left(\begin{array}{ccc}\exp[\pm\frac{2\pi i}{3}]&0&0\\
2295: 0&\exp[\mp\frac{2\pi i}{3}]&0\\
2296: 0&0&1\end{array}\right)\,,\ \ \ \ \ \ {\bf P}_4=\UM_3.
2297: \eae
2298: %*********
2299: ${\bf P}_4$ is a trivial representation of the center group.
2300: The set ${\bf P}^\pm_1, {\bf P}^\pm_2,{\bf P}^\pm_3$ closes
2301: under multiplication with the center elements
2302: ${\bf P}=\exp[\pm\frac{2\pi i}{3}]\UM_3,\UM_3$. It is a six
2303: dimensional, reducible representation of the center group.
2304: The two three dimensional irreducible representations, which collapse on one another, are spanned by
2305: %*********
2306: \eqb
2307: \label{irrcenter}
2308: \frac{1}{3}\UM_3\left({\bf P}^\pm_1+{\bf P}^\mp_2+{\bf P}^\pm_3\right)\,,\ \ \
2309: \frac{1}{3}\exp[\mp\frac{2\pi i}{3}]\UM_3\left({\bf P}^\pm_1+{\bf P}^\mp_2+{\bf P}^\pm_3\right)\,.
2310: \eqe
2311: %*********
2312: We conclude that the ground state has a
2313: $Z_3$ degeneracy: The electric $Z_3$ symmetry is dynamically broken and thus
2314: we have discussed a {\sl deconfining} phase.
2315:
2316: What about a gauge rotation to unitary gauge $a_\mu^{bg}=0$ and
2317: $\phi=|\phi|\lambda_3$ or $\phi=|\phi|\bar{\lambda}_3$ or $\phi=|\phi|\tilde{\lambda}_3$?
2318: Such a gauge transformation $\tilde{\Omega}$ is given as
2319: %***********
2320: \eqb
2321: \label{gatugsu3}
2322: \tilde{\Omega}=\left\{
2323: \begin{array}{c}\exp[\mp i\frac{\pi}{\beta}\tau\lambda_3]
2324: \exp[-i\frac{\pi}{4}\lambda_2]\,,\ \ \ \ \ \ \ \ \ \ \ \ \ (0\le\tau<\frac{\beta}{3})\nonumber\\
2325: \exp[\mp i\pi\frac{\pi}{\beta}(\tau-\frac{\beta}{3})\bar{\lambda}_3]
2326: \exp[-i\frac{\pi}{4}\bar{\lambda}_2]\,,\ \ \ \ \ (\frac{\beta}{3}\le\tau<\frac{2\beta}{3})\nonumber\\
2327: \exp[\mp i\frac{\pi}{\beta}(\tau-\frac{2\beta}{3})\tilde{\lambda}_3]
2328: \exp[-i\frac{\pi}{4}\tilde{\lambda}_2]\,,\ \ \ \ \ (\frac{2\beta}{3}\le\tau<\beta)\,.
2329: \end{array}\right.
2330: \eqe
2331: %************
2332: By construction $\tilde{\Omega}$ is periodic, at $\tau=\beta$ it jumps back
2333: to its value at $\tau=0$, and thus a
2334: fluctuation $\delta a_\mu$, which is periodic in winding gauge, is
2335: also periodic in unitary gauge.
2336:
2337: A comment concerning SU($N$) theories with $N\ge 4$ is in order. We only discuss the case when $N$ is even.
2338: Since at any $\tau$ the maximal number of {\sl independent} SU(2) subgroups contributing with
2339: calorons of topological charge one to the
2340: macroscopic field $\phi$ is $N/2$ the dynamical
2341: gauge-symmetry breaking is not as maximal as
2342: it is for SU(2) and SU(3). For example, SU(4) breaks only down to SU(2)$^2\times$U(1).
2343: The question is whether there is a single critical temperature $T_{c,E}$ where
2344: the unbroken nonabelian subgroups get broken to U(1) factors
2345: by the condensation of color magnetic monopoles into macroscopic
2346: adjoint Higgs fields and pure gauges, and where the abelian
2347: magnetic monopoles with respect to the remaining U(1) factors
2348: condense as soon as they are created, or whether this is
2349: a stepwise process. The lattice seems to
2350: favor the former situation \cite{LuciniTeperWenger2005} which,
2351: however, does not appear natural to the author.
2352:
2353:
2354: \subsection{Excitations}
2355:
2356: Now that the derivation of a macroscopic ground state
2357: for the electric phase is completed we are in a position to discuss
2358: the properties of its on-shell excitations and the role of
2359: residual quantum fluctuations. The temperature dependent
2360: mass is a measure for the strength of coupling between gauge modes and
2361: the nontrivial ground state. The evolution of this
2362: coupling with temperature is
2363: a manifestation of the thermodynamical selfconsistency
2364: of the separation into topological configurations
2365: and gauge modes with trivial topology. As we
2366: shall see, this evolution represents a decoupling between
2367: high and low temperature physics and, for temperatures
2368: sufficiently above the critical
2369: temperature $T_{c,E}$, indicates the conservation of
2370: magnetic charge associated with the
2371: isolated and screened BPS monopoles that are liberated by
2372: the dissociation of calorons with a large holonomy.
2373:
2374: \subsubsection{Mass spectrum of thermal quasiparticles}
2375:
2376: We first discuss some general aspects of the mass spectrum
2377: and then specialize to the cases SU(2) and SU(3). We refer to
2378: gauge modes which acquire a quasiparticle mass on tree level by the adjoint Higgs mechanism
2379: as tree-level heavy (TLH) and to those which remain
2380: massless as tree-level massless (TLM).
2381:
2382: In unitary gauge the mass spectrum calculates as
2383: %***********
2384: \eqb
2385: \label{massspectrum}
2386: m_a^2=-2e^2\,\mbox{tr}\,[\phi,t^{a}][\phi,t^{a}]\,,
2387: \eqe
2388: %***********
2389: where $t^a$ are the group generators normalized
2390: as tr\,$t^a t^b=\frac{1}{2}\delta^{ab}$. For SU($N$) off-diagonal generators
2391: can be chosen as
2392: %********
2393: \eab
2394: \label{TLgen}
2395: t^{IJ}_{rs}&=&1/2\,(\delta_r^I\delta_s^J+\delta_s^I\delta_r^J)\,,\ \ \
2396: {\bf t}^{IJ}_{rs}=-i/2\,(\delta_r^I\delta_s^J-\delta_s^I\delta_r^J)\,,\nonumber\\
2397: (I&=&1,\cdots,N;\,J>I;\,r,s=1,\cdots,N)\,.
2398: \eae
2399: %*********
2400: This yields
2401: %********
2402: \eqb
2403: \label{TLHmasses}
2404: m_{IJ}^2={\bf m}_{IJ}^2=e^2(\phi_I-\phi_J)^2\,
2405: \eqe
2406: %*********
2407: where $\phi_I,\phi_J$ denote the diagonal elements of $\phi$ in unitary gauge.\vspace{0.1cm}\\
2408: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
2409: In this case $\phi$ breaks the gauge symmetry dynamically down to U(1).
2410: Thus we have one TLM mode and two TLH modes whose degenerate
2411: masses are, according to Eqs.\,(\ref{backsu2}) and (\ref{TLHmasses}),
2412: given as
2413: %*********
2414: \eqb
2415: \label{n2masses}
2416: m_{12}^2={\bf m}_{12}^2=4\,e^2|\phi|^2=4\,e^2 \frac{\Lambda_E^3}{2\pi T}\,.
2417: \eqe
2418: %*********
2419: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2420: Here $\phi$ is diagonal in either one of the three SU(2) subgroups, and the gauge symmetry is
2421: dynamically broken to U(1)$^2$. We have
2422: %******
2423: \eab
2424: \label{n3masses}
2425: m_{12}^2&=&{\bf m}_{12}^2=4\,e^2\,\frac{\Lambda_E^3}{2\pi T}\,,\nonumber\\
2426: m_{13}^2&=&{\bf m}_{13}^2=m_{23}^2={\bf m}_{23}^2=e^2 \,\frac{\Lambda_E^3}{2\pi T}\,,
2427: \ \ \ \ \ \mbox{or}\nonumber\\
2428: m_{13}^2&=&{\bf m}_{13}^2=4\,e^2\,\frac{\Lambda_E^3}{2\pi T}\,,\nonumber\\
2429: m_{12}^2&=&{\bf m}_{12}^2=m_{23}^2={\bf m}_{23}^2=e^2 \,\frac{\Lambda_E^3}{2\pi T}\,,
2430: \ \ \ \ \ \mbox{or}\nonumber\\
2431: m_{23}^2&=&{\bf m}_{23}^2=4\,e^2\,\frac{\Lambda_E^3}{2\pi T}\,,\nonumber\\
2432: m_{12}^2&=&{\bf m}_{12}^2=m_{13}^2={\bf m}_{13}^2=e^2 \,\frac{\Lambda_E^3}{2\pi T}\,.
2433: \eae
2434: %******
2435: For $T$ sufficiently far above $T_{c,E}$ we will
2436: see in Sec.\,\ref{eveffgc} that $e$ is practically
2437: independent of $T$. As a consequence, the TLH masses
2438: die off according to the power law in
2439: Eqs.\,(\ref{n2masses}) and (\ref{n3masses}). Moreover, the energy density and the pressure of the
2440: ground state are only linear in $T$, $P^{gs}=-\rho^{gs}=-4\pi\,\La_E^3\,T$.
2441: We will show in Sec.\,\ref{Radcor} that the one-loop
2442: result for thermodynamical quantities is reliable on the $0.1$\% level
2443: throughout the electric phase. Thus the Stefan-Boltzmann
2444: limit $P=\frac{\pi^2}{15}T^4$ and $\rho=\frac{\pi^2}{5}T^4$ (SU(2))
2445: or $P=\frac{8\pi^2}{45}T^4$ and $\rho=\frac{8\pi^2}{15}T^4$ (SU(3))
2446: is reached very quickly apart from a factor arising from the extra polarizations of TLH modes. This
2447: result is in agreement with early
2448: lattice simulations using the differential method.
2449:
2450:
2451: \subsubsection{Thermodynamical selfconsistency}
2452:
2453: In this section we provide a conceptual basis for the notion of
2454: thermodynamical selfconsistency.
2455:
2456: As a result of the existence of a nontrivial macroscopic
2457: ground state, which is built of interacting calorons of
2458: topological charge-one, the ground-state physics and the
2459: properties of the excitations are temperature
2460: dependent. We have discussed in Sec.\,\ref{macgs}
2461: how the temperature dependence of the ground-state pressure
2462: and its energy density arises due to caloron
2463: interactions. These interactions are
2464: encoded in a pure-gauge configuration
2465: on the macroscopic level. For topologically
2466: trivial fluctuations $\delta a_\mu$ the two following properties are induced
2467: by the ground-state physics. First, in unitary gauge off-Cartan fluctuations
2468: are massive in a temperature dependent way, compare with
2469: Eqs.\,(\ref{n2masses}) and (\ref{n3masses}). Notice that the
2470: quasiparticle masses are related to the ground-state
2471: pressure by their respective dependences on temperature
2472: if the temperature dependence of the
2473: effective gauge coupling $e$ is known. Second, there are
2474: constraints for the maximal off-shellness of the fluctuations
2475: $\delta a_\mu$ arising from the compositeness of
2476: the ground-state field $\phi$. Namely, a fluctuation
2477: $\delta a_\mu$, which was generated in a thermal equilibrium situation
2478: by the ground state, is not capable of
2479: destroying this ground state: In a physical gauge vacuum fluctuations or scattering
2480: processes with momenta or momentum transfers larger
2481: than the (temperature dependent)
2482: scale $|\phi|$ are forbidden.
2483:
2484: Thermodynamical quantities such as the pressure,
2485: the energy density, or the
2486: entropy density are interrelated by Legendre transformations.
2487: These transformations can be derived from the partition function
2488: of the fundamental theory where the temperature dependence only occurs in
2489: an explicit way through the Boltzmann weight. A reformulation of the
2490: theory into a spatially coarse-grained Lagrangian, where certain parameters are
2491: temperature dependent (implicit temperature dependences)
2492: by virtue of a separation into a ground state and (very weakly interacting) excitations,
2493: see Eq.\,(\ref{actiontotal}), must not alter
2494: the Legendre transformations between thermodynamical quantities.
2495: For this to be true in the effective theory
2496: one needs to impose that in each transformation law the total
2497: derivatives with respect to temperature involves the explicit
2498: temperature dependences only. That is, derivatives with respect to implicit
2499: temperature dependences ought to cancel in a given Legendre
2500: transformation.
2501:
2502: A particular and essential Legendre transformation maps the total pressure onto the total
2503: energy density as
2504: %*********
2505: \eqb
2506: \label{rhoPree}
2507: \rho=T\frac{dP}{dT}-P\,.
2508: \eqe
2509: %*********
2510: If the effective theory has temperature dependent quasiparticle fluctuations of mass
2511: $m_a=c_a\,m$, where $c_a$ are dimensionless constants,
2512: and a ground-state pressure $P^{gs}$, which can be regarded a function of $m$,
2513: then the condition of thermodynamical selfconsistency
2514: is expressed as \cite{Gorenstein1995}
2515: %*********
2516: \eqb
2517: \label{TSCm}
2518: \pd_m P=0\,.
2519: \eqe
2520: %*********
2521: Eq.\,(\ref{TSCm}) assures that
2522: in Eq.\,(\ref{rhoPree}) only derivatives with respect to the
2523: explicit temperature dependence of $P$ contribute since
2524: $\frac{dP}{dT}=\pd_T P+\pd_m P\frac{d\,m}{d\,T}$. In $\pd_m P=0$ the
2525: derivative of the pressure contribution arising from fluctuations cancels against that
2526: arising from the ground state. Eq.\,(\ref{TSCm}) governs
2527: the temperature evolution of the effective coupling $e$
2528: at any loop order that $P$ is expanded in.
2529:
2530: The higher the loop order the more
2531: complicated the implementation of Eq.\,(\ref{TSCm}).
2532: In Sec.\,\ref{eveffgc} we perform an analysis on the
2533: one-loop level (noninteracting quasiparticles)
2534: which is more than sufficient
2535: for many practical purposes. A discussion of Eq.\,(\ref{TSCm}) on
2536: the two-loop level is given in Sec.\,\ref{loopexp}.
2537:
2538:
2539: \subsubsection{Compositeness constraint and pressure at one loop}
2540:
2541: We work in a
2542: physical gauge where TLM modes are transverse
2543: (two polarizations, Coulomb gauge with respect to the
2544: unbroken gauge group), and where TLH modes have three polarizations and do not
2545: interact with the pure-gauge configuration of the ground state (unitary gauge),
2546: for details see Sec.\,\ref{Radcor}.
2547: This is a physical gauge fixing which needs no introduction of additional fields
2548: since the Coulomb ghosts decouple from the dynamics.
2549:
2550: On the one-loop level, see Fig.\,\ref{diagpress}, there are no interactions
2551: between the fluctuations $\delta a_\mu$.
2552: %***********************
2553: \begin{figure}
2554: \begin{center}
2555: \leavevmode
2556: %\epsfxsize=9.cm
2557: \leavevmode
2558: %\epsffile[80 25 534 344]{}
2559: \vspace{2.5cm}
2560: \special{psfile=Fig-9.ps angle=0 voffset=-75
2561: hoffset=-100 hscale=60 vscale=60}
2562: \end{center}
2563: \caption{Diagrams contributing to the pressure when radiative corrections are
2564: ignored. A thick line corresponds to TLH modes and a
2565: thin one to TLM modes. The cross depicts the ground-state contribution
2566: that arises from interacting calorons.
2567: \label{diagpress}}
2568: \end{figure}
2569: %************************
2570: The only relevant compositeness
2571: constraint, related to the maximal off-shellness of a
2572: quantum fluctuation created by the ground state, thus is
2573: %**********
2574: \eab
2575: \label{comconnoi}
2576: |p^2-m_a^2|&\le & |\phi|^2\ \ \ \ (\mbox{Minkowskian signature})\nonumber\\
2577: p^2+m_a^2&\le & |\phi|^2\ \ \ \ (\mbox{Euclidean signature})\,.
2578: \eae
2579: %**********
2580: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
2581: Before we discuss the thermal contribution to the
2582: one-loop pressure let us investigate what the
2583: constraints (\ref{comconnoi}) imply for the one-loop
2584: quantum correction $-\Delta V_E$. Setting $m_a\equiv 0$ in (\ref{comconnoi}) and considering
2585: two polarizations for TLM and
2586: three polarizations for TLH modes,
2587: an upper bound on $|\Delta V_E|$ can be obtained as
2588: %*********
2589: \eqb
2590: \label{1loopvacbubble}
2591: |\Delta V_E|<\frac{1}{\pi^2}\int_0^{|\phi|}dp\,
2592: p^3\log\left(\frac{p}{|\phi|}\right)=\frac{|\phi|^4}{16\pi^2}\,.
2593: \eqe
2594: %********
2595: Thus we have
2596: %*******
2597: \eqb
2598: \label{ratioVE}
2599: \left|\frac{\Delta V_E}{V_E}\right|<
2600: \frac{1}{32\pi^2}\left(\frac{|\phi|}{\La_E}\right)^6=
2601: \frac{\lambda_E^{-3}}{32\pi^2}\,.
2602: \eqe
2603: %**********
2604: Since $\lambda_E>13.867$ when omitting $\Delta V_E$
2605: in the one-loop evolution of $e$, see Sec.\,\ref{eveffgc}, this omission
2606: is justified: One-loop quantum corrections
2607: to the pressure are suppressed as compared to the tree-level
2608: result by a factor less than $2\times 10^{-6}$.
2609:
2610: Omitting the tiny quantum part, the one-loop expression for
2611: the pressure, including the ground-state contribution, reads:
2612: %********
2613: \eqb
2614: \label{treelPsu2}
2615: P(\lambda_E)=-\La_E^4\left\{\frac{2\lambda_E^4}{(2\pi)^6}\left[2\bar{P}(0)+
2616: 6\bar{P}(2a)\right]+2\lambda_E\right\}\,
2617: \eqe
2618: %*******
2619: where $\lambda_E\equiv\frac{2\pi T}{\La_E}$ and
2620: %******
2621: \eqb
2622: \label{aofela}
2623: a\equiv\frac{m}{2T}=e\frac{|\phi|}{T}=e\sqrt{\frac{\La_E^3}{2\pi T^3}}=2\pi\,e\lambda_E^{-3/2}\,.
2624: \eqe
2625: %*******
2626: The (negative) function $\bar{P}(a)$ is defined as
2627: %**********
2628: \eqb
2629: \label{P(y)}
2630: \bar{P}(a)\equiv\int_0^\infty dx \,x^2\,\log[1-\exp(-\sqrt{x^2+a^2})]\,.
2631: \eqe
2632: %*****************
2633: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2634: Here a similar estimate for the quantum contribution to the
2635: one-loop pressure holds as in Eq.\,(\ref{1loopvacbubble}),
2636: and thus, again, we only have to consider the thermal part.
2637: It reads
2638: %********
2639: \eqb
2640: \label{treelPsu3}
2641: P(\lambda_E)=-\La_E^4\left\{\frac{2\lambda_E^4}{(2\pi)^6}\left[4\bar{P}(0)+3\left(
2642: 4\bar{P}(a)+2\bar{P}(2a)\right)\right]+
2643: 2\lambda_E\right\}\,
2644: \eqe
2645: %*******
2646: with the same definitions as in the SU(2) case.
2647:
2648:
2649: \subsubsection{One-loop evolution of effective gauge coupling\label{eveffgc}}
2650:
2651: Let us now implement the condition (\ref{TSCm}). We have
2652: %********
2653: \eqb
2654: \label{TSCa}
2655: \pd_m P=0\ \ \Leftrightarrow\ \ \pd_{(aT)} P=0\,.
2656: \eqe
2657: %********
2658: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
2659: Substituing Eq.\,(\ref{treelPsu2})
2660: into Eq.\,(\ref{TSCa}) yields the following evolution
2661: equation
2662: %********
2663: \eqb
2664: \label{eeq}
2665: \pd_a \lambda_E=-\frac{24\,\lambda_E^4\,a}{(2\pi)^6}\frac{D(2a)}{1+\frac{24\,\lambda_E^3\,a^2}{(2\pi)^6}\,D(2a)}\,,
2666: \eqe
2667: %*********
2668: where the function $D(a)$, see Fig.\,\ref{DAf},
2669: %***********************
2670: \begin{figure}
2671: \begin{center}
2672: \leavevmode
2673: %\epsfxsize=9.cm
2674: \leavevmode
2675: %\epsffile[80 25 534 344]{}
2676: \vspace{4.5cm}
2677: \special{psfile=Fig-10.ps angle=0 voffset=-140
2678: hoffset=-150 hscale=60 vscale=50}
2679: \end{center}
2680: \caption{The functions $a\,D(a)$ and $a^2\,D(a)$. \label{DAf}}
2681: \end{figure}
2682: %************************
2683: is defined as
2684: %**********
2685: \eqb
2686: \label{DA}
2687: D(a)\equiv \int_0^{\infty} dx\,
2688: \frac{x^2}{\sqrt{x^2+a^2}}\frac{1}{\exp(\sqrt{x^2+a^2})-1}\,.
2689: \eqe
2690: %**********
2691: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
2692: The evolution equation is obtained by substituting Eq.\,(\ref{treelPsu3}) into Eq.\,(\ref{TSCa}):
2693: %********
2694: \eqb
2695: \label{eeq3}
2696: \pd_a \lambda_E=-\frac{12\,\lambda_E^4\,a}{(2\pi)^6}
2697: \frac{D(a)+2\,D(2a)}{1+\frac{12\,\lambda_E^3\,a^2}{(2\pi)^6}\,\left(D(a)+2\,D(2a)\right)}\,.
2698: \eqe
2699: %*********
2700: Each of the Eqs.\,(\ref{eeq}) and (\ref{eeq3}) has two fixed points,
2701: one at $a=0$ and one at $a=\infty$, see Fig.\,\ref{DAf}. The points $\lambda_{P,E}\equiv\lambda_E(a\to 0)$ and
2702: $\lambda_{c,E}\equiv\lambda_E(a=\infty)$ are
2703: associated with the highest and the lowest temperatures, respectively, which are attainable
2704: in the electric phase.
2705:
2706: Above $\lambda_{P,E}$ the
2707: ground state would be trivial (topological fluctuations are absent) and thus
2708: no tree-level quasiparticle mass were generated by a caloron induced Higgs mechanism.
2709: This re-introduces the problem of the diverging loop expansion as it is encountered
2710: in thermal perturbation theory \cite{Linde1980} and
2711: thus makes the thermalized Yang-Mills theory inconsistent
2712: as an interacting field theory. We conclude that $\lambda_{P,E}$ marks the
2713: point in temperature where the field theoretic implementation
2714: of the gauge principle breaks down. For a physics model, whose gauge group is a product of
2715: SU(2) and/or SU(3) groups, we expect that
2716: $\lambda_{P,E}\sim \frac{2\pi M_P}{\La_E}$ where
2717: $M_P=1.22\times\,10^{19}\,$GeV is the Planck mass.
2718:
2719: As we shall see below, the point $\lambda_{c,E}$, where each
2720: tree-level quasiparticle mass diverges, marks a transition
2721: to a (pre-confining) phase with condensed magnetic monopoles, dynamically broken dual
2722: gauge symmetries U(1)$_D$ (SU(2)) and U(1)$_D^2$ (SU(3)), and isolated but instable
2723: center-vortex loops: The magnetic phase.
2724:
2725: Notice that the right-hand sides of Eqs.\,(\ref{eeq}) and (\ref{eeq3})
2726: are negative definite. As a consequence, the solutions $\lambda_E(a)$
2727: to Eqs.\,(\ref{eeq}) and (\ref{eeq3}) can be inverted, and, according to Eq.\,(\ref{aofela}),
2728: one obtains the evolution of the effective gauge coupling $e$ with temperature as
2729: %*******
2730: \eqb
2731: \label{coupP}
2732: e(\lambda_E)=\frac{1}{2\pi}a(\lambda_E)\lambda_E^{3/2}\,.
2733: \eqe
2734: %*******
2735: Inspecting the right-hand of Eqs.\,(\ref{eeq}) and (\ref{eeq3})
2736: in the vicinity of the point $\lambda_{c,E}$, it follows
2737: with Eq.\,(\ref{coupP}) that $e$ diverges logarithmically at $\lambda_{c,E}$:
2738: %*********
2739: \eqb
2740: \label{divloglae}
2741: e(\lambda_E)\sim -\log(\lambda_E-\lambda_{c,E})\,.
2742: \eqe
2743: %*********
2744: %***********************
2745: \begin{figure}
2746: \begin{center}
2747: \leavevmode
2748: %\epsfxsize=9.cm
2749: \leavevmode
2750: %\epsffile[80 25 534 344]{}
2751: \vspace{4.8cm}
2752: \special{psfile=Fig-11.ps angle=0 voffset=-146
2753: hoffset=-150 hscale=60 vscale=50}
2754: \end{center}
2755: \caption{The solution $\lambda_E(a)$ to Eq.\,(\protect\ref{eeq}) subject to the initial
2756: condition $\lambda_{P,E}=10^7$.\label{laofa}}
2757: \end{figure}
2758: %************************
2759: In Fig.\,\ref{laofa} a solution to Eq.\,(\ref{eeq}) subject to
2760: the initial condition $\lambda_{P,E}\equiv\lambda_E(a=0)=10^7$ is shown.
2761: We have noticed numerically that the low-temperature behavior of $\lambda_E(a)$
2762: is practically independent of the value $\lambda_{P,E}$ as long as $\lambda_{P,E}$ is
2763: sufficiently large. Let us show this analytically. For $a$ sufficiently smaller than unity we may
2764: expand the right-hand side of Eq.\,(\ref{eeq}) only taking the
2765: linear term in $a$ into account. In this regime, the
2766: inverse of the solution $\lambda_E(a)$ is of the following form
2767: %********
2768: \eqb
2769: \label{anasolv}
2770: a\propto \lambda_E^{-3/2}\sqrt{1-\left(\frac{\lambda_E}{\lambda_{E,P}}\right)^3}\,.
2771: \eqe
2772: %********
2773: If $\lambda_E$ is sufficiently smaller than $\lambda_{P,E}$ then this can be approximated as
2774: %********
2775: \eqb
2776: \label{anasolapp}
2777: a\propto\lambda_E^{-3/2}\,.
2778: \eqe
2779: %********
2780: Thus the dependence in Eq.\,(\ref{anasolapp})
2781: is an {\sl attractor} of the downward-in-temperature evolution
2782: as long as $a$ remains sufficiently small: If we are only interested
2783: in the behavior of the theory not too far above $\lambda_{c,E}$
2784: then it is {\sl irrelevant} what the value of
2785: $\lambda_{P,E}$ is as long as
2786: it is sizably larger than $\lambda_{c,E}$.
2787: This result is reminiscent of the
2788: ultraviolet-infrared perturbative decoupling property of the renormalizable,
2789: underlying theory. Notice that the dependence of $a$ on $\lambda_E$ in
2790: Eq.\,(\ref{anasolapp}) is canceled by the explcicit dependence of
2791: $e$ on $\lambda_E$ in Eq.\,(\ref{coupP}). Thus a plateau value $e(\lambda_E)=\mbox{const}$
2792: is reached rapidly.
2793:
2794: In Fig.\,\ref{eoflam} the temperature dependence of $e$
2795: for SU(2) and SU(3), subject to the initial condition $\lambda_{P,E}=10^7$, is shown
2796: for $\lambda_E\le 500$.
2797: %***********************
2798: \begin{figure}
2799: \begin{center}
2800: \leavevmode
2801: %\epsfxsize=9.cm
2802: \leavevmode
2803: %\epsffile[80 25 534 344]{}
2804: \vspace{5.5cm}
2805: \special{psfile=Fig-12.ps angle=0 voffset=-150
2806: hoffset=-150 hscale=60 vscale=50}
2807: \end{center}
2808: \caption{The temperature evolution of the gauge
2809: coupling $e$ in the electric phase for
2810: SU(2) (grey line) and SU(3) (black line). The gauge coupling
2811: diverges logarithmically, $e\propto -\log(\lambda_{E}-\lambda_{c,E})$, at
2812: $\lambda_{c,E}=13.867$ (SU(2)) and $\lambda_{c,E}=9.475$ SU(3).
2813: The respective plateau values are $e=8.89$ and $e=7.26$.\label{eoflam}}
2814: \end{figure}
2815: %************************
2816: Before we interpret our results a remark on the
2817: interpretation of the effective gauge coupling constant $e$ is in order.
2818: Since $e$ determines the
2819: strength of the interaction between topologically trivial gauge
2820: field fluctuations $\delta a_\mu$
2821: and the {\sl macroscopic} manifestation $\phi$ of interacting calorons in the ground state
2822: $e$ is {\sl not} equal to the perturbatively generated
2823: gauge coupling constant $\bar{g}$ of the fundamental Yang-Mills theory.
2824: From Fig.\,\ref{eoflam} we see that the effective gauge coupling $e$
2825: evolves to values larger than unity. Naively, one would
2826: conclude that the theory is strongly coupled and
2827: that the one-loop evolution of $e$ contradicts
2828: itself. This, however, is an incorrect conclusion. Because of
2829: compositeness constraints and the infrared regularization provided by
2830: TLH masses higher loop orders turn out to be very small
2831: as compared to the one-loop result for the pressure,
2832: see Sec.\,\ref{loopexp}.
2833:
2834: We interpret the fact that the
2835: gauge coupling constant $e$ is constant for $\lambda_E$ sizably
2836: larger than $\lambda_{c,E}$ as another indication for the
2837: existence of spatially isolated and conserved
2838: magnetic charges in the system. These
2839: charges are screened by calorons with a
2840: small holonomy, and thus $e$ measures
2841: the {\sl effective} magnetic charge of a BPS monopole which
2842: is given as $g=\frac{4\pi}{e}$.
2843: The plateau values for $e$ are $e\sim 8.89$ (SU(2)) and $e\sim 7.26$ (SU(3)).
2844: At $\lambda_{c,E}=13.867$ (SU(2)) or $\lambda_{c,E}=9.475$ (SU(3))
2845: both the core size $\sim \frac{\beta}{e}$ of a screened BPS
2846: monopole and its mass $\sim \frac{4\pi}{e}\,u_{\tiny\mbox{max}}=\frac{4\pi^2}{e\beta}$
2847: {\sl vanish}, see Sec.\,\ref{BPSM}. (Notice that monopoles are very massive at high temperatures.)
2848: Thus monopoles are not well separated anymore at $\lambda_{c,E}$ because it is extremely easy to move them. Local magnetic
2849: charge conservation is violated since
2850: in a typical volume $\beta^3$ the number
2851: of monopoles no longer is conserved. The increasing mobility of monopoles and the
2852: increasing violation of local
2853: charge conservation can be seen
2854: in the evolution of $e$ for
2855: $\lambda_E\searrow\lambda_{c,E}$
2856: where an increase of $e$ as
2857: compared to the plateau value is observed,
2858: see Fig.\,\ref{eoflam} and Eq.\,(\ref{divloglae}).
2859:
2860: For $\lambda_E\searrow\lambda_{c,E}$ TLH modes decouple (their masses diverge) and
2861: thus the (small) interaction between TLM modes
2862: dies off. This is the macroscopic
2863: manifestation of the fact that the magnetic
2864: charge of (dynamical and screened) BPS monopoles vanishes at $\lambda_{c,E}$
2865: making them unavailable as `scattering centers'
2866: for TLM modes. This is, indeed, seen as a result
2867: of a two-loop calculation of the SU(2)
2868: pressure in the electric phase \cite{HerbstHofmannRohrer2004,SHG2006},
2869: see Sec.\,\ref{loopexp}.
2870:
2871: \subsection{Radiative corrections for SU(2)\label{Radcor}}
2872:
2873: \subsubsection{Electric screening mass for TLM modes\label{elscmass}}
2874:
2875: In this section we investigate for SU(2) the one-loop contribution
2876: to the electric screening mass for the TLM mode.
2877:
2878: In unitary gauge the analytically continued propagator of
2879: a free TLH mode $D^{\tiny\mbox{TLH},IJ,0}_{\mu\nu,ab}(k,T)$
2880: is that of a massive vector boson \cite{LandsmanWeert1987}
2881: %***********
2882: \eab
2883: \label{TLHprop}
2884: D^{\tiny\mbox{TLH},IJ,0}_{\mu\nu,ab}(k,T)&=&-\delta_{ab}
2885: \left(g_{\mu\nu}-\frac{k_\mu k_\nu}{4\,(e|\phi|)^2}\right)\times\nonumber\\
2886: &&\left[\frac{i}{k^2-4\,(e|\phi|)^2}+2\pi\delta(k^2-4\,(e|\phi|)^2)n_B(|k_0|/T)\right]\,,
2887: \eae
2888: %**********
2889: where $n_B(x)=\frac{1}{\e^x-1}$ is the Bose distribution and
2890: $k_0=\pm\sqrt{\vec{k}^2+4\,(e|\phi|)^2}$. The electric (or Debye)
2891: screening mass $m_D$ is related \cite{Linde1980} to the 00-component of the polarization
2892: tensor $\Pi_{\mu\nu}(k)$ in the limit $k_0=0,\vec{k}\to 0$.
2893: $\Pi_{\mu\nu}(k)$ is calculated according to
2894: the diagrams in Fig.\,\ref{Fig3e}.
2895: %***********************
2896: \begin{figure}
2897: \begin{center}
2898: \leavevmode
2899: %\epsfxsize=9.cm
2900: \leavevmode
2901: %\epsffile[80 25 534 344]{}
2902: \vspace{3.5cm}
2903: \special{psfile=Fig-13.ps angle=0 voffset=-90
2904: hoffset=-130 hscale=50 vscale=40}
2905: \end{center}
2906: \caption{Diagrams for the TLM polarization tensor at one loop.
2907: Thick and thin lines denote TLH and TLM propagation, respectively.\label{Fig3e}}
2908: \end{figure}
2909: %************************
2910: The vertices for the interactions of TLH and TLM modes are the usual ones. In addition to
2911: the compositeness constraint (\ref{comconnoi}) we have constraints associated with
2912: the maximally allowed momentum transfer in a four vertex. Let the three independent momenta be $p_1, p_2$,
2913: and $p_3$. Then we have \cite{HofmannSept2006}:
2914: %**********
2915: \eab
2916: \label{comcon4vert}
2917: |(p_1+p_2)^2|&\le&|\phi|^2\,,\ \ \ (s\ \mbox{channel})\ \ \ \ \ \ \ \ \
2918: |(p_3-p_1)^2|\le|\phi|^2\,,\ \ \ (t\ \mbox{channel})\nonumber\\
2919: |(p_2-p_3)^2|&\le&|\phi|^2\,,\ \ \ (u\ \mbox{channel})\,.
2920: \eae
2921: %**********
2922: One can easily see that these
2923: constraints together with the constraint (\ref{comconnoi})
2924: do not allow for any (neither from quantum nor from thermal propagation of the intermediate states)
2925: contribution of the local (or tadpole) diagrams in Fig.\,\ref{Fig3e} in the
2926: limit $k_0=0,\vec{k}\to 0$ and for $e_{\tiny\mbox{plateau}}=8.89$ (SU(2)). A calculation of the nonlocal
2927: diagram reveals that only thermal intermediate
2928: states contribute in the limit $k_0=0$ and for $e=13.867$.
2929: Notice that in perturbation theory the nonlocal diagram does not contribute
2930: to $\Pi_{00}(k_0=0,\vec{k})$. The result for $\Pi_{00}(k_0=0,\vec{k})$ reads\footnote{The author
2931: would like to thank Ulrich Herbst for performing this calculation.}
2932: %***********
2933: \eab
2934: \label{Pi00limit}
2935: &&\Pi_{00}(k_0=0,\vec{k})=\nonumber\\
2936: &&\frac{ie^2}{2\pi}\left\{\frac{1}{2}\left(\frac{12}{|\vec{k}|}+
2937: 4\frac{|\vec{k}|}{4(e|\phi|)^2}+\frac{|\vec{k}|^3}{(4(e|\phi|)^2)^2}\right)\times\right.\nonumber\\
2938: &&\left.\int_{\frac{|\vec{k}|}{2}}^\infty d|\vec{p}|\,|\vec{p}|\sqrt{|\vec{p}|^2+4(e|\phi|)^2}
2939: \left[n_B\left(\frac{\sqrt{|\vec{p}|^2+4(e|\phi|)^2}}{T}\right)\right]^2-\right.\nonumber\\
2940: &&\left.\left(4|\vec{k}|+\frac{|\vec{k}|^3}{4(e|\phi|)^2}\right)
2941: \int_{\frac{|\vec{k}|}{2}}^\infty d|\vec{p}|\,\frac{|\vec{p}|}{\sqrt{|\vec{p}|^2+4(e|\phi|)^2}}
2942: \left[n_B\left(\frac{\sqrt{|\vec{p}|^2+4(e|\phi|)^2}}{T}\right)\right]^2\right\}\,.\nonumber\\
2943: \eae
2944: %***********
2945: Thus $\Pi_{00}(k_0=0,\vec{k})$ is purely
2946: {\sl imaginary} and {\sl diverges} for $|\vec{k}|\to 0$: The electric screening
2947: mass $m_D$, which is the positive square root of $\Pi_{00}(k_0=0,\vec{k}\to 0)$,
2948: has an infinite and positive real part at {\sl finite} coupling $e$: Static
2949: electric fields of long wavelength are completely screened
2950: by calorons. For $e\to\infty$ the Boltzmann factors
2951: in the integrals in Eq.\,(\ref{Pi00limit}) make $\Pi_{00}(k_0=0,\vec{k})$
2952: vanish at {\sl any} momentum $\vec{k}$.
2953:
2954:
2955: \subsubsection{Two-loop result for the SU(2) pressure\label{loopexp}}
2956:
2957: The nonvanishing two-loop diagrams contributing to the SU(2) pressure
2958: in the electric phase \cite{LandsmanWeert1987} are shown in Fig.\,\ref{looppress}.
2959: %***********************
2960: \begin{figure}
2961: \begin{center}
2962: \leavevmode
2963: %\epsfxsize=9.cm
2964: \leavevmode
2965: %\epsffile[80 25 534 344]{}
2966: \vspace{3.5cm}
2967: \special{psfile=Fig-14.ps angle=0 voffset=-90
2968: hoffset=-140 hscale=40 vscale=40}
2969: \end{center}
2970: \caption{Nonvanishing two-loop diagrams contributing to the pressure. Thick lines denote propagators of TLH modes,
2971: thin lines those of TLM modes.\label{looppress}}
2972: \end{figure}
2973: %************************
2974: Because of the strong screening of near-to-static electric
2975: modes, compare with Sec.\,\ref{elscmass}, the
2976: TLM propagator in Coulomb gauge can safely be approximated as
2977: %**************
2978: \eqb
2979: \label{TLMprop}
2980: D^{\tiny\mbox{TLM},0}_{\mu\nu,ab}(k,T)=-\delta_{ab}\left\{P^T_{\mu\nu}
2981: \left(\frac{i}{k^2}+2\pi\delta(k^2)n_B(|k_0|/T)\right)-i\frac{u_\mu u_\nu}{\vec{k}^2}\right\}\,
2982: \eqe
2983: %**************
2984: where
2985: %***********
2986: \eab
2987: \label{PT}
2988: P^T_{00}&=&P^T_{0i}=P^T_{i0}=0\,,\nonumber\\
2989: P^T_{ij}&=&\delta_{ij}-\frac{k_ik_j}{\vec{k}^2}\,,
2990: \eae
2991: %***********
2992: $k_0=\pm |\vec{k}|$, and $u_\mu=\delta_{0\mu}$.
2993:
2994: The result of a calculation of the diagrams in
2995: Fig.\,\ref{looppress} was published in \cite{HerbstHofmannRohrer2004}.
2996: We do only outline this (lengthy) calculation here.
2997:
2998: The following nomenclature is useful. Each diagram can be split into contributions
2999: arising from the vacuum ($v$), the Coulomb ($c$) ($\propto u_\mu u_\nu$ in Eq.\,(\ref{TLMprop})), and the thermal ($t$) parts of the
3000: involved propagators. Moreover, if a TLH propagator
3001: contributes to a given diagram then this situation is abreviated by
3002: $H$, in the other case by $M$. With $e$ being larger than the one-loop plateau value $e_{\tiny\mbox{plateau}}=8.89$
3003: the compositeness constraint in Eq.\,(\ref{comconnoi}) allows for the five following two-loop radiative
3004: corrections to the pressure only:
3005: %********
3006: \eqb
3007: \label{deltacorrpre}
3008: \frac{1}{8}\Delta P^{HH}_{tt}\,,\ \
3009: \frac{1}{8}\Delta P^{HM}_{tt}\,,\ \
3010: \frac{1}{8}\Delta P^{HM}_{tv}\,,\ \
3011: \frac{1}{8}\Delta P^{HM}_{tc}\,,\ \
3012: \frac{1}{4}\left(\Delta P^{HHM}_{ttv}+
3013: \Delta P^{HHM}_{ttc}\right)\,.
3014: \eqe
3015: %*********
3016: The ratio of the two-loop corrections and the one-loop result
3017: (excluding the ground-state contribution) is plotted in Fig.\,\ref{Fig7b}.
3018: %***********************
3019: \begin{figure}
3020: \begin{center}
3021: \leavevmode
3022: %\epsfxsize=9.cm
3023: \leavevmode
3024: %\epsffile[80 25 534 344]{}
3025: \vspace{10.0cm}
3026: \special{psfile=Fig-15.ps angle=0 voffset=-285
3027: hoffset=-180 hscale=60 vscale=50}
3028: \end{center}
3029: \caption{Ratios (a) $\frac{1}{8 P_{1-loop}}\Delta P^{HH}_{tt}$ ,
3030: (b) $\frac{1}{8 P_{1-loop}}\Delta P^{HM}_{tt}$, (c) $>\frac{1}{8 P_{1-loop}}\Delta P^{HM}_{tv}$,
3031: (d) $>\frac{1}{8 P_{1-loop}}\Delta P^{HM}_{tc}$, and (e) $\frac{1}{4 P_{1-loop}}\left(\Delta P^{HHM}_{ttv}+
3032: \Delta P^{HHM}_{ttc}\right)$ as functions
3033: of temperature.
3034: \label{Fig7b}}
3035: \end{figure}
3036: %************************
3037: For $\frac{1}{8 P_{1-loop}}\Delta P^{HM}_{tv}$ the plot represents only an
3038: upper bound for the modulus of the correction, all other plots
3039: are exact results. The dominating
3040: correction $\frac{1}{4 P_{1-loop}}\Delta P^{HHM}_{ttv}$
3041: arises from the nonlocal diagram in
3042: Fig.\,\ref{looppress}.
3043: It is negative. The dip is microscopically
3044: explained by the increasing effect of TLM modes scattering off of decreasingly massive
3045: magnetic monopoles close to the phase
3046: transition at $\lambda_{c,E}=13.867$. The effect
3047: of this scattering is suppressed for $\lambda_E\gg \lambda_{c,E}$
3048: since then the monopoles are dilute and massive scattering centers. (Recall that their
3049: mass is $\propto \frac{T}{e}$ after screening.)
3050: Notice, that the presence of massive but dilute
3051: scattering centers causes the contribution
3052: $\frac{1}{8 P_{1-loop}}\Delta P^{HM}_{tt}$ to remain finite but small
3053: for asymptotically high temperatures.
3054:
3055: A comment concerning thermodynamical selfconsistency on the two-loop level is in order.
3056: Recall that on the one-loop level
3057: we have obtained an evolution equation from the requirement of thermal
3058: selfconsistency $\pd_m P=0$. This gave a functional relation between temperature and mass
3059: which could be inverted for all temperatures in the electric phase.
3060: After the relation Eq.\,(\ref{coupP}) between
3061: coupling constant $e$ and mass $a$ was exploited we obtained
3062: a functional dependence of the effective gauge coupling constant
3063: $e$ on temperature. Equivalently, we could have
3064: demanded $\pd_e P=0$ since $e$ is the only variable parameter (apart from the scale $\La_E$)
3065: of the effective theory in the electric phase. This would {\sl directly} have
3066: generated an evolution equation for temperature
3067: as a function of $e$. Because each diagram comes with a prefactor $e^2$
3068: and due to the compositeness constraints
3069: radiative corrections $\Delta P$ to the
3070: pressure have a separate dependence on $a$ and $e$,
3071: %*********
3072: \eqb
3073: \label{radpressure}
3074: \Delta P=T^4 \Delta\tilde{P}(e,a,\lambda_E)\,,
3075: \eqe
3076: %**********
3077: where $\Delta\tilde{P}$ is a dimensionless function of its dimensionless arguments.
3078: To implement thermodynamical selfconsistency by demanding $\pd_m P=0$ one has to
3079: express the explicitly appearing $e$ in Eq.\,(\ref{radpressure})
3080: in terms of $a$ by means of Eq.\,(\ref{coupP}) and distinguish temperature
3081: dependences arising from a simple rescaling and those
3082: arising from the $T$ dependent ground-state physics. For SU(2) we have
3083: %*********
3084: \eqb
3085: \label{e(a)}
3086: \frac{m^2}{|\phi|^2}\equiv e^2(a,\lambda_E)=T^2\,\times\frac{a^2}{|\phi|^2}=
3087: \frac{\lambda_E^2}{4\pi^2}\times a^2\lambda_E\,.
3088: \eqe
3089: %*******
3090: The first factor on the right-hand sides of Eq.\,(\ref{e(a)}) arises
3091: from rescaling, so only the second factor needs to be differentiated:
3092: %*********
3093: \eqb
3094: \label{e(a)D}
3095: \pd_a e^2(a,\lambda_E)=\frac{\lambda_E^2}{4\pi^2}\times\left(2a\lambda_E+a^2\pd_a\lambda_E\right)\,.
3096: \eqe
3097: %*******
3098: After solving $\pd_m P=0$ for the term $\pd_a\lambda_E$
3099: we obtain a modified right-hand side of the
3100: evolution equation Eq.\,(\ref{eeq}). The two-loop evolution of $e$
3101: is work in progress \cite{RohrerHofmann2005}. The
3102: investigation of the screening masses and of the two-loop
3103: pressure for SU(3) is left for future work. The latter
3104: will be of a similar magnitude as in the SU(2) case.
3105:
3106:
3107: \section{The magnetic phase\label{MP}}
3108:
3109: The electric phase is terminated at the temperature $T_{c,E}$ by the
3110: condensation of massless magnetic monopoles. These condensates can be described macroscopically
3111: by quantum mechanically and
3112: statistically inert complex scalar fields and pure gauges. The latter describe
3113: the interactions between monopoles which induce a
3114: negative pressure by the generation of
3115: isolated but collapsing magnetic flux loops (center-vortex loops).
3116: At $T_{c,E}$ one can not distinguish between the unbroken U(1) and the
3117: dual gauge group U(1)$_D$ (SU(2)) or the unbroken U(1)$^2$ and the
3118: dual gauge group U(1)$^2_D$ (SU(3)): An exact
3119: electric-magnetic duality occurs. For SU(2)
3120: the point $T_{c,E}$ is stabilized by a dip of the energy density. For
3121: SU(3) a similar stabilization takes place but with a much larger slope of the
3122: energy density: The phase transition appears to be
3123: weakly first order. It turns out that the electric
3124: $Z_2$ (SU(2)) or $Z_3$ (SU(3))
3125: degeneracy of the ground state as it was observed in the electric phase
3126: is lifted: A unique
3127: ground state characterizes
3128: the magnetic phase. On the other hand, the expectation of the
3129: Polyakov loop, though strongly suppressed as compared to that
3130: in the electric phase, does not vanish entirely
3131: in the magnetic phase. This is a manifestation
3132: of the fact that the ground state in the
3133: magnetic phase does allow for the propagation
3134: of massive dual gauge modes
3135: despite the confinement of fundamental, fermionic, and
3136: heavy test charges by the monopole condensates.
3137: In that sense, the magnetic phase is only preconfining.
3138:
3139: The critical
3140: behavior in the vicinity of $T_{c,E}$ is
3141: investigated and compared with results that seem to be
3142: related to the Yang-Mills theory by universality arguments.
3143:
3144: The monopole condensates are characterized by infinite
3145: correlation lengths. In a thermodynamical simulation on a finite-size
3146: lattice performed in the magnetic phase not much can be learned about thermodynamical
3147: quantities such as the pressure or the energy density which are sensitive to the infrared behavior of the
3148: theory.
3149:
3150:
3151: \subsection{Prerequisites}
3152:
3153: \subsubsection{The BPS monopole\label{BPSM}}
3154:
3155: Here we provide some facts about the 't Hooft-Polyakov
3156: monopole \cite{'tHooft1974,Polayakov1974} in the BPS limit \cite{PrasadSommerfield1975}
3157: since we will need them in
3158: Sec.\,\ref{defphismag}. We consider an SU(2) gauge model
3159: with the Lagrangian of Eq.\,(\ref{actiontotal})
3160: with the modification that the potential is absent.
3161: The BPS monopole is a
3162: static, particle-like solution to the equations of motion of this model
3163: saturating the BPS bound on
3164: the mass $M$. When centered at $\vec{x}=0$ it is given as
3165: %**************
3166: \eqb
3167: \label{BPSmonopole}
3168: \phi^a=\hat{x}_a|\phi|\,F(\rho)\,,\ \ \ \ a^a_4=0\,,\ \ \ \ \ \ \ a_i^a=\frac{{\epsilon}_{aij}}{er}\hat{x_j}\,G(\rho)\,,
3169: \eqe
3170: %************
3171: where $r\equiv |\vec{x}|$, $\hat{x}_i\equiv\frac{x_i}{r}$, $|\phi|\equiv\sqrt{\phi^a\phi^a}(r\to\infty)$, and
3172: $\rho=e|\phi|r$. The antimonopole solution is obtained by letting
3173: $\hat{x}\to -\hat{x}$ in Eq.\,(\ref{BPSmonopole}). The functions
3174: $F$ and $G$ can be determined analytically. They are given as \cite{PrasadSommerfield1975}
3175: %*********
3176: \eqb
3177: \label{FG}
3178: F(\rho)=\coth\rho-\frac{1}{\rho}\,,\ \ \ \ \ \ \ \ \ \ \ \ G(\rho)=1-\frac{\rho}{\sinh\rho}\,.
3179: \eqe
3180: %**********
3181: The mass $M$ of a BPS monopole or antimonopole calculates as
3182: %******
3183: \eqb
3184: \label{BPSmass}
3185: M=\frac{4\pi}{e}|\phi|\,.
3186: \eqe
3187: %******
3188: (Notice that in Eq.\,(\ref{BPSmass}) $|\phi|$ is replaced by $u\sim u_{\tiny\mbox{max}}=\frac{\pi}{\beta}$
3189: for a screened BPS monopole generated by the dissociation of a large-holonomy caloron, compare
3190: with Eq.\,(\ref{Vmu}).) The magnetic charge $g$ is obtained by integrating the 't Hooft tensor
3191: %********
3192: \eqb
3193: \label{tHoofttensor}
3194: {\cal F}_{\mu\nu}=\pd_\mu(\hat{\phi}^a a_\nu^a)-\pd_\nu(\hat{\phi}^a
3195: a_\mu^a)-\frac{1}{e}\epsilon^{abc}\hat{\phi}^a\pd_\mu\hat{\phi}^b\pd_\nu\hat{\phi}^c\,,\ \ \ \ \ (\hat{\phi^a}=
3196: \frac{\phi^a}{\sqrt{\phi^b\phi^b}})\,,
3197: \eqe
3198: %********
3199: over a two-sphere $S_2$ with infinite radius which is centered at $\vec{x}=0$.
3200: It is given as
3201: %*******
3202: \eqb
3203: \label{chargeg}
3204: g=\int_{S_2,R=\infty}\,d\Sigma_{\mu\nu}\,{\cal F}_{\mu\nu}=\pm\frac{4\pi}{e}\,.
3205: \eqe
3206: %********
3207: In Eq.\,(\ref{chargeg}) $d\Sigma_{\mu\nu}$ denotes the differential surface element.
3208:
3209: In unitary gauge, where the color orientation of $\phi^a$ is `combed' into a fixed
3210: direction in adjoint color space, the magnetic field
3211: $B_i=\mp \frac{\hat{x}_i}{er^2}$ associated with Eq.\,(\ref{tHoofttensor})
3212: is accompanied by a Dirac string
3213: along this direction in position space:
3214: If the monopole lies inside an $S_2$
3215: then the magnetic flux through this $S_2$ of the Dirac string
3216: precisely cancels that of the hedge-hog magnetic field. If a monopole or antimonopole lies
3217: outside of an $S_2$ with infinite radius an finite distance $b$ away from its surface
3218: and the monopole's or the antimonopole's Dirac string does not pierce the surface then the magnetic flux $F$
3219: through the surface, see Fig.\,\ref{Fig30}, calculates as
3220: %********
3221: \eab
3222: \label{magfluxs2}
3223: F&=&\int_{\tiny\mbox{plane}}d\Sigma_{\mu\nu}\,{\cal F}_{\mu\nu}\nonumber\\
3224: &=&\pm\frac{1}{e}\int_0^{2\pi}d\beta\int_{-\infty}^{\infty}dx\,|x|\cos\alpha(x,b)\frac{1}{x^2+b^2}\nonumber\\
3225: &=&\pm\frac{4\pi}{e}\,b\int_{0}^{\infty}dx\,\frac{|x|}{(x^2+b^2)^{3/2}}\nonumber\\
3226: &=&\pm\frac{4\pi}{e}\,.
3227: \eae
3228: %*********
3229: We conclude that static monopoles or antimonopoles lying inside an $S_2$ of
3230: {\sl infinite} radius do not contribute to the flux through $S_2$
3231: while they may contribute to the flux when situated
3232: outside of this $S_2$.
3233: %***********************
3234: \begin{figure}
3235: \begin{center}
3236: \leavevmode
3237: %\epsfxsize=9.cm
3238: \leavevmode
3239: %\epsffile[80 25 534 344]{}
3240: \vspace{4.9cm}
3241: \special{psfile=Fig-16.ps angle=0 voffset=-140
3242: hoffset=-120 hscale=40 vscale=40}
3243: \end{center}
3244: \caption{A BPS monopole in unitary gauge outside of an $S_2$ with
3245: infinite radius. The Dirac string does not pierce the surface of the $S_2$.
3246: \label{Fig30}}
3247: \end{figure}
3248: %************************
3249:
3250:
3251: \subsubsection{Derivation of the phases of
3252: macroscopic complex scalar fields\label{defphismag}}
3253:
3254: There is one species of magnetic monopoles in the SU(2) case
3255: while there are two independent species
3256: of magnetic monopoles for SU(3).
3257: \vspace{0.1cm}\\
3258: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
3259: In unitary gauge we consider an isolated system of a monopole and an antimonopole, which both are
3260: at rest and do not interact, outside of an $S_2$ with {\sl infinite} radius.
3261: We characterize their Dirac strings by
3262: unit vectors $\hat{x}_m$ and $\hat{x}_a$ which point away
3263: from the core of the monopole and the antimonopole, respectively.
3264: Let $\Pl$ be the plane perpendicular to $S_2$ such that the intersection line
3265: $L=\Pl\cap S_2$ coincides with the intersection line of $S_2$ with
3266: the plane spanned by $\hat{x}_m$ and $\hat{x}_a$. (The case where $\hat{x}_m$ and $\hat{x}_a$ lie
3267: in $S_2$ is inessential for what follows.) Whether or not this system contributes
3268: to the magnetic flux through $S_2$ depends on the angle
3269: $\delta=\angle(\hat{x}_m,\hat{x}_a)$ and on the angle
3270: $\gamma$ which the projection of $\hat{x}_m$ onto $\Pl$
3271: forms with $L$.
3272:
3273: A magnetic flux through $S_2$ is generated if and only
3274: if either $\hat{x}_m$ or $\hat{x}_a$ {\sl alone} pierces $S_2$.
3275: For a given angle $\delta$ the angle $\gamma$
3276: is uniformly distributed. In the absence of a heat bath
3277: the probability of measuring a flux $\frac{4\pi}{e}$ or
3278: a flux $-\frac{4\pi}{e}$ through $S_2$ thus is given as
3279: $\frac{\delta}{2\pi}$. We conclude that for a given angle $\delta$ the average plus or minus
3280: flux through $S_2$ reads
3281: %********
3282: \eqb
3283: \label{avflux}
3284: \bar{F}_{\pm}=\pm\frac{\delta}{2\pi}\frac{4\pi}{e}=\pm\frac{2\delta}{e}\,,\ \ \ \ \ \ (0\le\delta\le\pi)\,.
3285: \eqe
3286: %********
3287: Notice that $\delta$ and $2\pi-\delta$ generate the same average flux $\bar{F}_{\pm}$,
3288: thus the restriction $0\le\delta\le\pi$ in Eq.\,(\ref{avflux}).
3289:
3290: So far we have discussed the flux through $S_2$ which is
3291: generated by an isolated monopole-antimonopole system with no interactions.
3292: To derive the phase of the macroscopic complex field $\varphi$ describing the Bose
3293: condensate of such systems we couple the system to the heat bath,
3294: project onto zero-momentum states (condensate) of the monopole-antimonopole
3295: system such that each constituent does not carry momentum
3296: and perform the massless limit $e\to\infty$ which takes place for $T\le T_{c,E}$, compare
3297: with Eq.\,(\ref{divloglae}). (The rare case of a zero-momentum
3298: state with opposite and finite momenta of the constituents generates a
3299: closed and instable magnetic flux line, see Fig.\,\ref{Fig0}.
3300: This situation takes place if a large number of large-holonomy
3301: calorons dissociate into monopole-antimonopole pairs
3302: almost simultaneously inside a small
3303: spatial volume. On the macroscopic level, the thermal average over these flux loops,
3304: which collapse as soon as they are created, will later be described by a pure gauge configuration.)
3305:
3306: The monopole and antimonopole
3307: are generated by the dissociation of a large-holonomy caloron. According to
3308: Eq.\,(\ref{massesMonLL}) the sum of
3309: monopole and antimonopole mass, $M_{m+a}$, is, after screening, given as
3310: %**********
3311: \eqb
3312: \label{massesma}
3313: M_{m+a}=\frac{8\pi^2}{e\beta}\,.
3314: \eqe
3315: %**********
3316: The thermally averaged flux of the zero-momentum system at
3317: finite coupling $e$ is obtained as
3318: %*****
3319: \eab
3320: \label{avfluxsys}
3321: \bar{F}_{\pm,\tiny\mbox{th}}(\delta)&=&\,4\pi\,\int d^3p\,\delta^{(3)}(\vec{p})\, n_B(\beta |E(\vec{p})|)\,\bar{F}_{\pm}\nonumber\\
3322: &=&\pm\frac{8\pi\,\delta}{e}\int d^3p\,\frac{\delta^{(3)}(\vec{p})}{\exp\left[\beta\sqrt{M_{m+a}^2+\vec{p}^2}\right]-1}\,.
3323: \eae
3324: %*********
3325: After setting $\vec{p}=0$ in $\left(\exp\left[\beta\sqrt{M_{a+b}^2+\vec{p}^2}\right]-1\right)$ and by
3326: appealing to Eq.\,(\ref{massesma}), the expansion of this term reads
3327: %*********
3328: \eqb
3329: \label{expexpon}
3330: \lim_{\vec{p}\to 0}\left(\exp\left[\beta\sqrt{M_{m+a}^2+\vec{p}^2}\right]-1\right)=\frac{8\pi^2}{e}\left(1+\frac{1}{2}\frac{8\pi^2}{e}+
3331: \frac{1}{6}\left(\frac{8\pi^2}{e}\right)^2+\cdots\right)\,.
3332: \eqe
3333: %********
3334: Appealing to Eq.\,(\ref{expexpon}), the limit $e\to\infty$ can
3335: now safely be performed in Eq.\,(\ref{avfluxsys}). We have
3336: %**********
3337: \eqb
3338: \label{avfluxsys,einf}
3339: \lim_{e\to\infty} \bar{F}_{\pm,\tiny\mbox{th}}(\delta)=\pm\frac{\delta}{\pi}\,,\ \ \ \ \ \ (0\le\delta\le\pi)\,.
3340: \eqe
3341: %**********
3342: The right-hand side of Eq.\,(\ref{avfluxsys,einf}) defines the
3343: argument of the complex and periodic
3344: function $f$ with
3345: %*********
3346: \eqb
3347: \label{fmag}
3348: f(\frac{\delta}{\pi})\equiv C\frac{\varphi}{|\varphi|}(\frac{\delta}{\pi})
3349: \eqe
3350: %*********
3351: where $C$ is an undetermined (complex) constant. (Recall, that $\delta$ is an angle). Since $f$'s argument
3352: was obtained by a projection onto zero spatial
3353: momentum the only admissible nontrivial periodic dependence is that on the
3354: Euclidean time $\tau$. Without restriction of generality
3355: we can thus set $\frac{\delta}{\pi}=\frac{\tau}{\beta}$.
3356: Since $f$ is periodic, $f(\tau=0)=f(\tau=\beta)$, it can be expanded into a
3357: Fourier series:
3358: %*********
3359: \eqb
3360: \label{FourSer}
3361: f(\frac{\tau}{\beta})=\sum_{n=-\infty}^{n=\infty} f_n \exp\left[2\pi i n\frac{\tau}{\beta}\right]\,
3362: \eqe
3363: %*********
3364: where $f_n$ are (complex) constants. According to Eq.\,(\ref{fmag}) $f\bar{f}=|C|^2$
3365: is constant in $\tau$. Thus the only possibility in Eq.\,(\ref{FourSer}) is $f_n=C\,\delta_{mn}$ {\sl or}
3366: $f_n=\bar{C}\,\delta_{(-m)n}$ for a
3367: fixed value of $m$. Moreover, only $m=1$ is allowed since the
3368: physical situation generating the continuous parameter $\delta$
3369: does not repeat itself for $0\le\delta\le\frac{\pi}{m}$,
3370: $\frac{\pi}{m}\le\delta\le\frac{2\pi}{m},\cdots$ if $m>1$ because no
3371: higher-charge monopoles exist. (Recall that only calorons with a
3372: large-holonomy and topological charge unity are allowed to contribute to
3373: the thermodynamics in the electric phase.) We conclude that the equation of motion satisfied by $f$ is:
3374: %*********
3375: \eqb
3376: \label{eomfm}
3377: \pd_\tau^2 f(\frac{\tau}{\beta})+\left(\frac{2\pi}{\beta}\right)^2\,f(\frac{\tau}{\beta})=0\,.
3378: \eqe
3379: %*********
3380:
3381: \subsubsection{Derivation of the modulus of
3382: macroscopic complex scalar fields\label{defphismagmod}}
3383:
3384: What about $\varphi$'s modulus? The reasoning is completely analoguous to that for the derivation of
3385: $\phi$'s modulus. First, since $\varphi$ is composed of massless, noninteracting
3386: monopoles being at rest its $\tau$ dependence ought to be BPS saturated.
3387: Second, for $\varphi$ to have the phase $\exp[\pm 2\pi i\frac{\tau}{\beta}]$
3388: the right-hand side of its BPS equation ought
3389: to be {\sl linear} in $\varphi$. Third, for the existence of smooth
3390: deformations $\beta\to\beta+\delta\beta$, subject to a
3391: perturbative expansion in $\frac{\delta\beta}{\beta}$ away from a
3392: phase boundary, the right-hand side of the BPS equation must
3393: be analytic. Fourth, we assume that a scale $\La_M$ is given externally.
3394: Fifth, no explicit dependence on $\beta$ may appear on the
3395: right-hand side of the BPS equation since at zero momentum the temperature dependence of the
3396: mass of a monopole cancels against the factor $\beta$ in the
3397: Boltzmann weight. The only possibility for the BPS equation satisfying these conditions is
3398: %*******
3399: \eqb
3400: \label{BPSvarphi}
3401: \pd_\tau \varphi=\pm i\frac{\La_M^3\,\varphi}{|\varphi|^2}=\pm i\frac{\La_M^3}{\bar{\varphi}}\,.
3402: \eqe
3403: %*******
3404: From Eq.\,(\ref{BPSvarphi}) it follows that $\varphi$'s modulus
3405: is given as
3406: %*******
3407: \eqb
3408: \label{BPSvarphimod}
3409: |\varphi|=\sqrt{\frac{\La_M^3\beta}{2\pi}}\,.
3410: \eqe
3411: %*******
3412: The right-hand side of Eq.\,(\ref{BPSvarphi}) defines the square root of the
3413: potential $V_M$. In the absence of interactions
3414: between (screened) monopoles the effective theory for $\varphi$ thus reads
3415: %**********
3416: \eqb \label{actionvarphi}
3417: S_{\varphi} = \int_0^\beta d\tau \int d^3x
3418: \left( \frac{1}{2}\,\overline{\partial_\tau \varphi} \partial_\tau \varphi +
3419: \frac{1}{2}\,\frac{\La_M^6}{\bar{\varphi}\varphi} \right) \,.
3420: \eqe
3421: %***********
3422: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
3423: Here we have two independent monopole species which
3424: do not interact. The situation for each species is
3425: completely analogous to the SU(2) case. The two macroscopic
3426: fields $\varphi_1$ and $\varphi_2$ both satisfy the
3427: BPS equation (\ref{BPSvarphi}) and are given as
3428: %**********
3429: \eqb
3430: \label{varphi12}
3431: \varphi_1(\tau)=\varphi_2(\tau)=\sqrt{\frac{\La_M^3\beta}{2\pi}}\,\exp\left[\pm 2\pi i\frac{\tau}{\beta}\right]\,.
3432: \eqe
3433: %**********
3434: In the absence of interactions
3435: between monopoles the effective theory for $\varphi_1,\varphi_2$ reads
3436: %**********
3437: \eqb
3438: \label{actionvarphi12}
3439: S_{\varphi} = \int_0^\beta d\tau \int d^3x
3440: \left( \frac{1}{2}\,\overline{\partial_\tau \varphi_1} \partial_\tau \varphi_1 +
3441: \frac{1}{2}\,\overline{\partial_\tau \varphi_2} \partial_\tau \varphi_2 +
3442: \frac{1}{2}\,\frac{\La_M^6}{\bar{\varphi_1}\varphi_1} + \frac{1}{2}\,
3443: \frac{\La_M^6}{\bar{\varphi_2}\varphi_2} \right) \,.
3444: \eqe
3445: %**********
3446:
3447: \subsection{A macroscopic ground state\label{gsmag}}
3448:
3449: In close analogy to the electric phase we now derive the full macroscopic
3450: ground-state dynamics. First, we establish the quantum mechanical and statistical
3451: inertness of the fields $\varphi$ (SU(2)) and $\varphi_1,\varphi_2$
3452: (SU(3)). Subsequently, we solve the equations of motion for the
3453: topologically trivial dual gauge field in these backgrounds to
3454: obtain pure-gauge configurations describing, on a macroscopic level, the interaction between monopoles in the
3455: ground state. It will turn out that the Polyakov loops when evaluated on
3456: these pure-gauge configurations are trivial: The electric $Z_2$ (SU(2)) and $Z_3$ (SU(3))
3457: degeneracies observed in the electric phase no longer exist. Thus the magnetic
3458: phase confines fundamental test charges.
3459:
3460: The ratios of the mass-squared of potential $\varphi$-field fluctuations with
3461: $|\varphi|^2$ and $T^2$ are
3462: %**************
3463: \eqb
3464: \label{qsflphiM}
3465: \frac{\pd^2_{|\varphi|} V_{M}(\varphi)}{|\varphi|^2}=6\,\lambda_M^3\,,\
3466: \ \ \ \ \ \ \ \frac{\pd^2_{|\varphi|} V_{M}(\varphi)}{T^2}=24\pi^2\,,
3467: \eqe
3468: %*********
3469: where $\lambda_M\equiv\frac{2\pi T}{\La_M}$. We will show below
3470: that $\lambda_M\ge 7.075$ (SU(2)) and $\lambda_M\ge 6.467$ (SU(3)).
3471: Thus $\varphi$-field (SU(2)) and $\varphi_1,\varphi_2$-field (SU(3))
3472: fluctuations neither exist on-shell nor off-shell.
3473:
3474: \subsubsection{Pure-gauge configurations\label{pgmag}}
3475:
3476: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
3477: The topologically trivial sector is coupled to $\varphi$ in a minimal fashion, and
3478: the following effective action arises
3479: %**********
3480: \eqb
3481: \label{effactM2}
3482: S=\int_0^{\beta}
3483: d\tau\int d^3x\,\left[\frac{1}{4}\,
3484: G^D_{\mu\nu}G^D_{\mu\nu}+
3485: \frac{1}{2}\overline{{\cal D}_{\mu}\varphi}
3486: {\cal D}_{\mu}\varphi+\frac{1}{2}\frac{\La_M^6}{\bar{\varphi}\varphi}\right]\,,
3487: \eqe
3488: %**********
3489: where $G^D_{\mu\nu}=\pd_\mu a^D_\nu-\pd_\nu a^D_\mu$ denotes the Abelian field strength
3490: of the dual gauge field $a^D_\mu$ and ${\cal D}_{\mu}\equiv \pd_\mu+ig\,a^D_\mu$
3491: denotes the covariant derivative involving the magnetic gauge coupling $g$.
3492:
3493: Since the field $\varphi$ does not fluctuate it is a background
3494: to the macroscopic gauge-field equations of motion which follows from Eq.\,(\ref{effactM2}):
3495: %*********
3496: \eqb
3497: \label{eomdualG2}
3498: \pd_\mu G^D_{\mu\nu}=ig\left[\overline{{\cal D}_{\nu}\varphi}\varphi-\bar{\varphi}
3499: {\cal D}_{\nu}\varphi\right]\,.
3500: \eqe
3501: %*********
3502: A pure-gauge solution to Eq.\,(\ref{eomdualG2}) with $D_\mu\varphi\equiv0$
3503: is given as
3504: %*******
3505: \eqb
3506: \label{pgsolM2}
3507: a^{D,bg}_{\mu}=\pm\delta_{\mu 4}\frac{2\pi}{g\beta}\,.
3508: \eqe
3509: %********
3510: On $\varphi$ and on $a^{D,bg}_{\mu}$ only the potential does not vanish in
3511: Eq.\,(\ref{effactM2}): Interactions
3512: between magnetic monopoles create a nonvanishing energy density $\rho^{gs}$
3513: and pressure $P^{gs}$ where
3514: %*********
3515: \eqb
3516: \label{Pmag2}
3517: \rho^{gs}=\pi\,\La_M^3\,T=-P^{gs}\,.
3518: \eqe
3519: %***********
3520: We shall see in Sec.\,\ref{ANOBPS}, compare with Eq.\,(\ref{Pvort}), that the negative
3521: ground-state pressure in Eq.\,(\ref{Pmag2}) originates from center-vortex
3522: loops which collapse as soon as they are created.
3523:
3524: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
3525: The situation is the same as for SU(2) except that
3526: we have gauge dynamics subject to U(1)$_D^2$ and not only
3527: U(1)$_D$. The effective action reads
3528: %**********
3529: \eab
3530: \label{effactM3}
3531: S&=&\int_0^{\beta}
3532: d\tau\int d^3x\,\left[\frac{1}{4}\,
3533: G^D_{\mu\nu,1}G^D_{\mu\nu,1}+\frac{1}{4}\,
3534: G^D_{\mu\nu,2}G^D_{\mu\nu,2}+\right.\nonumber\\
3535: &&\left.\frac{1}{2}\overline{{\cal D}_{\mu,1}\varphi_1}
3536: {\cal D}_{\mu,1}\varphi_1+\frac{1}{2}\overline{{\cal D}_{\mu,2}\varphi_2}
3537: {\cal D}_{\mu,2}\varphi_2+\frac{1}{2}\frac{\La_M^6}{\bar{\varphi_1}\varphi_1}
3538: +\frac{1}{2}\frac{\La_M^6}{\bar{\varphi_2}\varphi_2}\right]\,.
3539: \eae
3540: %**********
3541: The Abelian field strengths $G^D_{\mu\nu,1}, G^D_{\mu\nu,2}$
3542: and the covariant derivatives ${\cal D}_{\mu,1},{\cal D}_{\mu,2}$
3543: are defined as for the SU(2) case with the replacements $a^D_\mu\to a^D_{\mu,1},a^D_\mu\to a^D_{\mu,2}$,
3544: respectively. (The magnetic gauge coupling $g$ is universal since both species of
3545: monopoles couple with the same strength to their respective gauge field.)
3546:
3547: The equations of motion for the fields $a^D_{\mu,1},a^D_{\mu,2}$ in the background
3548: of the fields $\varphi_1,\varphi_2$ are
3549: %*********
3550: \eqb
3551: \label{eomdualG3}
3552: \pd_\mu G^D_{\mu\nu,1}=ig\left[\overline{{\cal D}_{\nu,1}\varphi_1}\varphi-\bar{\varphi_1}
3553: {\cal D}_{\nu,1}\varphi_1\right]\,,\ \
3554: \pd_\mu G^D_{\mu\nu,2}=ig\left[\overline{{\cal D}_{\nu,2}\varphi_2}\varphi-\bar{\varphi_2}
3555: {\cal D}_{\nu,2}\varphi_2\right]\,.
3556: \eqe
3557: %*********
3558: Pure-gauge solutions to these equations with
3559: ${\cal D}_{\nu,1}\varphi_1={\cal D}_{\nu,2}\varphi_2=0$ are given as
3560: %*******
3561: \eqb
3562: \label{pgsolM3}
3563: a^{D,bg}_{\mu,1}=\pm\delta_{\mu 4}\frac{2\pi}{g\beta}\,,\ \ \ \ \ \ \
3564: a^{D,bg}_{\mu,2}=\pm\delta_{\mu 4}\frac{2\pi}{g\beta}\,.
3565: \eqe
3566: %********
3567: On $\varphi_1,\varphi_2$ and on $a^{D,bg}_{\mu,1},a^{D,bg}_{\mu,2}$
3568: only the potentials do not vanish in Eq.\,(\ref{effactM3}): Interactions
3569: between magnetic monopoles create a nonvanishing energy density $\rho^{gs}$
3570: and pressure $P^{gs}$ where
3571: %*******
3572: \eqb
3573: \label{Pmag3}
3574: \rho^{gs}=2\pi\,\La_M^3\,T=-P^{gs}\,.
3575: \eqe
3576: %********
3577: Again, the negative ground-state pressure in Eq.\,(\ref{Pmag3}) originates from center-vortex
3578: loops which collapse as soon as they are created.
3579:
3580: \subsubsection{Polyakov loop and rotation to unitary gauge\label{PUGM}}
3581:
3582: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
3583: The Polyakov loop ${\bf P}\equiv \exp\left[ig\int_0^\beta d\tau\, a^D_4\right]$, when
3584: evaluated on the pure-gauge configuration in Eq.\,(\ref{pgsolM2}), reads
3585: %*******
3586: \eqb
3587: \label{pGPol2}
3588: {\bf P}=\exp[\pm ig\int_0^\beta d\tau\,\frac{2\pi}{g\beta} ]=1\,.
3589: \eqe
3590: %*******
3591: A gauge rotation $a_\mu^{D,bg}\to a_\mu^{D,bg}+\frac{i}{g}\left(\pd_\mu \Omega^\dagger\right)\Omega$
3592: to unitary gauge $\varphi=|\varphi|, a_\mu^{D,bg}=0$ is mediated by the U(1) group element
3593: $\Omega=\exp\left[\pm 2\pi i\frac{\tau}{\beta}\right]$: The Polyakov loop {\bf P} is invariant
3594: under this gauge transformation. We conclude that the electric
3595: $Z_2$ ground-state degeneracy, which was observed in the electric
3596: phase, no longer exists in the magnetic phase: The ground state confines
3597: fundamentally charged, heavy and fermionic test charges.\vspace{0.1cm}\\
3598: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
3599: Here the Polyakov loop ${\bf P}$ is a product of the Polyakov
3600: loops ${\bf P}_1$ and ${\bf P}_2$ computed on the respective pure-gauge
3601: configurations $a^{D,bg}_{\mu,1}$ and $a^{D,bg}_{\mu,1}$ in Eq.\,(\ref{pgsolM3}). We have
3602: %*******
3603: \eqb
3604: \label{pGPol3}
3605: {\bf P}={\bf P}_1{\bf P}_2=\exp[\pm 2ig\int_0^\beta d\tau\,\frac{2\pi}{g\beta} ]=1\,\ \ \ \ \mbox{or}\ \ \ \ \
3606: {\bf P}={\bf P}_1{\bf P}_2=\exp[0]=1\,.
3607: \eqe
3608: %*******
3609: Gauge rotations to unitary gauge
3610: $\varphi_1=|\varphi_1|=\varphi_2, a_{\mu,1}^{D,bg}=a_{\mu,2}^{D,bg}=0$ are
3611: mediated by the U(1) group elements
3612: $\Omega_1=\exp\left[\pm 2\pi i\frac{\tau}{\beta}\right]=\Omega_2$.
3613: Again, ${\bf P}$ is invariant under these gauge rotations. We conclude that the electric
3614: $Z_3$ ground-state degeneracy does not exist in the magnetic phase: Fundamentally charged, heavy and fermionic
3615: test charges are confined by the monopole condensates.
3616:
3617:
3618: \subsection{Excitations\label{ExcM}}
3619:
3620: \subsubsection{Mass spectrum of thermal quasiparticles\label{MSTQPM}}
3621:
3622: The dual Abelian Higgs mechanism generates tree-level
3623: quasiparticle masses $m$ and $m_1,m_2$ for the fluctuations $\delta a^D_{\mu}$ (SU(2))
3624: and $\delta a^D_{\mu,1},\delta a^D_{\mu,2}$ (SU(3)), respectively. We have
3625: %********
3626: \eqb
3627: \label{massspecM}
3628: m=g|\varphi|=m_1=m_2=a\,T\,,\ \ \ \ \ \ \ (a\equiv 2\pi\,g\,\lambda_M^{-3/2})\,
3629: \eqe
3630: %*********
3631: where $\lambda_M\equiv\frac{2\pi T}{\La_M}$.
3632:
3633: \subsubsection{Thermodynamical selfconsistency and evolution equation\label{TSCevM}}
3634:
3635: Due to the absence of interactions between the dual gauge fields
3636: the thermodynamics of the magnetic phase is exact on the one-loop level.
3637: Again, a magnetic modification of the compositeness condition Eq.\,(\ref{comconnoi}) applies.
3638:
3639: Let us first compare the contribution $\Delta V_M$ of
3640: quantum fluctuations to the pressure arising from dual gauge modes with the tree-level
3641: result $-1/2\,V_M=-\pi\,\La_M^3 T$ (SU(2)) and $-1/2\,V_M=-2\pi\,\La_M^3 T$ (SU(3)).
3642: In both cases we have
3643: %**********
3644: \eqb
3645: \label{deltatreeM}
3646: \frac{\Delta V_M}{V_M}=\frac{\lambda_M^{-3}}{24\pi^2}\,.
3647: \eqe
3648: %**********
3649: Considering that $\lambda_M\ge 7.075$ (SU(2)) and $\lambda_M\ge 6.467$ (SU(3)) this
3650: is smaller than $1.2\times 10^{-5}$ and $1.6\times 10^{-5}$, respectively.
3651: Thus the quantum contribution to the
3652: one-loop pressure can safely be neglected.
3653:
3654: For SU(2) the thermal contribution to the pressure reads
3655: %*********
3656: \eqb
3657: \label{Pre2M}
3658: P(\lambda_M)=-\La_M^4\left[\frac{6\lambda_M^4}{(2\pi)^6}\bar{P}(a)+\frac{\lambda_M}{2}\right]\,
3659: \eqe
3660: %*******
3661: where the (negative) function $\bar{P}(a)$ is defined in Eq.\,(\ref{P(y)}). The SU(3)
3662: pressure is just twice the SU(2) pressure. From the condition $\pd_a P=0$ of
3663: thermodynamical selfconsistency the following evolution equation arises for both SU(2) and SU(3):
3664: %********
3665: \eqb
3666: \label{evol23M}
3667: \pd_a\lambda_M=-\frac{12\lambda_M^4\,a}{(2\pi)^6}\,\frac{D(a)}{1+\frac{12\lambda_M^3\,a^2}{(2\pi)^6}\,D(a)}
3668: \eqe
3669: %**********
3670: where the (positive) function $D(a)$ is defined in Eq.\,(\ref{DA}).
3671: In analogy to the electric phase the $\lambda_M$ dependence of the gauge coupling constants $g$
3672: is obtained by inverting the solutions to Eq.\,(\ref{evol23M})
3673: and by subsequently using the relation between $g$, $\lambda_M$ and $a$ in
3674: Eq.\,(\ref{massspecM}). The temperature evolution of $g$ is
3675: shown in Fig.\,\ref{gevol}.
3676: %***********************
3677: \begin{figure}
3678: \begin{center}
3679: \leavevmode
3680: %\epsfxsize=9.cm
3681: \leavevmode
3682: %\epsffile[80 25 534 344]{}
3683: \vspace{5.0cm}
3684: \special{psfile=Fig-17.ps angle=0 voffset=-155
3685: hoffset=-140 hscale=60 vscale=50}
3686: \end{center}
3687: \caption{The evolution of the effective gauge coupling $g$ in the magnetic phase for
3688: SU(2) (thick grey line) and SU(3) (thick black line). At
3689: $\lambda_{c,M}=7.075$ (SU(2)) and $\lambda_{c,M}=6.467$ (SU(3)) $g$ diverges logarithmically,
3690: $g\sim -\log(\lambda_{M}-\lambda_{c,M})$. \label{gevol}}
3691: \end{figure}
3692: %************************
3693:
3694: \subsubsection{Interpretation of results\label{intM}}
3695:
3696: The magnetic gauge coupling $g$ increases continuously from $g=0$ at the electric-magnetic
3697: phase boundary ($T=T_{c,E}$) until it diverges logarithmically at $T_{c,M}$.
3698: A variation of the magnetic
3699: coupling with temperature is not in contradiction with magnetic charge conservation
3700: since no {\sl isolated} magnetic charges appear
3701: in the magnetic phase: Magnetic monopoles either are condensed
3702: or they conspire to form instable magnetic flux loops (center-vortex loops),
3703: see Fig.\,\ref{Fig0}. From Fig.\,\ref{gevol} one can see that the
3704: magnetic phase is more narrow for SU(3) than it is for SU(2). Despite the fact
3705: that the ground-state degeneracies with respect to the electric $Z_2$ symmetry (SU(2)) and
3706: the electric $Z_3$ symmetry (SU(3)) are absent in the magnetic phase
3707: the fully averaged Polyakov loop (including the massive dual gauge-mode excitations)
3708: does not vanish at $T_{c,E}$, see Sec.\,\ref{polyaloop}. The expectation of the Polyakov loops only
3709: vanishes at $T_{c,M}$ where {\sl all} dual gauge modes are decoupled because of a
3710: diverging mass. If we take the mass of the dual
3711: gauge modes to be the order parameter for the dynamical
3712: breaking of U(1)$_D$ (SU(2)) and U(1)$_D^2$ (SU(3))
3713: then the electric-magnetic transition is second order with
3714: mean-field exponents in both cases. The best one
3715: can do to relate this order-parameter to electric
3716: $Z_2$ or $Z_3$ restoration is to look for the point
3717: where its exponent is least sensitive to the length $\Delta\lambda_M$ of the
3718: fitting interval. The expectation of the 't Hooft loop, which is a dual order parameter for complete
3719: confinement and which is nonzero if center-vortices are condensed,
3720: vanishes inside the magnetic phase where center-vortex loops
3721: are isolated and instable defects. At $T_{c,M}$ the expectation of the 't Hooft loop
3722: jumps to a finite value. The associated transition
3723: is, however, neither second nor first order
3724: but of the Hagedorn type see Sec.\,\ref{Hagedorn}. Because of the
3725: infinite correlation length (massless condensed magnetic monopoles)
3726: a finite-size lattice simulation of the order parameter as well as infrared sensitive
3727: thermodynamical quantities such as the pressure is problematic.
3728:
3729:
3730: \subsection{Polyakov loop in the electric and the magnetic phase\label{polyaloop}}
3731:
3732: In this section we show, on a qualitative level, that the expectation of the
3733: Polyakov loop $\la{\bf P}\ra$, which is an order parameter
3734: for the confinement-deconfinement transition, is finite both in the electric
3735: phase and the magnetic phase. Deep in the magnetic phase $\la{\bf P}\ra$ is, however,
3736: strongly suppressed as compared to its value at $T_{c,E}$.
3737:
3738: In each phase the Polyakov-loop expectation $\la{\bf P}\ra$
3739: can be computed in unitary(-Coulomb) gauge. Let us first discuss the electric phase.
3740: The sector $S_{f,E}$ in the effective action
3741: (\ref{actiontotal}), which involves fluctuating fields, is given
3742: as
3743: %********
3744: \eqb
3745: \label{SfE}
3746: S_{f,E}=\int_0^{\beta}d\tau\,d^3x\left[\frac14\,G^a_{\mu\nu}G^a_{\mu\nu}+
3747: \frac12\sum_a m_a^2\,(\delta a^a_\mu)^2\right]\,.
3748: \eqe
3749: %********
3750: Here $a=1,2,3$ and $m_1=m_2>0, m_3=0$ for SU(2) and $a=1,\cdots,8$
3751: and $m_1=m_2=\frac{1}{2}m_3=\frac{1}{2}m_4=\frac{1}{2}m_5=\frac{1}{2}m_6>0,
3752: m_7=m_8=0$ for SU(3). Since in unitary gauge the Polyakov loop is
3753: unity in the ground state no direct ground-state contribution arises
3754: in the expectation $\la{\bf P}\ra$: The associated factor in the numerator
3755: cancels that in the denominator. We have
3756: %***********
3757: \eqb
3758: \label{Ployeff}
3759: \la{\bf P}\ra=Z_{f,E}^{-1}\times
3760: \int \left[d\delta a_\mu\right]\,\exp\left[ie\int_0^{\beta} d\tau\,\delta a_4\right] \exp[-S_{f,E}]\,,
3761: \eqe
3762: %**********
3763: where $\left[d\delta a_\mu\right]$ denotes the path-integral measure and
3764: $Z_{f,E}\equiv \int \left[d\delta a_\mu\right] \exp[-S_{f,E}]$.
3765:
3766: Since the fluctuations $\delta a_\mu$
3767: are periodic in $\tau$ they can be decomposed into a Matsubara sum:
3768: %********
3769: \eqb
3770: \label{fourierb}
3771: \delta a_\mu(\tau,{\vec x})=\sum_{n=-\infty}^{n=\infty}
3772: \exp\left[2\pi in \frac{\tau}{\beta}\right]\,\delta\bar{a}_{\mu,n}({\vec x})\,.
3773: \eqe
3774: %********
3775: Modes with $n\not=0$ in Eq.\,(\ref{fourierb}) render the Polyakov-loop
3776: phase in Eq.\,(\ref{Ployeff}) to be unity and are
3777: action-suppressed. Zero modes ($n=0$) contribute
3778: to $\la{\bf P}\ra$ sizably
3779: if they are not action-suppressed. This is the case if and only if
3780: both of the following conditions are met: (i) as compared to $T$ some or all masses
3781: $m_a$ in Eq.\,(\ref{SfE}) are small and (ii)
3782: $\pd_i\delta\bar{a}_{\mu,0}({\vec x})$ is small compared with $T^2$ and the field configuration
3783: is still localized in space. Here $\pd_i$ denotes a spatial derivative. In the electric
3784: phase there are massless modes, the conditions (i) and (ii) can be satisfied, and
3785: thus a finite Polyakov-loop expectation emerges.
3786:
3787: A similar consideration can be performed for the magnetic phase.
3788: Deep inside this phase condition (i) is badly violated since the mass of {\sl all} dual
3789: gauge modes is much larger than $T$ by virtue of Fig.\,\ref{gevol}
3790: and the mass formula in Eq.\,(\ref{massspecM}). Thus $\la{\bf P}\ra$, though nonvanishing, is strongly
3791: suppressed deep inside the magnetic phase as compared
3792: to its value at $T_{c,E}$. For quantitative results the average $\la{\bf P}\ra$ can be performed
3793: on a lattice or analytically by using the respective
3794: effective theory for the electric and the magnetic phase.
3795:
3796:
3797: \subsection{Critical behavior at the electric-magnetic transition\label{critexp}}
3798:
3799: The electric-magnetic transition, which goes with the dynamical breakdown
3800: of U(1)$_D$ (SU(2)) and U(1)$^2_D$ (SU(3)), is second
3801: order in both cases. The difference is that
3802: in the SU(3) case the magnetic phase is more narrow than for SU(2),
3803: see Fig.\,\ref{gevol}, and that the peak-value
3804: of the specific heat per volume is much larger for SU(3) than it is for SU(2),
3805: see Fig.\,\ref{SH}. In addition, the entropy density, which measures
3806: the mobility of dual gauge modes and which is used to extract an
3807: apparent latent heat on the lattice \cite{Brown1988},
3808: drops much more rapidly in the magnetic phase for SU(3) as compared to the SU(2) case.
3809: This is a plausible explanation for the apparent first-order nature of the confinement-deconfinement
3810: transition observed in lattice simulations \cite{LuciniTeperWenger2003,LuciniTeperWenger2005}.
3811:
3812: The order parameter for the electric-magnetic transition is the
3813: mass $\propto a\lambda_M$ of the dual gauge bosons. (The monopole mass vanishes
3814: like an inverse {\sl logarithm} on the electric side of the transition, and thus it is not an order parameter.)
3815: The following model applies to the
3816: behavior of $a\,\lambda_M$ close to the critical temperature $\lambda_{c}=9.24$ (SU(2))
3817: and $\lambda_{c}=6.81$ (SU(3)):
3818: %********
3819: \eqb
3820: \label{modeldata}
3821: a\,\lambda_M (\lambda_M)=K\,|\lambda_M-\lambda_{c}|^\nu\,,
3822: \eqe
3823: %********
3824: where $K$ and $\nu$ are constants, and $\lambda_{c}$ is the critical
3825: temperature $T_{c,E}$ in units of $\frac{\Lambda_M}{2\pi}$. By demanding continuity of
3826: the pressure across the electric-magnetic transition, see Sec.\,\ref{MCC},
3827: we derive $\lambda_{c}=8.478$ (SU(2)) and $\lambda_{c}=7.376$ (SU(3)).
3828:
3829: The magnetic phase is not
3830: entirely confining (dual gauge bosons propagate although fundamental, heavy,
3831: and fermionic test charges are confined), and thus the expectation of the Polyakov
3832: loop $\la{\bf P}\ra$ is not exactly zero. The best one can do in order
3833: to compare the behavior of $a\,\lambda_M$ to the behavior of $\la{\bf P}\ra$ inferred
3834: from universality arguments \cite{SvetitskyYaffe1982-1,SvetitskyYaffe1982-2}
3835: is to look for the point where the fitted value of $\nu$
3836: in Eq.\,(\ref{modeldata}) is least sensitive to a variation of the length $\Delta\lambda_M$
3837: of the fitting interval.
3838:
3839: To perform the fit to the model in Eq.\,(\ref{modeldata}) we
3840: have used Mathematica's NonlinearFit function which is contained in the
3841: statistics package. The function $a(\lambda_M)$, subject to the
3842: initial conditions $a(\lambda_{c}=8.478)=0$ (SU(2)) and $a(\lambda_{c}=7.376)=0$
3843: (SU(3)), was generated by an inversion of the corresponding numerical
3844: solutions to Eq.\,(\ref{evol23M}). (A step-size $\delta a=5\times 10^{-9}$
3845: was used in the Runge-Kutta algorithm\footnote{The author would like to thank
3846: Jochen Rohrer for performing the numerical calculation.}.) In Fig.\,\ref{expfit} the
3847: fitted exponent $\nu$ is shown as a function of $\Delta\lambda_M$.
3848: %***********************
3849: \begin{figure}
3850: \begin{center}
3851: \leavevmode
3852: %\epsfxsize=9.cm
3853: \leavevmode
3854: %\epsffile[80 25 534 344]{}
3855: \vspace{4.5cm}
3856: \special{psfile=Fig-18.ps angle=0 voffset=-140
3857: hoffset=-195 hscale=50 vscale=50}
3858: \end{center}
3859: \caption{The critical exponent $\nu$ for the mass of the dual gauge modes and for
3860: the magnetic-electric transition as a function of the length $\Delta\lambda_M$
3861: of the fitting interval. The left panel is the SU(2) and the
3862: right panel the SU(3) result. \label{expfit}}
3863: \end{figure}
3864: %************************
3865: Two things are important to observe. First, the magnetic-electric
3866: transition is second order with the mean-field
3867: exponent $\nu=0.5$ for both SU(2) and SU(3). Second, for each case there
3868: is a point $\Delta\lambda^*_M$ of least sensitivity for $\nu$ under
3869: variations in $\Delta\lambda_M$. For SU(2) we have $\Delta\lambda^*_M=0.28 \pm 0.03$ and
3870: $\nu(\Delta\lambda^*_M)=0.576\pm 0.008$, and
3871: for SU(3) $\Delta\lambda^*_M=0.16 \pm 0.02$ and
3872: $\nu(\Delta\lambda^*_M)=0.572 \pm 0.006$.
3873: By universality we expect $\nu(\Delta\lambda^*_M)$ for SU(2) to
3874: be close to the exponent $\nu_{\tiny\mbox{IM}}$ for the corresponding order
3875: parameter of a 3D Ising model \cite{SvetitskyYaffe1982-1,SvetitskyYaffe1982-2}.
3876: One has $\nu_{\tiny\mbox{IM}}\sim 0.63$. The SU(2) exponent $\nu(\Delta\lambda^*_M)$ only
3877: deviates by about 8.5\% from $\nu_{\tiny\mbox{IM}}$.
3878:
3879:
3880: \section{The center phase\label{CVCM}}
3881:
3882: \subsection{Prerequisites}
3883:
3884: \subsubsection{ANO vortex in the BPS limit\label{ANOBPS}}
3885:
3886: Just like the isolated defects in the electric phase
3887: are screened BPS monopoles, the isolated defects in
3888: the magnetic phase are screened and closed
3889: magnetic flux lines (vortex-loops). In Fig.\,\ref{Fig0}, see also \cite{Olejnik1997}, we have given
3890: a figurative interpretation of these flux
3891: lines: They are composed of magnetic monopoles and
3892: antimonopoles which move into opposite directions. Thus there is a net magnetic
3893: current in the vortex core. A vortex core can
3894: be viewed as locations in space where U(1)$_D$ (SU(2)) or one of the factors in
3895: U(1)$^2_D$ (SU(3)) are restored. Hence the picture of
3896: isolated monopoles, contributing to the magnetic current, applies to the vortex core.
3897:
3898: We have explained in Sec.\,\ref{intro} why the magnetic flux carried by a vortex-loop
3899: is independent of the state of motion of a
3900: particular monopole contributing
3901: to the vortex: The amount of flux carried by a vortex solely is a function of the charge
3902: of each BPS monopole contributing to it. Because
3903: large-holonomy calorons of topological charge one and
3904: thus monopoles with one unit of magnetic charge only are
3905: thermodynamically excited Abrikosov-Nielsen-Olesen (ANO) vortices are
3906: center vortices in the magnetic phase of an SU(2) or an SU(3)
3907: Yang-Mills theory.
3908:
3909: A mesoscopic description of a static ANO vortex in the BPS limit
3910: is given by the action Eq.\,(\ref{effactM2}) when omitting
3911: the potential for $\varphi$. (This potential measures the
3912: energy density of the {\sl macroscopic} ground state and thus must
3913: be subtracted when discussing the typical energy of a
3914: solitonic configuration on a mesoscopic level.)
3915: The following cylindrically symmetric (with axis along the $x_3$
3916: direction) and static ansatz for the gauge field $a^D_\mu$ is made to describe the vortex
3917: \cite{NielsenOlesen1973}:
3918: %*******
3919: \eqb
3920: \label{ansatzaM}
3921: a^D_4=0\,,\ \ \ \ \ \ \ \ a^D_i=\epsilon_{ijk}\hat{r}_j \,e_k A(r)\,,
3922: \eqe
3923: %*******
3924: where $\hat{r}$ is a radial unit vector in the $x_1x_2$ plane, and $\vec e$ is a
3925: unit vector along the $x_3$ direction. Writing
3926: $\varphi=|\varphi|(r)\exp[i\theta]$, the equations
3927: of motion for $|\varphi|(r)$ and $A(r)$ read
3928: %**********
3929: \eab
3930: -\frac{1}{r}\frac{d}{dr}\left(r\frac{d}{dr}|\varphi|\right)+
3931: \left(\frac{1}{r}-g\,A\right)^2|\varphi|&=&0\,,\label{phir}\\
3932: -\frac{d}{dr}\left(\frac{1}{r}\frac{d}{dr}(rA)\right)+
3933: g\,|\varphi|^2\left(g\,A-\frac{1}{r}\right)&=&0\,.\label{Ar}
3934: \eae
3935: %***********
3936: We keep in mind that $|\varphi|(r\to\infty)=\sqrt{\frac{\La_M^3\beta}{2\pi}}$,
3937: see Eq.\,(\ref{BPSvarphimod}). Let us first search for BPS saturated solutions to the system
3938: (\ref{phir},\ref{Ar}). The question is under what
3939: condition a solution to the first-order equation
3940: %*********
3941: \eqb
3942: \label{fovarphi}
3943: \frac{d}{dr}|\varphi|=\left(\frac{1}{r}-g\,A\right)|\varphi|
3944: \eqe
3945: %*********
3946: also solves Eq.\,(\ref{phir}). Substituting Eq.\,(\ref{fovarphi})
3947: into Eq.\,(\ref{phir}), we observe that
3948: %*********
3949: \eqb
3950: \label{foA}
3951: A=-r\frac{d}{dr}\,A\,.
3952: \eqe
3953: %********
3954: The solution to Eq.\,(\ref{foA}) is
3955: %*********
3956: \eqb
3957: \label{solfoA}
3958: A(r)=\frac{\mbox{const}}{r}\,.
3959: \eqe
3960: %*********
3961: Substituting Eq.\,(\ref{solfoA}) into Eq.\,(\ref{Ar}),
3962: the constant in Eq.\,(\ref{solfoA}) is determined to be
3963: $\frac{1}{g}$. Eq.\,(\ref{fovarphi}) is solved for $r>0$ by
3964: %**********
3965: \eqb
3966: \label{fovarphisol}
3967: |\varphi|(r)\equiv\sqrt{\frac{\La_M^3\beta}{2\pi}}\,.
3968: \eqe
3969: %***********
3970: We have found an analytical solution to the system
3971: (\ref{phir},\ref{Ar}) for $r>0$ which has one unit of
3972: magnetic flux $F_v(r)\equiv\frac{2\pi}{g}=\oint_C dz_\mu\,a^D_{\mu}$, where $C$
3973: is a circular curve of radius $r$ in the $x_1 x_2$ plane
3974: centered at $x_1=x_2=0$, and which has a vanishing vortex core.
3975: The energy density $\rho_v(r)$, when evaluated
3976: on the configuration $A(r)=\frac{1}{g\,r}$, $|\varphi|(r)\equiv\sqrt{\frac{\La_M^3\beta}{2\pi}}$,
3977: reduces to that of the magnetic field $H(r)=\frac{1}{2\pi r}\frac{d}{dr}\,F_v(r)$:
3978: (By Stoke's theorem the magnetic field $H$ must be proportional to a
3979: two-dimensional $\delta$-function at $r=0$. Thus the energy per unit
3980: vortex length diverges on the configuration (\ref{solfoA}) and (\ref{fovarphisol}).)
3981: %*********
3982: \eqb
3983: \label{endensmag}
3984: \rho_v(r)=\frac12\,H^2(r)\equiv 0\,,\ \ \ \ \ \ (r>0)\,.
3985: \eqe
3986: %*********
3987: The (isotropic in the $x_1x_2$ plane) pressure
3988: $P_v(r)$ outside the vortex core is given as
3989: %**********
3990: \eqb
3991: \label{Pvort}
3992: P_v(r)=-\frac12\,\frac{\La_M^3\beta}{2\pi}\,\frac{1}{g^2\,r^2}\,,\ \ \ \ \ \ \ \ (r>0)\,.
3993: \eqe
3994: %***********
3995: Eq.\,(\ref{Pvort}) is the mesoscopic reason for the macroscopic
3996: results in Eqs.\,(\ref{Pmag2}) and (\ref{Pmag3}). Because of the {\sl negative} pressure in
3997: Eq.\,(\ref{Pvort}) vortex-{\sl loops} start to collapse as soon as
3998: they are created at finite coupling $g$. (The pressure is more negative inside
3999: than outside of the vortex-loop.) Notice that in the limit $g\to\infty$ we have
4000: $P_v(r)\to 0$: For temperatures below $T_{c,M}$
4001: vortex-loops do exist as particle-like excitations.
4002:
4003: \subsubsection{Leaving the BPS limit}
4004:
4005: Let us now discuss how the solutions in Eqs.\,(\ref{fovarphisol}) and
4006: (\ref{solfoA}) are deformed when the BPS limit is left at
4007: finite coupling $g$. In this case only approximate analytical
4008: solutions to the second-order system (\ref{phir}) and (\ref{Ar})
4009: are known \cite{NielsenOlesen1973}. Assuming $|\varphi|$ to be constant,
4010: which is viable sufficiently far away from the core around $r=0$, the
4011: solution to Eq.\,(\ref{Ar}) reads
4012: %**********
4013: \eqb
4014: \label{arnoBPS}
4015: A(r)=\frac{1}{gr}-|\varphi|\,K_1(g|\varphi|r)\to \frac{1}{gr}-|\varphi|\sqrt{\frac{\pi}{2g|\varphi|r}}\,
4016: \exp[-g|\varphi|r]\,,\ \ \ \ (r\to\infty)\,,
4017: \eqe
4018: %***********
4019: where $K_1$ is a modified Bessel function. (Notice that the $1/r$ divergence at $r=0$
4020: of the solution in Eq.\,(\ref{solfoA}) is absent in the configuration in Eq.\,(\ref{arnoBPS}).)
4021: Now $|\varphi|$ is not constant inside the vortex
4022: core but smoothly approaches zero for $r\to 0$. So there
4023: is a gradient contribution from $|\varphi|$ to the energy per unit length $\frac{E_v}{2\pi R}$
4024: along the vortex where $R$ denote the typical radius
4025: of a vortex loop. Let us first calculate the magnetic energy per unit length $\frac{E_{m,v}}{2\pi R}$.
4026: One has \cite{NielsenOlesen1973}
4027: %*******
4028: \eqb
4029: \label{magenANO}
4030: \frac{E_{m,v}}{2\pi R}=\frac12\,\int_0^\infty dr\,2\pi r H^2(r)=
4031: \pi\,|\varphi|^2\int_0^\infty dy\,K^2_0(y)\,y=\frac{\pi}{2}\,|\varphi|^2\,,
4032: \eqe
4033: %********
4034: where $|\varphi|$ is given in Eq.\,(\ref{fovarphisol}). The gradient contribution
4035: $\frac{E_{\varphi,v}}{2\pi R}$ is comparable to $\frac{E_{m,v}}{2\pi R}$.
4036: Thus the typical energy $E_v$ of the vortex loop is obtained by
4037: multiplying $\frac{E_{m,v}}{2\pi R}+\frac{E_{\varphi,v}}{2\pi R}$ with
4038: the typical circumference $2\pi R\sim \frac{1}{g|\varphi|}$ of the loop. We have
4039: %**********
4040: \eqb
4041: \label{EANOvortex}
4042: E_v\sim 2\,\frac{\pi}{2}\,|\varphi|^2\times \frac{1}{g|\varphi|}=
4043: \pi\,\frac{|\varphi|}{g}\,.
4044: \eqe
4045: %*********
4046: From Eq.\,(\ref{EANOvortex}) we conclude that vortex loops
4047: become massless in the limit $g\to\infty$.
4048: For $r\gg \frac{1}{g|\varphi|}$ the (isotropic in the $x_1x_2$ plane) pressure
4049: $P_v(r)$ of the vortex configuration is still given by Eq.\,(\ref{Pvort}):
4050: For finite coupling $g$ vortex loops collapse as soon as they are created.
4051:
4052:
4053: \subsection{Derivation of the phase of a macroscopic complex scalar field}
4054:
4055: We consider a pair of center-vortex loops that are pierced by a circular
4056: contour $C$ of infinite radius, see Fig.\,\ref{Fig35}. The total
4057: center flux $F_{\pm,0}$ through the minimal surface spanned by
4058: $C$ is
4059: %********
4060: \eqb
4061: \label{centerfluxC}
4062: F_{\pm,0}=\left\{\begin{array}{c}\pm\frac{2\pi}{g}\\
4063: 0\end{array}\right.
4064: \eqe
4065: %********
4066: depending on whether at finite coupling $g$ the loop $A$ ($B$)
4067: collapses to nothing well before the loop $B$ ($A$) or whether this roughly happens
4068: at the same time. (In any case, vortex loops which start out {\sl without} getting pierced
4069: by $C$ do not contribute a center flux through $C$.)
4070:
4071: Unlike in the case of a pair of a
4072: monopole and an antimonopole discussed in
4073: Sec.\,\ref{defphismag} the flux $F_{\pm,0}$ takes {\sl discrete} values.
4074: %***********************
4075: \begin{figure}
4076: \begin{center}
4077: \leavevmode
4078: %\epsfxsize=9.cm
4079: \leavevmode
4080: %\epsffile[80 25 534 344]{}
4081: \vspace{5.2cm}
4082: \special{psfile=Fig-19.ps angle=0 voffset=-150
4083: hoffset=-80 hscale=50 vscale=50}
4084: \end{center}
4085: \caption{Two center-vortex loops of opposite flux being
4086: pierced by an $S_1$ of infinite radius.\label{Fig35}}
4087: \end{figure}
4088: %************************
4089: In analogy to the derivation
4090: of a macroscopic monopole condensate we may investigate the thermally averaged
4091: flux of the vortex-antivortex (spin-0) system in the limit where
4092: there is no spatial momentum of this system and where
4093: $g\to\infty$:
4094: %*****
4095: \eab
4096: \label{avfluxsysV}
4097: \lim_{g\to\infty}\,F_{\pm,0;\tiny\mbox{th}}(\delta)&=&
4098: \,4\pi\,\int d^3p\,\delta^{(3)}(\vec{p})\, n_B(\beta |2\,E_v(\vec{p})|)\,F_{\pm,0}\nonumber\\
4099: &=&0,\pm\frac{2}{\beta|\varphi|}=0,\pm\frac{\lambda^{3/2}_{c,M}}{\pi}\,.
4100: \eae
4101: %*********
4102: The phase of a macroscopic complex field $\Phi$ is defined as
4103: %********
4104: \eqb
4105: \label{phasePhi}
4106: \Gamma\frac{\Phi}{|\Phi|}(x)\equiv \lim_{g\to\infty}\la \exp[ig\oint_{{C}(x)}dz_\mu\,
4107: (a^D)^\mu]\ra
4108: \eqe
4109: %**********
4110: where $\Gamma$ is a complex constant, and $(a^D)^\mu$ denotes
4111: the gauge field of a center vortex. The expectation on the right-hand side of Eq.\,(\ref{phasePhi})
4112: is proportional to the expectation of 't Hooft's
4113: loop operator \cite{'tHooft1978}. (Green functions of this
4114: operator change their phase by $-1$ (SU(2)) and $\exp[\pm \frac{2\pi i}{3}]$ (SU(3))
4115: under an exchange of the order of any two of their arguments:
4116: A feature which is familiar from Green functions of fermionic fields.)
4117: The possible values of $\Phi$'s phase are parametrized by the average center
4118: flux $\lim_{g\to\infty}\,F_{\pm,0;\tiny\mbox{th}}(\delta)$ in Eq.\,(\ref{avfluxsysV}).
4119:
4120: According to Eq.\,(\ref{avfluxsysV}) the condensate $\Phi$ of center-vortex loops
4121: is determined by discrete parameter values which can be
4122: normalized as $\hat{\tau}=\pm 1,0$. Recall that at $\lambda_{c,M}$
4123: vortex loops start to be stable excitations since
4124: their pressure $P_v(r)$ is zero outside the (infinitely thin)
4125: vortex core. Once the field $\Phi$
4126: acquires a nonzero modulus its phase is observed to jump locally
4127: in space. (Each jump corresponds to a stable vortex loop travelling in from infinity and getting
4128: pierced by ${\cal C}$.) Thus we are led to interpret jumps in
4129: $\Phi$'s phase as creation processes for (fermionic)
4130: particles. (There are two degenerate
4131: polarizations of these particles: The two possible directions of center
4132: flux in a given vortex-loop. By travelling in from infinity the
4133: vortex loop makes $\Phi$'s phase jump twice: A created unit of
4134: flux is associated with a forward jump (piercing by $C$) while a backward jump corresponds to
4135: minus this flux (no piercing by $C$, center-vortex loop lies inside $C$).) If a single
4136: center-vortex loop is created with sufficiently large momentum then a part of
4137: its kinetic energy can be converted
4138: into the mass of self-intersection points by subsequent twisting. Self-intersection points are
4139: $Z_2$ magnetic monopoles, each contributing $\sim \Lambda_C$
4140: to the mass of the state where $\Lambda_C$ is a mass scale. Twisting does
4141: not alter the fact that only two possible polarizations (spin-1/2 fermions) occur.
4142: (The magnetic flux is reversed by a $Z_2$ monopole \cite{Reinhardt2001}, see Fig.\,\ref{intersect}.)
4143: %***********************
4144: \begin{figure}
4145: \begin{center}
4146: \leavevmode
4147: %\epsfxsize=9.cm
4148: \leavevmode
4149: %\epsffile[80 25 534 344]{}
4150: \vspace{4.5cm}
4151: \special{psfile=Fig-20.ps angle=0 voffset=-120
4152: hoffset=-170 hscale=60 vscale=50}
4153: \end{center}
4154: \caption{The creation of an isolated $Z_2$ monopole by self-intersection
4155: of a center-vortex loop. \label{intersect}}
4156: \end{figure}
4157: %************************
4158: We conclude that the mass spectrum $m_n$ of
4159: fermionic excitations is equidistant:
4160: %*********
4161: \eqb
4162: \label{massspectferm}
4163: m_n\sim n\,\Lambda_C\,,\ \ \ \ \ (n=0,1,2,\cdots)\,.
4164: \eqe
4165: %*********
4166: The process of fermion creation violates the spatial homogeneity
4167: of the system and thus thermal equilibrium. Fermion creation, that is, the process of
4168: sucking in stable center-vortex loops from infinity, can only go on so long as
4169: the energy density provided by the field $\Phi$
4170: is finite. Thus the field
4171: $\Phi$ must eventually relax to one of the possible zero-energy
4172: minima of its potential. This phenomenon is
4173: generally known as tachyonic pre- and re-heating \cite{nonequ}.
4174:
4175: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
4176: The symmetry, which is dynamically broken by center-vortex condensation,
4177: is a local magnetic $Z_2$. After a relaxation of
4178: $\Phi$ to zero energy density the ground state must exhibit
4179: the associated $Z_2$ degeneracy. We conclude that for
4180: SU(2) the parameters $\hat{\tau}=\pm 1$
4181: must be identified: They describe the same
4182: minimum of $\Phi$'s potential. The parameter value
4183: $\hat{\tau}=0$ corresponds to the other degenerate minimum.
4184: The center flux carried by a given
4185: flux line is associated with the differences in $\Phi$'s phase in each
4186: minimum of $\Phi$'s potential.\vspace{0.1cm}\\
4187: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
4188: For SU(3) the dynamically broken symmetry is a local magnetic $Z_3$.
4189: As a consequence, each of the three possible values
4190: $\hat{\tau}=\pm 1,0$ describes
4191: one of the three possible, distinct minima of $\Phi$'s potential.
4192: Before relaxation to zero energy density local jumps of $\Phi$'s phase
4193: generate two distinct species of flux loops: Each associated with the three differences
4194: in $\Phi$'s phase modulo three. (A short jump between two neighbouring unit roots is
4195: equivalent to a long jump into the opposite direction involving
4196: the third unit root as a brief stop-over.)
4197:
4198: \subsection{The potential of the macroscopic complex scalar field}
4199:
4200: At $T_{c,M}$ dual gauge modes decouple. Moreover, a
4201: condensate of (Cooper-like) pairs of single center-vortex--center-antivortex loops confines fundamental electric and fermionic
4202: test charges. This happens
4203: because each condensed center-vortex loop represents an
4204: electric dipole. A condensate of such dipoles squeezes an externally applied
4205: electric field into a flux tube: Oppositely charged test particles are subject to
4206: a linear potential at large distances. Thus the center phase
4207: is confining both test charges and {\sl all} gauge modes:
4208: There is complete confinement. (The Polyakov loop
4209: expectation is zero below $T_{c,M}$, the 't Hooft loop expectation $\Phi$, which is the
4210: dual order parameter for confinement, jumps to a finite value.)
4211:
4212: The effective action for
4213: the center phase thus is only a functional of $\Phi$ and $\bar{\Phi}$. Moreover,
4214: thermal equilibrium (that is, periodicity in Euclidean time)
4215: is no longer applicable to constrain $\Phi$'s potential. According to our discussion in
4216: the last section $\Phi$'s potential $V_C$ must
4217: be (i) invariant under center jumps only (invariance under a larger
4218: (continuous or discontinuous) symmetry is forbidden), (ii) it must allow for
4219: fermion creation by center jumps, and (iii) center-degenerate minima of
4220: zero energy density have to occur. Moreover, (iv) we insist on the occurrence
4221: of a single mass scale $\La_C$ only. From (i) we conclude that $V_C$ can not
4222: be a function of $\bar{\Phi}\Phi$ alone.
4223:
4224: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
4225: The only potential $V_C$ satisfying (i),(ii), (iii), and (iv) is given by
4226: %*********
4227: \eqb
4228: \label{2potC}
4229: V_C=\overline{v_C}\,v_C\equiv\overline{\left(\frac{\Lambda_C^3}{\Phi}-\La_C\,\Phi\right)}\,
4230: \left(\frac{\Lambda_C^3}{\Phi}-\La_C\,\Phi\right)\,.
4231: \eqe
4232: %*********
4233: The zero-energy minima of $V_M$
4234: are at $\Phi=\pm \Lambda_C$. It is easy to show that
4235: adding or subtracting powers $(\Phi^{-1})^{2l+1}$ or
4236: $\Phi^{2k+1}$ in $v_C$, where $k,l=0,1,2,3,\cdots$,
4237: generates additional minima and thus destroys
4238: the center degeneracy of the ground state after relaxation
4239: and/or violates the demand for zero energy-density at a finite value of $\Phi$
4240: in these minima (requirement (iii)). \vspace{0.1cm}\\
4241: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
4242: The only potential $V_C$ satisfying (i),(ii), (iii), and (iv) is given by
4243: %*********
4244: \eqb
4245: \label{3potC}
4246: V_C=\overline{v_C}\,v_C\equiv\overline{\left(\frac{\Lambda_C^3}{\Phi}-\Phi^2\right)}\,
4247: \left(\frac{\Lambda_C^3}{\Phi}-\Phi^2\right)\,.
4248: \eqe
4249: %*********
4250: The zero-energy minima of $V_C$
4251: are at $\Phi=\Lambda_C\exp\left[\pm\frac{2\pi i}{3}\right]$ and $\Phi=\Lambda_C$.
4252: Again, adding or subtracting powers $({\Phi}^{-1})^{3l+1}$ or $(\Phi)^{3k-1}$
4253: in $v_C$, where $l=0,1,2,3,\cdots$ and $k=1,2,3,4,\cdots$, violates requirement (iii).
4254:
4255: In Fig.\,\ref{Fig37} plots of the potentials in Eq.\,(\ref{2potC}) and Eq.\,(\ref{3potC}) are shown.
4256: %***********************
4257: \begin{figure}
4258: \begin{center}
4259: \leavevmode
4260: %\epsfxsize=9.cm
4261: \leavevmode
4262: %\epsffile[80 25 534 344]{}
4263: \vspace{5.5cm}
4264: \special{psfile=Fig-21.ps angle=0 voffset=-150
4265: hoffset=-190 hscale=50 vscale=50}
4266: \end{center}
4267: \caption{The potential $V_C=\overline{v_C(\Phi)}v_C(\Phi)$ for the center-vortex
4268: condensate $\Phi$. Notice the regions of negative tangential curvature
4269: inbetween the minima.\label{Fig37}}
4270: \end{figure}
4271: %************************
4272: The ridges of negative tangential curvature are classically forbidden:
4273: The field $\Phi$ tunnels through these ridges, and a
4274: phase change, which is determined by an element of the center $Z_2$ (SU(2)) or $Z_3$ (SU(3)),
4275: occurs locally in space. This is the afore-mentioned generation of one unit of center flux.
4276:
4277: \subsection{Thermodynamics close to the Hagedorn transition \label{Hagedorn}}
4278:
4279: The action describing the process of relaxation of $\Phi$
4280: to one of $V_M$'s minima is
4281: %**********
4282: \eqb \label{actionPhi}
4283: S = \int d^4x
4284: \left(\frac{1}{2}\,\overline{\partial_\mu \Phi} \partial^\mu \Phi - \frac12\, V_C \right) \,.
4285: \eqe
4286: %***********
4287: In contrast to the electric and magnetic phases the action $S$ in Eq.\,(\ref{actionPhi})
4288: does not determine a classical, macroscopic ground state if $\Phi$ is not
4289: in one of $V_C$'s minima. Though tunneling processes occur
4290: in real time they can be described by a Euclidean simulation,
4291: WKB-like approximations are thinkable. Alternatively, the computation of fermion creation
4292: rates can be performed on a finite-temperature
4293: lattice based on the theory (\ref{actionPhi}). An interesting object to be measured
4294: is $\Phi$'s two-point correlator $\Pi(x)\equiv\la\bar{\Phi}(x)\Phi(0)\ra$.
4295: Projecting onto a given spatial momentum $\vec{p}$ at a given temperature, the strength of intermediate tachyonic
4296: modes can be extracted by a Fourier analysis
4297: of the $\tau$ dependence in $\int d^3x\,\exp[i\vec{p}\cdot \vec{x}]\,\Pi(\tau,\vec{x})$. This gives a measure for the
4298: density of states $\rho_n$ for fermions of
4299: mass $\sim n\,\Lambda_C$ and spatial momentum $\vec{p}$ \cite{HofmannSchefflerStamatescu2005}.
4300:
4301: Let us estimate $\rho_n$ for SU(2).
4302: The multiplicity of fermion states with $n$ self-intersections is given by twice the number $L_n$
4303: of bubble diagrams with $n$ vertices in a scalar $\lambda \phi^4$ theory.
4304: In \cite{BenderWu1969} the minimal number of such diagrams $L_{n,min}$ was
4305: estimated to be $L_{n,min}=n!3^{-n}$. Using Stirling's formula
4306: this can be approximated as
4307: %*******
4308: \eab
4309: \label{Stirling}
4310: L_{n,min}&\sim&\frac{1}{3}\sqrt{2\pi n}\,\left(3\e\right)^{-n}\,n^{n}\nonumber\\
4311: &=&\frac{\sqrt{2\pi}}{3}\,\exp\left[n\Big(\log n-(\log3+1)\Big)+\frac12\log n\right]\nonumber\\
4312: &\sim&\frac{\sqrt{2\pi}}{3}\,\exp\left[n\log n\right]\,
4313: \eae
4314: %*******
4315: for $n\gg 1$. So the number $F_n$ of fermion states with mass $m_n\sim n\,\La_C$ is
4316: bounded from below roughly by
4317: %********
4318: \eqb
4319: \label{Fn}
4320: F_n\sim 2\times \frac{\sqrt{2\pi}}{3}\,\exp\left[n\log n\right]=
4321: \frac{\sqrt{8\pi}}{3}\,\exp\left[n\log n\right]\,.
4322: \eqe
4323: %*******
4324: We now estimate the density $\rho_{n,0}$ of fermion states at
4325: rest $\vec{p}=0$ (or $\tilde{\rho}(E=n\La_C)$)
4326: by differentiating $F_n$ with respect to
4327: $n$ and dividing the result by the level-spacing
4328: $\delta m_n=\Lambda_C$. We have
4329: %******
4330: \eab
4331: \label{statdes}
4332: \rho_{n,0}&>&\frac{\sqrt{8\pi}}{3\La_C}\,\exp[n\log n]\Big(\log n+1\Big)\,\ \ \ \ \ \mbox{or}\nonumber\\
4333: \tilde{\rho}(E)&>&\frac{\sqrt{8\pi}}{3\La_C}\exp[\frac{E}{\La_C}\log\frac{E}{\La_C}]
4334: \Big(\log\frac{E}{\La_C}+1\Big)\,.
4335: \eae
4336: %*******
4337: Eq.\,(\ref{statdes}) tells us that the density of static fermion
4338: states is more than exponentially increasing with energy $E$. The partition function
4339: $Z_{\Phi}$ thus is estimated as
4340: %******
4341: \eab
4342: \label{partfunctionphi}
4343: Z_{\Phi}&>&\int_{E^*}^\infty dE\,\tilde{\rho}(E)\,n_F(\beta E)\nonumber\\
4344: &>&\frac{\sqrt{8\pi}}{3\La_C}\,\int_{E^*}^\infty dE\,\exp\left[\frac{E}{\La_C}\right]\,
4345: \exp[-\beta E]\,,
4346: \eae
4347: %*******
4348: where $E^*\gg \Lambda_C$ is the energy where we start to trust
4349: our approximations. Thus $Z_{\Phi}$ diverges at some
4350: temperature $T_H<\Lambda_C$. (Strictly speaking, $T_H=0$
4351: according to Eq.\,(\ref{statdes}). This is an artefact of
4352: our assumption that all states are infinitely narrow.
4353: There are, however, finite widths for higher-charge states ($n>1$) since contact interactions
4354: exist between vortex lines and intersection points.
4355: Moreover, in the real world higher-charge states are even broader
4356: due to their decay and their mutual annihilation into charge-one and
4357: charge-zero states. This happens because an SU(2) theory, which is not
4358: confining at the present temperature of the Universe,
4359: mixes with the theory under consideration
4360: at large temperatures and thus couples its massless gauge mode -- the photon -- to the
4361: $Z_2$ charges of the latter. The larger $n$ the
4362: broader the associated state and the less reliable our
4363: assumption of infinitely narrow states.) This is the
4364: celebrated Hagedorn transition. (At $T_H$ the entropy
4365: diverges and the system condenses self-intersecting center-vortex
4366: loops into a new ground state: The monopole condensate of
4367: the magnetic phase. The process of monopole condensation from below violates the spatial
4368: homogeneity of the system: $Z_2$ charges loose their
4369: mass by dense packing only. Thus the Hagedorn transition is
4370: genuinely nonthermal.)
4371:
4372: Once $\Phi$ has is settled into $V_C$'s
4373: minima $\Phi_{\tiny\mbox{min}}$ there are no
4374: quantum fluctuations $\delta\Phi$. Writing $\Phi=|\Phi|\exp\left[i\frac{\theta}{\La_c}\right]$,
4375: this is a consequence of the following fact:
4376: %*******
4377: \eqb
4378: \label{minimacur}
4379: \left.\frac{\pd^2_{\theta} V_C(\Phi)}{|\Phi|^2}\right|_{\Phi_{\tiny\mbox{min}}}=
4380: \left.\frac{\pd^2_{|\Phi|} V_C(\Phi)}
4381: {|\Phi|^2}\right|_{\Phi_{\tiny\mbox{min}}}
4382: =\left\{\begin{array}{c}8\,\ \ \ \ \ (\mbox{SU(2)})\\
4383: 18\,\ \ \ \ \ (\mbox{SU(3)})\end{array}\right.\,.
4384: \eqe
4385: %*******
4386: Thus radial {\sl and} tangential
4387: fluctuations around $\Phi_{\tiny\mbox{min}}$ would have a mass $m_{\delta\Phi}$ which is
4388: sizably larger than the compositeness scale $|\Phi_{\tiny\mbox{min}}|$
4389: for both SU(2) and SU(3). Since $|p^2_{\delta\Phi}+m^2_{\delta\Phi}|\le
4390: |\Phi_{\tiny\mbox{min}}|^2$ for any allowed Euclidean momentum
4391: $p^2_{\delta\Phi}>0$ this means that the fluctuations $\delta\Phi$
4392: are absent: After relaxation the ground state of the
4393: center phase has a vanishing pressure and a vanishing energy density.
4394:
4395:
4396: \section{Matching the pressure\label{MCC}}
4397:
4398: The mass scales $\La_E$ and $\La_M$, which determine the modulus of the
4399: adjoint Higgs field $\phi$ and the moduli of the
4400: monopole condensates $\varphi$ (SU(2)) and $\varphi_1,\varphi_2$ (SU(3)), respectively, are
4401: related. This derives from the fact that across a second-order
4402: transition the pressure is continuous. We have
4403: %********
4404: \eab
4405: \label{laElaM}
4406: \Lambda_M&=&\left(4+\frac{\lambda_{c,E}^3}{720\pi^2}\right)^{1/3}\La_E\,,\ \ \ \ \ \ \ (\mbox{SU(2)})\,,\nonumber\\
4407: \Lambda_M&=&\left(2+\frac{\lambda_{c,E}^3}{720\pi^2}\right)^{1/3}\La_E\,,\ \ \ \ \ \ \ (\mbox{SU(3)})\,.
4408: \eae
4409: %********
4410: Across the magnetic-center transition the pressure is not
4411: continuous. We may, however, estimate the scale
4412: $\Lambda_C$ by assuming thermal equilibrium at the
4413: {\sl onset} of this transition. (This assumption underlies
4414: Eq.\,(\ref{avfluxsysV}).) This gives
4415: %********
4416: \eqb
4417: \label{laMlaC}
4418: \Lambda_M\sim 2^{1/3}\,\La_C\,,\ \ \ \ (\mbox{SU(2)})\,\ \ \ \ \ \mbox{and}\ \ \ \ \
4419: \Lambda_M\sim \La_C\,,\ \ \ \ (\mbox{SU(3)})\,.
4420: \eqe
4421: %********
4422:
4423:
4424: \section{Thermodynamical quantities\label{PEEN}}
4425:
4426: \subsection{Results}
4427:
4428: In this section we present our numerical results for one-loop
4429: temperature evolutions of thermodynamical quantities throughout the
4430: electric and magnetic phase.\vspace{0.1cm}\\
4431: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
4432: \noindent In the electric phase the ratio of pressure $P$ and $T^4$ is given as
4433: %**********
4434: \eqb
4435: \label{pressureEP2}
4436: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_E^4}\left[\frac{2\lambda_E^4}{(2\pi)^6}\left(2\bar{P}(0)+
4437: 6\bar{P}(2a)\right)+2\lambda_E\right]\,
4438: \eqe
4439: %************
4440: where the function $\bar{P}(a)$ and the dimensionless mass parameter
4441: $a$ are defined in Eq.\,(\ref{P(y)}) and
4442: Eq.\,(\ref{aofela}), respectively. In
4443: the magnetic phase we have
4444: %**********
4445: \eqb
4446: \label{pressureMP2}
4447: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_M^4}
4448: \left[\frac{6\lambda_M^4}{(2\pi)^6}\bar{P}(a)+\frac{\lambda_M}{2}\right]\,
4449: \eqe
4450: %**********
4451: where $a$ is defined in Eq.\,(\ref{massspecM}).
4452:
4453: \noindent In the electric phase the ratio of energy density $\rho$
4454: and $T^4$ is given as
4455: %************
4456: \eqb
4457: \label{rhoEP2}
4458: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_E^4}
4459: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left(2\bar{\rho}(0)+
4460: 6\bar{\rho}(2a)\right)+2\lambda_E\right]\,
4461: \eqe
4462: %**********
4463: where the function $\bar{\rho}(a)$ is defined as
4464: %**********
4465: \eqb
4466: \label{rhobar}
4467: \bar{\rho}(a)\equiv \int_{0}^{\infty}dx\,x^2 \frac{\sqrt{x^2+a^2}}{\exp(\sqrt{x^2+a^2})-1}\,.
4468: \eqe
4469: %**********
4470: In the magnetic phase we have
4471: %**********
4472: \eqb
4473: \label{rhoMP2}
4474: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_M^4}
4475: \left[\frac{6\lambda_M^4}{(2\pi)^6}\bar{\rho}(a)+\frac{\lambda_M}{2}\right]\,.
4476: \eqe
4477: %**********
4478: The ratio of entropy density and $T^3$ is given as
4479: %************
4480: \eqb
4481: \label{ent}
4482: \frac{s}{T^3}=\frac{1}{T^4}\left(\rho+P\right)\,.
4483: \eqe
4484: %************
4485: Because the ground-state contributions cancel in $\frac{s}{T^3}$ this quantity
4486: is not as infrared sensitive as, e. g., $\frac{\rho}{T^4}$ or $\frac{P}{T^4}$:
4487: Lattice simulations are in a position to correctly measure the
4488: entropy density at low temperatures.
4489: \vspace{0.1cm}\\
4490: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
4491: In the electric phase we have
4492: %**********
4493: \eqb
4494: \label{pressureEP3}
4495: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_E^4}
4496: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left(4\bar{P}(0)+
4497: 3(4\,\bar{P}(a)+2\,\bar{P}(2a))\right)+
4498: 2\lambda_E\right]\,
4499: \eqe
4500: %***********
4501: and
4502: %**********
4503: \eqb
4504: \label{rhoEP3}
4505: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_E^4}
4506: \left[\frac{2\lambda_E^4}{(2\pi)^6}\left(4\bar{\rho}(0)+
4507: 3(4\,\bar{\rho}(a)+2\,\bar{\rho}(2a))\right)+
4508: 2\lambda_E\right]\,.
4509: \eqe
4510: %***********
4511: In the magnetic phase we have
4512: %**********
4513: \eqb
4514: \label{pressureMP3}
4515: \frac{P}{T^4}=-\frac{(2\pi)^4}{\lambda_M^4}
4516: \left[\frac{12\lambda_M^4}{(2\pi)^6}
4517: \bar{P}(a)+\lambda_M\right]\,
4518: \eqe
4519: %**********
4520: and
4521: %**********
4522: \eqb
4523: \label{rhoMP3}
4524: \frac{\rho}{T^4}=\frac{(2\pi)^4}{\lambda_M^4}
4525: \left[\frac{12\lambda_M^4}{(2\pi)^6}
4526: \bar{\rho}(a)+\lambda_M\right]\,.
4527: \eqe
4528: %**********
4529: The ratio of entropy density and $T^3$ is given in Eq.\,(\ref{ent}) where
4530: now the SU(3)-expressions for $P$ and $\rho$ have to be used.
4531:
4532: The result for $\frac{P}{T^4}$ is plotted in Fig.\,\ref{pressure} as
4533: a function of temperature throughout the electric and
4534: magnetic phase, Fig.\,\ref{pressureLat} depicts SU(3)-lattice results.
4535: %***********************
4536: \begin{figure}
4537: \begin{center}
4538: \leavevmode
4539: %\epsfxsize=9.cm
4540: \leavevmode
4541: %\epsffile[80 25 534 344]{}
4542: \vspace{4.5cm}
4543: \special{psfile=Fig-22.ps angle=0 voffset=-140
4544: hoffset=-195 hscale=50 vscale=50}
4545: \end{center}
4546: \caption{\protect{$\frac{P}{T^4}$ as a function of temperature for SU(2) (left panel) and SU(3) (right panel).
4547: The horizontal lines indicate the respective asymptotic values, the dashed vertical lines are the phase boundaries.
4548: \label{pressure}}}
4549: \end{figure}
4550: %************************
4551: %***********************
4552: \begin{figure}
4553: \begin{center}
4554: \leavevmode
4555: %\epsfxsize=9.cm
4556: \leavevmode
4557: %\epsffile[80 25 534 344]{}
4558: \vspace{4.5cm}
4559: \special{psfile=Fig-23.ps angle=0 voffset=-140
4560: hoffset=-140 hscale=50 vscale=40}
4561: \end{center}
4562: \caption{\protect{$\frac{P}{T^4}$ as a function of temperature for SU(3) as obtained on
4563: a ($16^3\times 4$)-lattice using the differential method with a universal two-loop
4564: perturbative $\beta$ function \protect\cite{Brown1988,Deng1988} and using the integral
4565: method (solid line) \protect\cite{EngelsFingberg1990}. The figure is taken
4566: from \protect\cite{EngelsFingberg1990}.\label{pressureLat}}}
4567: \end{figure}
4568: %************************
4569: Notice that the pressure is negative
4570: in the electric phase close to $\lambda_{E,c}$ and
4571: even more so in the magnetic phase where the ground state
4572: strongly dominates the thermodynamics of infrared sensitive quantities. Notice also the
4573: negative pressure in Fig.\,\ref{pressureLat} obtained close to the phase transition
4574: when the differential method is used in the lattice simulation.
4575: (For a discussion of differential versus integral
4576: method see Sec.\,\ref{DVI}.) We conclude that the finite-size
4577: constraints of realistic lattices have a severe effect on the obtained values of the
4578: pressure shortly above the first confining transition
4579: and even more so below this transition.
4580:
4581: Let us now discuss the behavior close to $T_{c,M}$ where the thermodynamical relation
4582: %**********
4583: \eqb
4584: \label{pdt}
4585: dP=S\,dT\,
4586: \eqe
4587: %**********
4588: starts to be violated. Eq.\,(\ref{pdt}) implies that
4589: in a homogeneous, thermalized system the pressure needs to be a strictly
4590: monotonic increasing function of temperature since $S$ is never negative.
4591: Crossing the point $T_{c,M}$ from above,
4592: the pressure jumps from a negative to a positive value. (On the magnetic side of the phase
4593: boundary the ground state strongly dominates the excitations, on the
4594: center side the vortex-condensate has zero pressure while fermionic
4595: excitations give a positive contribution.) There are two ways
4596: of seeing that thermal equilibrium must break down close to
4597: $T_{c,M}$. First, spatial homogeneity starts to be badly violated by
4598: discontinuous and local
4599: phase changes of the field $\Phi$ as soon as the system starts to
4600: condense center-vortex loops. The derivation of Eq.\,(\ref{pdt}),
4601: however, relies on thermal equilibrium and thus on spatial homogeneity.
4602: Second, one may assume thermal equilibrium and then lead this assumption to a contradiction.
4603: In thermal equilibrium the spectral density $\rho(E)$ in the center phase is more than
4604: exponentially increasing with energy $E$, see Eq.\,(\ref{statdes}). Thus the
4605: partition function diverges at $T=T_{H}<\Lambda_C$: A
4606: homogeneous system would need an infinite amount of energy per volume
4607: to increase its temperature beyond $T_{H}$. But this is a contradiction to the fact that
4608: a magnetic phase exists for $T>T_{H}$. (In an extended thermalized
4609: system the transition from the center phase to the magnetic phase is accomplished
4610: by an increase of the overall energy density: The excitation
4611: of very massive dual gauge modes, though very unlikely, is furnished energetically
4612: by the large energy residing in the system. If
4613: the considered system is of a small spatial extent, such as the interaction
4614: vertex in a scattering process, then the total energy of the system, e.g.,
4615: the center-of-mass energy being deposited into a vertex, needs to be larger than the
4616: very large mass of the dual gauge modes on the magnetic
4617: side of the phase boundary.) Thus thermal equilibrium breaks down
4618: for $T\sim T_H$.
4619:
4620: The result for $\frac{\rho}{T^4}$ as
4621: a function of temperature throughout the electric and
4622: magnetic phase is shown in Fig.\,\ref{rho}. Fig.\,\ref{En1} depicts an
4623: SU(2)-lattice result \cite{EnKaSaMo1982}. Notice the (small) discontinuities at
4624: $\lambda_{c,E}$. Their occurrence is explained by the fact that by crossing the
4625: electric-magnetic phase boundary the system needs to generate an
4626: extra polarization for each dual gauge mode compared to the two polarizations
4627: of a TLM mode. Extra polarizations are extra fluctuating
4628: degrees of freedom which increase the energy
4629: density on the magnetic side of the phase boundary.
4630: The situation is somewhat peculiar: On the one hand, the order parameter for the dynamical breaking
4631: of the dual gauge groups U(1)$_D$ (SU(2)) and U(1)$_D^2$ (SU(3))
4632: is continuous. On the other hand, there is a small amount of
4633: latent heat being released across the {\sl magnetic-electric} transition. (That is, by heating up
4634: the system starting in the magnetic phase.) As we shall see in
4635: Sec.\,\ref{Apps}, this is the reason for a dynamical stabilization of
4636: the temperature of the cosmic microwave background
4637: against gravitational expansion. (Thus we may look forward to
4638: enjoy the privilege of the photon's masslessness
4639: for another sizable fraction of today's age of the Universe \cite{GH2005}.) Again,
4640: the energy density is dominated by the
4641: ground-state contribution in the electric phase close to the electric-magnetic transition
4642: and even more so in the magnetic
4643: phase. Notice also that $\rho=-P$ at the point, where the system starts
4644: to condense center-vortex loops, that the magnetic phase is
4645: narrower for SU(3) than it is for SU(2), and that $\frac{\rho}{T^4}$
4646: dips in a much steeper way at the electric-magnetic transition for SU(3) than for SU(2).
4647: %***********************
4648: \begin{figure}
4649: \begin{center}
4650: \leavevmode
4651: %\epsfxsize=9.cm
4652: \leavevmode
4653: %\epsffile[80 25 534 344]{}
4654: \vspace{5.5cm}
4655: \special{psfile=Fig-24.ps angle=0 voffset=-150
4656: hoffset=-195 hscale=50 vscale=50}
4657: \end{center}
4658: \caption{$\frac{\rho}{T^4}$ as a function of temperature for SU(2) (left panel) and SU(3) (right panel).
4659: The horizontal lines indicate the respective asymptotic values, the dashed
4660: vertical lines are the phase boundaries.\label{rho}}
4661: \end{figure}
4662: %************************
4663: %***********************
4664: \begin{figure}
4665: \begin{center}
4666: \leavevmode
4667: %\epsfxsize=9.cm
4668: \leavevmode
4669: %\epsffile[80 25 534 344]{}
4670: \vspace{6cm}
4671: \special{psfile=Fig-25.ps angle=0 voffset=-180
4672: hoffset=-220 hscale=80 vscale=80}
4673: \end{center}
4674: \caption{$\frac{\rho}{T^4}$ as obtained from the SU(2)-lattice simulation in \protect\cite{EnKaSaMo1982}.\label{En1}}
4675: \end{figure}
4676: %************************
4677:
4678: The result for the interaction measure
4679: $\frac{\Delta}{T^4}\equiv\frac{\left(\rho-3P\right)}{T^4}$ is shown in
4680: Fig.\,\ref{IM}. Figs.\,\ref{En3} and \ref{B1} are lattice results.
4681: Notice the rapid approach to the free-gas limit in Fig.\,\ref{IM} and
4682: the large values of $\frac{\Delta}{T^4}$ in the magnetic phase. Interestingly, there is a small
4683: bump to the left of the phase boundary in Fig.\,\ref{En3}.
4684:
4685: The result for the ratio of the specific heat per unit volume $c_V\equiv \frac{d\rho}{dT}$ and
4686: $T^3$ is shown in Fig.\,\ref{SH}, Fig.\,\ref{En2} is an SU(2)-lattice result \cite{EnKaSaMo1982}.
4687: The quantity $\frac{c_V}{T^3}$ peaks both at the electric-magnetic and
4688: the magnetic-center transition. The finite peak at the former phase boundary is
4689: in agreement with the electric-magnetic transition being second-order. Moreover,
4690: we have $\left.\frac{c_V}{T^3}\right|_{T_{c,E};\tiny\mbox{SU(3)}}\sim 3\,
4691: \left.\frac{c_V}{T^3}\right|_{T_{c,E};\tiny\mbox{SU(2)}}$. This
4692: explains why lattice simulations prefer to identify the confining transition
4693: as weakly first order for SU(3), see \cite{LuciniTeperWenger2003,LuciniTeperWenger2005}
4694: and references therein. Now,
4695: $3\not=\infty$ but in the vicinity of $T_{c,E}$ lattice results
4696: for infrared sensitive quantities, such as $\frac{c_V}{T^3}$, are
4697: not reliable anyway.
4698: %***********************
4699: \begin{figure}
4700: \begin{center}
4701: \leavevmode
4702: %\epsfxsize=9.cm
4703: \leavevmode
4704: %\epsffile[80 25 534 344]{}
4705: \vspace{5.5cm}
4706: \special{psfile=Fig-26.ps angle=0 voffset=-150
4707: hoffset=-195 hscale=50 vscale=50}
4708: \end{center}
4709: \caption{The interaction measure $\frac{\Delta}{T^4}$ as a function of temperature
4710: for SU(2) (left panel) and SU(3) (right panel). The asymptotic value in both cases is $\frac{\Delta}{T^4}=0$, the
4711: dashed vertical lines are the phase boundaries.\label{IM}}
4712: \end{figure}
4713: %************************
4714: %***********************
4715: \begin{figure}
4716: \begin{center}
4717: \leavevmode
4718: %\epsfxsize=9.cm
4719: \leavevmode
4720: %\epsffile[80 25 534 344]{}
4721: \vspace{7.5cm}
4722: \special{psfile=Fig-27.ps angle=0 voffset=-220
4723: hoffset=-220 hscale=80 vscale=50}
4724: \end{center}
4725: \caption{$\frac{\Delta}{T^4}$ as obtained from the SU(2)-lattice
4726: simulation in \protect\cite{EnKaSaMo1982}.\label{En3}}
4727: \end{figure}
4728: %************************
4729: %***********************
4730: \begin{figure}
4731: \begin{center}
4732: \leavevmode
4733: %\epsfxsize=9.cm
4734: \leavevmode
4735: %\epsffile[80 25 534 344]{}
4736: \vspace{12.8cm}
4737: \special{psfile=Fig-28.ps angle=0 voffset=-380
4738: hoffset=-220 hscale=80 vscale=50}
4739: \end{center}
4740: \caption{$\frac{P}{T^4}$ and $\frac{\Delta}{T^4}$ as obtained from the SU(3)-lattice
4741: simulation in \protect\cite{Bielfeld1996}.\label{B1}}
4742: \end{figure}
4743: %************************
4744:
4745: The result for $\frac{S}{T^3}$ is shown in Fig.\,\ref{ST3}.
4746: Fig.\,\ref{ST3lat} depicts a lattice result for SU(3)
4747: obtained with the differential method \cite{Brown1988}.
4748: The entropy density $S$ is a measure for the `mobility' of gauge modes. Notice
4749: the jump of $S/T^3$ which, again, is explained by the additional
4750: polarization of the dual gauge mode in the magnetic phase. Notice also that $\frac{S}{T^3}$ vanishes at
4751: the point $T_{c,M}$ where the system condenses center vortices. At this point dual gauge modes are
4752: infinitely heavy: The thermodynamics is completely determined by the ground state. The numerical
4753: agreement between the lattice result (d) (largest lattice) in Fig.\,\ref{ST3lat} and the SU(3)-result
4754: in Fig.\,\ref{ST3} is striking. The two data points to the left of the jump
4755: in (d) indicate that an ambiguity exists for the value of $\frac{S}{T^3}$
4756: very close to the transition.
4757: The jump itself corresponds to the large slope of $\frac{S}{T^3}$ on the electric side of the phase
4758: boundary in Fig.\,\ref{ST3}. The observed agreement is explained by the small sensitivity of
4759: the quantity $\frac{S}{T^3}$ on the ground-state physics making the finite-volume
4760: lattice simulation reliable close to the electric magnetic transition.
4761:
4762: %***********************
4763: \begin{figure}
4764: \begin{center}
4765: \leavevmode
4766: %\epsfxsize=9.cm
4767: \leavevmode
4768: %\epsffile[80 25 534 344]{}
4769: \vspace{5.5cm}
4770: \special{psfile=Fig-29.ps angle=0 voffset=-155
4771: hoffset=-200 hscale=50 vscale=50}
4772: \end{center}
4773: \caption{$\frac{c_V}{T^3}$ as a function of temperature for SU(2) (left panel) and SU(3) (right panel).
4774: The horizontal lines signal the respective asymptotic values, the dashed
4775: vertical lines are the phase boundaries.\label{SH}}
4776: \end{figure}
4777: %************************
4778: %***********************
4779: \begin{figure}
4780: \begin{center}
4781: \leavevmode
4782: %\epsfxsize=9.cm
4783: \leavevmode
4784: %\epsffile[80 25 534 344]{}
4785: \vspace{6.0cm}
4786: \special{psfile=Fig-30.ps angle=0 voffset=-180
4787: hoffset=-250 hscale=70 vscale=50}
4788: \end{center}
4789: \caption{$\frac{c_V}{T^3}$ as obtained from the SU(2)-lattice simulation in \protect\cite{EnKaSaMo1982}.\label{En2}}
4790: \end{figure}
4791: %************************
4792: The data files needed to generate the plots in Figs.\,\ref{pressure}, \ref{rho}, \ref{IM}, \ref{SH},
4793: and \ref{ST3} are provided by the author upon request.
4794:
4795: \subsection{Comparison with the lattice\label{complat}}
4796:
4797: \subsubsection{Specific observations}
4798:
4799: \noindent\underline{SU(2) case:}\vspace{0.1cm}\\
4800: The results of an early lattice measurements of the energy density $\rho$
4801: and the interaction measure $\Delta\equiv \rho-3P$ in a pure SU(2) gauge theory
4802: were reported in \cite{EnKaSaMo1982}. In that work the critical temperature $T_c$ for
4803: the deconfinement transition was determined from the critical behavior of the
4804: Polyakov-loop expectation and the peak position
4805: of the specific heat using a Wilson action. The function $\Delta(T)$ was extracted by multiplying the
4806: lattice $\beta$ function with the difference of
4807: plaquette expectations at finite and zero temperature (symmetric lattice).
4808: This assures that $\Delta$ vanishes for $T\to 0$. What is subtracted in \cite{EnKaSaMo1982}
4809: at finite $T$ is, however, {\sl not} the value $\Delta(T=0)$ since the plaquette expectation on the symmetric lattice
4810: is multiplied with the {\sl finite}-$T$ value of the $\beta$ function. Apart from this
4811: approximation, the use of a perturbative $\beta$ function was assumed for all
4812: temperatures. The simulation was carried out on a (rather small)
4813: $(10^3\times 3)$-lattice.
4814:
4815: Let us compare our results with those of \cite{EnKaSaMo1982}.
4816: %***********************
4817: \begin{figure}
4818: \begin{center}
4819: \leavevmode
4820: %\epsfxsize=9.cm
4821: \leavevmode
4822: %\epsffile[80 25 534 344]{}
4823: \vspace{5.0cm}
4824: \special{psfile=Fig-31.ps angle=0 voffset=-150
4825: hoffset=-195 hscale=50 vscale=50}
4826: \end{center}
4827: \caption{$\frac{S}{T^3}$ as a function of temperature for SU(2) (left panel) and SU(3) (right panel).
4828: The horizontal lines signal the respective asymptotic values. \label{ST3}}
4829: \end{figure}
4830: %************************
4831: %***********************
4832: \begin{figure}
4833: \begin{center}
4834: \leavevmode
4835: %\epsfxsize=9.cm
4836: \leavevmode
4837: %\epsffile[80 25 534 344]{}
4838: \vspace{5.0cm}
4839: \special{psfile=Fig-32.ps angle=0 voffset=-150
4840: hoffset=-190 hscale=50 vscale=50}
4841: \end{center}
4842: \caption{$\frac{S}{T^3}$ as a function of $\beta$
4843: obtained in SU(3) lattice gauge theory using the differential method and a
4844: perturbative beta function \protect\cite{Brown1988}. The simulations were performed on (a) $16^3\times 4$, (b)
4845: ($24^3\times 4$)-, (c) ($16^3\times 6$)- (open circles) and ($20^3\times 6$)- (closed circles), and
4846: (d) ($24^3\times 6$)-lattices. Using the ($24^3\times 6$)-lattice, the critical
4847: value of $\beta$ is between 5.8875 and 5.90. \label{ST3lat}}
4848: \end{figure}
4849: %************************
4850: The lattice results for $\rho$ in \cite{EnKaSaMo1982} differ drastically from
4851: our results for temperatures close the first confinement, that is, the electric-magnetic
4852: transition. (The lattice is doomed to produce incorrect results for infrared sensitive quantities
4853: close to the electric-magnetic transition and in the magnetic phase: A finite spatial
4854: lattice size $L$ cuts off
4855: {\sl physical} correlations on length scales $>L$ since the correlation length
4856: $l_M=M_{m}^{-1}>L$ close to the electric-magnetic transition
4857: and $l_M=\infty$ in the magnetic phase. Here $M_{m}$ denotes the
4858: mass of a magnetic monopole.)
4859:
4860: \noindent We obtain
4861: %*****
4862: \eqb
4863: \label{rhooverrhoSBc}
4864: \left.\frac{\rho}{\rho_{\tiny\mbox{SB}}}\right|_{T\sim 1.5\,T_{c,E}}\sim 1.27\,,
4865: \eqe
4866: %*****
4867: where $\rho_{\tiny\mbox{SB}}\equiv\frac{\pi^2}{5}T^4$ denotes the Stefan-Boltzmann limit
4868: (ideal gas of three species of massless gluons with two polarizations each). On
4869: the lattice this ratio is measured to be
4870: smaller than unity: $\left.\frac{\rho}{\rho_{\tiny\mbox{SB}}}\right|_{T\sim 1.5\,T_{c,E}}=0.84$.
4871: At $T\sim 5T_{c,E}$ we obtain
4872: %*****
4873: \eqb
4874: \label{rhooverrhoSB5}
4875: \left.\frac{\rho}{\rho_{\tiny\mbox{SB}}}\right|_{T\sim 5\,T_{c,E}}\sim 1.33\,
4876: \eqe
4877: %*****
4878: while the lattice measures $\left.\frac{\rho}{\rho_{SB}}\right|_{T\sim 5\,T_{c,E}}=0.85$.
4879:
4880: Our results for the pressure $P$ are {\sl negative} for $T$
4881: close to $T_{c,E}$ (see Fig.\, \ref{pressure}) -- much in
4882: contrast to the positive values obtained
4883: in \cite{EnKaSaMo1982}. At $T\sim 5\,T_{c,E}$ we obtain
4884: %*****
4885: \eqb
4886: \label{PoverrhoSB5}
4887: \left.\frac{P}{P_{\tiny\mbox{SB}}}\right|_{T\sim 5\,T_{c,E}}\sim 1.31\,
4888: \eqe
4889: %*****
4890: while the lattice measures $\left.\frac{P}{P_{\tiny\mbox{SB}}}\right|_{T\sim 5\,T_{c,E}}\sim 0.88$. (On the lattice
4891: $P$ is extracted from the measured values of $\Delta$ and $\rho$, and
4892: $P_{\tiny\mbox{SB}}\equiv\frac{\pi^2}{15}T^4$ denotes the
4893: Stefan-Boltzmann limit.)
4894:
4895: Notice that the results in Eq.\,(\ref{rhooverrhoSB5}) and Eq.\,(\ref{PoverrhoSB5}) are very close
4896: to the ratio $\frac{2\times 3+2}{6}\frac{4}{3}$ of the number of degrees of freedom in a gas of
4897: two species of (nearly) massless gluons (three polarizations per species) and one massless species
4898: and a gas of massless gluons. (At $\lambda_E\sim 75$ the value of the mass
4899: parameter is $a\sim 2\pi\frac{5.5}{650}\sim 0.086$. Thus the Boltzmann suppression is small for TLH modes,
4900: compare also with Fig.\,\ref{IM}.) At extremely high temperatures a TLH mode `remembers' its massiveness
4901: at low temperatures in terms of an extra polarization. The latter originates from a
4902: tiny mass which solves the infrared problem of loop expansions, for formal arguments see \cite{HofmannSept2006}.
4903:
4904: The peak-value of the specific heat is about
4905: $\left.\frac{c_V}{T^3}\right|_{T_{c,E}}\sim 20$ while
4906: it is measured to be $\sim 8$ on the lattice. Moreover, we have
4907: $\left.\frac{\Delta}{T^4}\right|_{T_{c,E}}\sim 4.8$ while the
4908: lattice obtains a value $\sim 0.85$. The much lower values obtained on the lattice are
4909: not surprising: Finite lattice sizes cut off existing long-range correlations at $T_{c,E}$.
4910:
4911: No result for the entropy density was directly reported in \cite{EnKaSaMo1982}.\vspace{0.1cm}\\
4912: \noindent\underline{SU(3) case:}\vspace{0.1cm}\\
4913: Here we discuss the results obtained in \cite{Bielfeld1996} with a
4914: Wilson action on the lattice of the largest
4915: time extension, $N_\beta=8$, and the results obtained
4916: in \cite{Deng1988,Brown1988}.
4917:
4918: In the vicinity of the transition point $T_{c,E}$
4919: the situation for both $\rho$ and $P$ is similar to the SU(2) case:
4920: Drastic differences between the lattice measurements and
4921: our results occur. Again, $P$ is negative close to $T_{c,E}$
4922: contradicting the positive values obtained with the integral method in
4923: \cite{Bielfeld1996}. A lattice simulation \cite{Deng1988} of $P$, which used the differential method,
4924: has reported negative pressure for $T$ shortly above the transition
4925: already in 1988. The most negative value of $P/T^4\sim -0.5$ obtained
4926: in \cite{Deng1988} very close to the phase transition is down by a factor
4927: of about $0.19$ as compared to our
4928: result at the electric-magnetic transition,
4929: see Figs.\,\ref{pressure} and \ref{pressureLat}. Again, this is
4930: explained by the finite lattice-size cutoff on physical long-range correlations.
4931: The lattice-result
4932: obtained with the integral method \cite{Bielfeld1996} is by construction
4933: positive definite, see Sec.\,\ref{DVI}, and thus
4934: it is {\sl unphysical}. For that reason we
4935: renounce a (useless) comparison of our results for pressure,
4936: energy density, and interaction measure with those obtained in \cite{Bielfeld1996}. In
4937: \cite{Deng1988} only the dependence of $\frac{P}{T^4}$ on the
4938: lattice coupling was presented. One can use Fig.\,\ref{pressureLat} and the temperature
4939: dependence of $\frac{P}{T^4}$ in Fig.\,\ref{B1} to
4940: gauge particular values of this quantity against temperature. (In both simulations
4941: \cite{Deng1988} and \cite{Bielfeld1996} the universal part
4942: of the two-loop perturbative $\beta$ function was used to relate lattice
4943: coupling to lattice spacing.) For example, a value of $\frac{P}{T^4}\sim 1.6$ in
4944: \cite{Deng1988} corresponds to a value $\frac{P}{T^4}\sim 1.5$ in \cite{Bielfeld1996}.
4945: The latter is associated with a temperature $T=3.2\,T_{c,E}$ by virtue of Fig.\,\ref{B1}.
4946: We have
4947: %**********
4948: \eqb
4949: \label{rhoN3}
4950: \left.\frac{P}{T^4}\right|_{T\sim 3.2\,T_{c,E}}\sim 2.2\,.
4951: \eqe
4952: %**********
4953: This is larger than the result obtained in \cite{Deng1988} and explained by the
4954: insufficient account of infrared correlations in the lattice simulation. These correlations generate masses for six
4955: out of eight gluon species, thus extra polarizations, and therefore a larger
4956: value for $\frac{P}{T^4}$.
4957:
4958: Our asymptotic ($\lambda_E=35$)
4959: values for $P$ and $\rho$ are
4960: %*****
4961: \eqb
4962: \label{PoverrhoSB3a}
4963: \left.\frac{P}{P_{SB}}\right|_{as}\sim 1.30\,,\ \ \ \ \ \ \left.\frac{\rho}{\rho_{SB}}\right|_{as}\sim 1.37\,,
4964: \eqe
4965: %*****
4966: where $\rho_{SB}=\frac{8}{15}\pi^2\,T^4=3\,P_{SB}$.
4967: Both values in Eq.\,(\ref{PoverrhoSB3a}) are close to the ratio $R=\frac{11}{8}=1.375$
4968: of the numbers of polarization in a free gluon gas, where six gluon species have a tiny mass, and in a free
4969: gluon gas where all gluon species are massless.
4970:
4971: The entropy density approaches zero for $T\searrow T_{c,M}$, see Fig.\,(\ref{ST3}).
4972: This expresses the fact that dual gauge modes
4973: are decoupled (infinite masses): The ground state strongly dominates the thermodynamics.
4974:
4975:
4976: \subsubsection{Differential versus integral method\label{DVI}}
4977:
4978: What are the reasons for the qualitative difference
4979: between the pressure-results obtained
4980: in \cite{Bielfeld1996,EngelsKarschScheideler1999} using the integral method and in
4981: \cite{Brown1988,Deng1988} using the differential method?
4982: While the differential method is based on the definition
4983: %*********
4984: \eqb
4985: \label{ptd}
4986: P=T\frac{\pd\ln Z}{\pd V}\,,
4987: \eqe
4988: %*********
4989: which is proper for a lattice of {\sl finite} volume $V$, the integral method assumes
4990: the thermodynamical limit $V\to\infty$ from the start.
4991: In this limit one has
4992: %*********
4993: \eqb
4994: \label{ptdL}
4995: P=T\frac{\ln Z}{V}\,,
4996: \eqe
4997: %*********
4998: and thus the pressure equals minus the free energy density.
4999: In Eqs.\,(\ref{ptd}) and (\ref{ptdL}) $Z$ denotes the partition function.
5000:
5001: The official reason for the introduction of the
5002: integral method, see for example \cite{EngelsFingberg1990}, was
5003: that one wanted to avoid the use of the
5004: imprecisely known $\beta$
5005: function in the strong-coupling regime of the theory. (Based on the definition in Eq.\,(\ref{ptd}),
5006: the $\beta$ function multiplies the sum of spatial and time plaquette averages in the
5007: expression for the pressure.) When using the definition in Eq.\,(\ref{ptdL}),
5008: the derivative of the pressure with respect to the bare coupling $\bar{\beta}$
5009: ($\bar{\beta}=\frac{6}{\bar{g}^2}$ for SU(3))
5010: can be expressed as an expectation over minus the sum of spatial and time-like plaquettes
5011: without the beta-function prefactor. Thus the pressure is, up to an unknown
5012: integration constant, obtained
5013: in terms of an integral of a sum of
5014: plaquette averages over $\beta$. The integration constant
5015: is chosen in such a way that the pressure vanishes at a temperature well
5016: below $T_c$. Instead of only integrating over minus the sum of spatial and
5017: time-like plaquette expectations an extra term was added to the
5018: {\sl integrand} \cite{Bielfeld1996,EngelsKarschScheideler1999}
5019: to assure that the pressure vanishes at $T=0$. The added term equals twice
5020: the plaquette expectation taken on a symmetric lattice (the expectation at $T=0$ ).
5021: We stress that this prescription does not
5022: follow from the definition in Eq.\,(\ref{ptdL}).
5023: Moreover, the assumption that $P=0$ for $T\sim 0.8\,T_{c}$ or
5024: so is a strong bias. (There are massless fermionic particles
5025: in the center phase which keep the total pressure
5026: positive for temperatures comparable to this value.)
5027:
5028: The results for $P(T)$ obtained when using the
5029: integral method show a rather large dependence on the spatial size and the time extent $N_\tau$ of the lattice
5030: \cite{Bielfeld1996}. We believe that this
5031: reflects the considerable deviation from the assumed
5032: thermodynamical limit for realistic lattice sizes. The problem
5033: was addressed in \cite{EngelsKarschScheideler1999} where a
5034: correction factor $r$ was introduced to relate $P$,
5035: obtained with the integral method, to $P$, obtained
5036: with the differential method. For a given value of $N_\tau$
5037: the factor $r$ was determined from the pressure-ratio at $\bar{g}=0$.
5038: Subsequently, this value of $r$ was used at {\sl finite} coupling $\bar{g}$
5039: to extract the spatial anisotropy
5040: coefficient $c_\sigma$ (essentially the $\beta$ function)
5041: by demanding the equality of the pressure obtained
5042: with the integral and the differential method.
5043: In doing so, twice the plaquette expectation at $T=0$ was, again, added to minus
5044: the sum of spatial and time-like plaquette expectations in the differential-method expression
5045: for the pressure. It may be questioned
5046: whether a simple correction factor $r$ does correctly
5047: account for finite-size effects and, if yes,
5048: whether it is justified to determine $r$ in the limit
5049: of noninteracting gluons. (The $c_\sigma$-values obtained in this way
5050: do not coincide with those obtained in \cite{Klassen1998}.)
5051: In addition, it seems that the imprecise knowledge of the $\beta$
5052: function, which contains information about fluctuations in the
5053: ultraviolet, is much less of a problem for a lattice simulation of the pressure
5054: than the missing infrared physics is (finite lattice size) \cite{Blum2004}.
5055:
5056: Using the universal part of the two-loop perturbative beta function in the differential method,
5057: negative values for the pressure were obtained for $T$ close to
5058: $T_c$ in \cite{Deng1988}. Moreover, a rapid approach of $\rho$ and
5059: $P$ to their respective free-gas limits was observed. This is in
5060: qualitative (but not quantitative) agreement
5061: with our results, see Figs.\,\ref{IM}, \ref{pressure}, and \ref{rho}.
5062:
5063:
5064: \section{Implications for particle physics and cosmology\label{Apps}}
5065:
5066: In this section we provide outlooks on the implications
5067: of the nonperturbative approach to SU(2) and
5068: SU(3) Yang-Mills thermodynamics in view of so-far unexplained phenomena
5069: in particle physics and cosmology. The way of
5070: how selected problems are addressed in this section
5071: is preliminary, mostly qualitative and thus should not be understood
5072: as the final say on the matter. Rather, we try to
5073: provide a certain amount of stimulus for
5074: future developments.
5075:
5076: \subsection{A Planck-scale axion: Cosmic coincidence today and $CP$ violation\label{PSA}}
5077:
5078: Among the gauge groups SU($N$) ($N$ finite) we regard SU(2) and
5079: SU(3) as particular due to their unique phase
5080: diagrams. We have come to appreciate that nature seems to prefer situations with a
5081: unique outcome. Thus we tend to believe that dynamics subject to a
5082: finite gauge symmetry, that is, dynamics below
5083: the Planck scale $M_P\sim 1.2\times 10^{19}\,$GeV,
5084: obeys the SU(2) or SU(3) gauge principle. A possible scenario would
5085: be that at $M_P$ an SU($N$) gauge symmetry ($N=\infty$) is
5086: dynamically broken into a four-dimensional low-energy manifestation
5087: involving several SU(2) and SU(3) factor groups, which can behave
5088: in an electric-magnetically dual way to one another, and into
5089: nonfluctuating gravity.
5090:
5091: This set-up may, in fact, be described by
5092: the low-energy sector of a bosonic string theory whose vacuum instability
5093: is resolved in terms of tachyon
5094: condensation in the presence of a D-brane
5095: \cite{Sen2000}. A low-energy Kaluza-Klein \cite{Kaluza1921,Klein1926}
5096: compactification of the Weyl-invariant and thus
5097: $d$-dimensional bosonic string theory (in a flat background $d=26$) to four
5098: dimensions yields gauge symmetries which are associated with the isometries of the $d-4$ dimensional
5099: compactification manifold: We assume that these isometries are products of
5100: SU(2) and SU(3) corresponding to a compactification
5101: manifold which, locally, is $S_3\times (S_3\times S_5)\times \cdots$ \cite{Steenrod1951,Aguilar1999}.
5102: In addition, there are
5103: a low-mass scalar dilaton field $\varphi$ and a massless
5104: antisymmetric tensor $B_{\mu\nu}$. While the former may drive the
5105: monopole condensation process within, say, an SU(2) theory of
5106: Yang-Mills scale $\Lambda_1\sim M_P$ and thus may trigger the
5107: inflation of the Universe (making it spatially flat) the latter may be
5108: responsible for adiabatically generated density
5109: perturbations \cite{Prokopec2005}.
5110:
5111: The scales of the SU(2) and SU(3)
5112: factors would dynamically bet set into a certain hierarchy:
5113: $\La_1\sim M_P>\La_2>\cdots>\La_{\tiny\mbox{CMB}}$. The scale $\La_{\tiny\mbox{CMB}}$
5114: is associated with an SU(2) theory that is not confining at the present temperature of the Universe:
5115: Being {\sl at} the electric-magnetic transition this theory
5116: generates the photon as its only massless excitation,
5117: see Sec.\,\ref{EWSB}. Being in its center phase at a temperature $T\sim \La_1\sim M_P$,
5118: $SU(2)_1$ generates fermions by re-heating (single and self-intersecting
5119: center-vortex loops). Because $SU(2)_1$ used to be part of SU($N=\infty$) and thus was mixing
5120: with the other factors these fermions couple to the gauge fields of these factors.
5121: Since higher-charge states (self-intersecting center-vortex loops) with opposite charges are generated in
5122: equal amounts we would expect that they quickly annihilate
5123: into charge-zero states (no self-intersections). (Re-creation of higher-charge states
5124: after annihilation is less likely due to the redshift of the spectrum by the
5125: rapid power-law expansion of the Universe furnished by a free-gas
5126: equation of state. The latter originates from the gauge-mode
5127: excitations of the other factors). Although the
5128: massless fermions (single center-vortex loops)
5129: do not couple to the propagating gauge fields of the other factors by naked gauge charges
5130: they do so by their dipole moments. Considering these massless fermions
5131: to be fundamental, there is a global, axial U(1)
5132: symmetry which, however, is dynamically broken \cite{AtiyahSinger1984} and anomalous
5133: due to the calorons of the other factors
5134: \cite{BellJackiw1969,Fujikawa1979}.
5135:
5136: Integrating out SU(2)$_1$'s massless fermions, the relevant composite field
5137: is a (canonically normalized) axion field $a$ whose
5138: coupling to the gauge fields of the other factors is
5139: %********
5140: \eqb
5141: \label{axialanomaly}
5142: {\cal L}_{a.a.}=\frac{a}{F}\, \sum_{i=2}^{\tiny{\mbox{CMB}}}\mbox{tr}\,\tilde{F}_{\mu\nu,i}F_{\mu\nu,i}
5143: \eqe
5144: %********
5145: where $F\sim M_P$ denotes the Peccei-Quinn scale \cite{PecceiQuinn1977}: The scale at
5146: which $SU(2)_1$'s fermions come into existence and which measure the magnitude of
5147: the Cooper-pair condensate involving the massless species. The sum is over the
5148: other factors, and $\tilde{F}_{\mu\nu,i}=\frac12\,\epsilon_{\mu\nu\alpha\beta}\,F_{\alpha\beta,i}$
5149: is the dual field strength. Eq.\,(\ref{axialanomaly}) represents
5150: a term in the action for the other factors (deconfining at $T\sim M_P$)
5151: which violates parity (P) and charge conjugation (C) symmetries.
5152:
5153: While $a$ would be a massless phase if the axial anomaly was absent and
5154: the axial U(1) only was broken dynamically the anomaly-induced coupling in
5155: Eq.\,(\ref{axialanomaly}) gives rise to an axion
5156: potential $V_a=\sum_{i=2}^{\tiny{\mbox{CMB}}}V_{a,i}$.
5157: The operator $\mbox{tr}\,\tilde{F}_{\mu\nu,i}F_{\mu\nu,i}$ measures the
5158: average topological charge density carried by
5159: {\sl propagating} gauge fields. This is a conserved quantity for $T\gg \La_i$, and
5160: thus it is independent of temperature. Integrating
5161: over topologies and accounting for dimensional transmutation
5162: we have
5163: %*********
5164: \eqb
5165: \label{axmass}
5166: V_{a,i}\sim \left\{\begin{array}{c}\left(1-\cos\frac{a}{F}\right)\,\La_i^4\,,\ \
5167: \ \ \ (\mbox{theory $i$ in electric or magnetic phase})\,,\\
5168: 0\,\ \ \ \ \ (\mbox{theory $i$ in center phase})\,.\end{array}\right.
5169: \eqe
5170: %**********
5171: For $a$ smaller (but not much smaller) than $F$ the cosine in Eq.\,(\ref{axmass}) can be expanded,
5172: and the axion mass-squared at $T\sim M_P$ reads
5173: %*********
5174: \eqb
5175: \label{axmassreally}
5176: m_a^2\sim\frac{1}{F^2}\,\sum_{i=2}^{\tiny{\mbox{CMB}}}
5177: \La_i^4\sim \sum_{i=2}^{\tiny{\mbox{CMB}}}\frac{\La_i^4}{M_P^2}\,.
5178: \eqe
5179: %**********
5180: As the temperature of the Universe falls below $\La_2$ the
5181: associated theory fails to contribute to the
5182: axion mass-squared and so forth.
5183:
5184: Now in an expanding Friedmann-Robertson-Walker Universe the (spatially homogeneous)
5185: axion field $a$ satisfies the following equation
5186: of motion
5187: %********
5188: \eqb
5189: \label{hubbleexp}
5190: \ddot{a}+3\,H\dot{a}+m_a^2\,a=0\,.
5191: \eqe
5192: %********
5193: The Hubble parameter $H\equiv\frac{\dot{R}}{R}=\sqrt{\frac{8\pi\rho(T)}{3\,M_P^2}}$
5194: is determined by the energy density $\rho(T)=\sum_{i=2}^{\tiny{\mbox{CMB}}}\rho_i(T)$ of
5195: the Universe. At $T\sim M_P$ this energy density is given by the Stefan-Boltzmann
5196: limit of the theories $i=2,\cdots,\mbox{CMB}$ if the hierarchy between $M_P$ and $\Lambda_2$
5197: is sufficiently large. (The SU(2) theory with $i=1$ went
5198: center, subsequently experienced a strong dilution of
5199: its massive excitations, and does not produce a ground-state
5200: contribution to $\rho$ at $T\sim M_P$.) We have $\rho_i(T)=\frac{4\pi^2}{15}\,T^4$ (SU(2)) and
5201: $\rho_i(T)=\frac{11\pi^2}{15}\,T^4$ (SU(3)). Thus $H$ is radiation-dominated
5202: at $T\sim M_P$ and $H^2\gg m_a^2$. But this means
5203: that $a$ is frozen to the slope of its potential with an
5204: amplitude $a\sim M_P$. For $T\sim \La_2$ the Universe's energy density remains
5205: radiation-dominated. The axion mass-squared, however, is reduced by the value
5206: $\frac{\La_2^4}{M_P^2}$ since the theory with $i=2$ went center.
5207: If $\La_2\ll M_P$ then the mass-squared $\sim
5208: \sum_{i=3}^{\tiny{\mbox{CMB}}}\frac{\La_i^4}{\La_2^2}$ of the axion generated by the theory with
5209: $i=2$ is much larger than $H^2$. By Eq.\,(\ref{hubbleexp}) this means
5210: that this axion rapidly relaxes to the minimum of its potential
5211: and thus is irrelevant for subsequent cosmology. Moreover, the
5212: fermions generated by the center transition of the theory
5213: with $i=2$ exhibit a large asymmetry in fermion number.
5214: This is qualitatively true since
5215: all three Sakharov conditions \cite{Sakharov1967} are satisfied: (i) the center transition is nonthermal (Hagedorn),
5216: (ii) there is a local violation of fermion number since fermions
5217: are nonlocal objects, and (iii) the generation of
5218: fermions takes place in the presence of
5219: CP violation (the frozen-in Planck-scale axion $a$).
5220:
5221: This goes on until $T=T_{\tiny\mbox{CMB}}$
5222: where the last theory in the chain, SU(2)$_{\tiny\mbox{CMB}}$, is close to its
5223: center transition (more specifically, at the
5224: electric-magnetic phase boundary today), see Fig.\,\ref{Fig40}.
5225: %***********************
5226: \begin{figure}
5227: \begin{center}
5228: \leavevmode
5229: %\epsfxsize=9.cm
5230: \leavevmode
5231: %\epsffile[80 25 534 344]{}
5232: \vspace{5.0cm}
5233: \special{psfile=Fig-33.ps angle=0 voffset=-150
5234: hoffset=-120 hscale=50 vscale=50}
5235: \end{center}
5236: \caption{The fate of a Planck-scale axion along the Universe's evolution. At temperatures sizably
5237: larger than $T_{\tiny\mbox{CMB}}$ the axion is frozen to the slope of
5238: its potential by cosmological friction ($H\gg m_a$), for $T\sim T_{\tiny\mbox{CMB}}$
5239: axion mass and Hubble parameter become comparable: The axion starts to roll
5240: down its potential.\label{Fig40}}
5241: \end{figure}
5242: %************************
5243: Here $H^2$ is dominated by the ground-state
5244: energy. We have
5245: %*********
5246: \eqb
5247: \label{coinc}
5248: H^2\sim \frac{8\pi}{3}\frac{4\pi T_{\tiny\mbox{CMB}}\La_{\tiny\mbox{CMB}}^3+\rho_V+\rho_K}{M_P^2}
5249: \stackrel{\sim}>m_a^2\sim \frac{\La_{\tiny\mbox{CMB}}^4}{M_P^2}\,.
5250: \eqe
5251: %*********
5252: In Eq.\,(\ref{coinc}) $\rho_V$ denotes the energy density
5253: associated with the value of the axion potential at $T_{\tiny\mbox{CMB}}$, and
5254: $\rho_K$ is an energy density due to axion rolling.
5255: Both contributions are comparable since $H$ is not much larger than $m_a$,
5256: compare with Eq.\,(\ref{hubbleexp}). The formerly frozen-in axion field $a$
5257: slowly starts to roll down its potential.
5258: While $\rho_V$ has an equation of state $\rho_V=-P_V$ the kinetic contribution
5259: $\rho_K$ is associated with an equation of state $P_K=0$ which is
5260: the same as that of nonrelativistic matter.
5261:
5262: Cosmic coincidence may have an explanation in terms of a Planck-scale axion and an SU(2) Yang-Mills
5263: theory of scale $\La_{\tiny\mbox{CMB}}$ comparable to the present
5264: temperature of the CMB. For related ideas see \cite{Wilczek2004}.
5265: The alert reader may object that the
5266: $\theta$ angle in Quantum Chromodynamics (QCD) is constrained
5267: to be an extremely small number by a measurement
5268: of the neutron's electric dipole moment. On the other hand, the mass
5269: of the $\eta^\prime$ is much larger than the pion mass. Thus one would
5270: conclude that the Planck-scale axion is
5271: irrelevant in the former while it is relevant
5272: in the latter case. What is
5273: the resolution of this puzzle? The electric dipole moment of the
5274: neutron is measured with a photon of momentum much smaller than
5275: the QCD confinement scale. Thus this photon does not probe a
5276: phase of QCD with propagating gauge bosons: The operator $\tilde{F}_{\mu\nu}
5277: F_{\mu\nu}$ has a vanishing expectation. The $\eta^\prime$, on the other hand,
5278: is generated in a scattering process involving propagating gluons:
5279: Inside the vertex the operator $\tilde{F}_{\mu\nu}F_{\mu\nu}$ has a finite
5280: expectation.
5281:
5282: We will see in Sec.\,\ref{EWSB} that the ground-state
5283: contribution of SU(2)$_{\tiny\mbox{CMB}}$ in the {\sl absence} of the Planck-scale
5284: axion is small in comparison with the measured value of today's cosmological constant.
5285: The scenario outlined above does not yet explain the origin of {\sl clustering} dark matter as it is
5286: observed in the anomalous rotation curves of galaxies but we
5287: will see below that the decoupled TLH modes of SU(2)$_{\tiny\mbox{CMB}}$ are
5288: candidates for this form of matter.
5289:
5290:
5291: \subsection{Electroweak sector of the Standard Model: Nature of leptons,
5292: electroweak symmetry breaking, masses of intermediary vector bosons,
5293: intergalactic magnetic fields, and solar wind\label{EWSB}}
5294: Here we would like to propose a formulation of the electroweak
5295: sector of the Standard Model in terms of pure SU(2) Yang-Mills theories.
5296: \vspace{0.1cm}\\
5297: \noindent\underline{SU(2)$_{\tiny\mbox{CMB}}$:}\vspace{0.1cm}\\
5298: Let us first discuss the U(1)$_Y$ factor of the electroweak gauge
5299: group SU(2)$_W\times$U(1)$_Y$. We claim that this factor
5300: is the unbroken subgroup of an SU(2) Yang-Mills
5301: theory of scale comparable to that of the CMB
5302: temperature: $T_{\tiny\mbox{CMB}}=2.728\,K=2.351\times 10^{-4}\,$eV.
5303: Only one point in the phase diagram of this theory exists, the boundary between the electric and magnetic phases,
5304: where this claim is in accord with observations: The photon
5305: is unscreened and practically massless ($m_\gamma<10^{-14}\,$eV from a precision
5306: measurement of the Coulomb potential \cite{Williams1971}, see \cite{Dvali2003} for a
5307: discussion on why stronger bounds are unreliable), see Fig.\,\ref{looppress}.
5308: Thus we identify $T_{\tiny\mbox{CMB}}=T_{c,E}$. Notice that
5309: isolated charges in the electric phase of SU(2)$_{\tiny\mbox{CMB}}$
5310: have a dual interpretation: What is a magnetic monopole
5311: in SU(2) is an electrically charged particle w.r.t. U(1)$_Y$.
5312:
5313: The energy density $\rho^{gs}$ of the ground-state at $T_{\tiny\mbox{CMB}}$
5314: is $\rho^{gs}=4\pi\,T_{\tiny\mbox{CMB}}\La_E^3=4\pi\times (2.351\times 10^{-4}\,\mbox{eV})\,\La_E^3=
5315: 2\lambda_{c,E}\La_E^4=27.7\,\La_E^4$. Moreover, we have
5316: $\La_E=1.065\times 10^{-4}\,$eV. Substituting this into $\rho^{gs}$, we have
5317: $\rho^{gs}=\left(2.444\times 10^{-4}\,\mbox{eV}\right)^4$. This is about 0.36\% of the
5318: commonly accepted value of today's dark energy density $(10^{-3}\,\mbox{eV})^4$.
5319: The dominating, missing part would be generated by a
5320: slowly rolling Planck-scale axion, see Sec.\,\ref{PSA}.
5321:
5322: An immediate question to answer is why the masslessness and
5323: the unscreened propagation of the photon is a singled-out situation. The answer to this
5324: question is encoded in Fig.\,\ref{rho}: The energy density of an SU(2) Yang-Mills theory
5325: dips at the electric-magnetic phase boundary. On the electric side this is explained by
5326: the thermodynamical decoupling of TLH modes, on the
5327: magnetic side an extra polarization, which costs energy, needs to be generated
5328: for the photon. To facilitate the jump in energy density for SU(2)$_{\tiny{CMB}}$
5329: to reach the magnetic phase thermal equilibrium needs to be violated by the
5330: eventually fast rolling Planck-scale axion. The photon acquires
5331: mass and the ground state becomes superconducting (electrically charged monopoles condense).
5332: It is suggestive that the occurrence of intergalactic magnetic fields is
5333: related to the Universe being slightly out of thermal
5334: equilibrium due to (slow) axion rolling today.
5335:
5336: At $T_{\tiny\mbox{CMB}}$ TLH modes decouple
5337: thermodynamically. Recall that their mass is given
5338: by $m_{\tiny\mbox{TLH}}=2e\,|\phi|$ and $e_{dec}=\infty$ at
5339: $T_{c,E}=T_{\tiny\mbox{CMB}}$. In the real world $e_{dec}$ is
5340: large but not infinite because SU(2)$_{\tiny\mbox{CMB}}$ is not
5341: the only Yang-Mills theory in the Universe. Taking
5342: $e_{dec}\sim 10^6$, say, the mass of a decoupled TLH mode is
5343: $m_{\tiny\mbox{TLH}}\sim 57\,$eV. (The reason why we chose this
5344: value for $e_{dec}$ is mildly justified by our
5345: discussion of the gauge group SU(2)$_e$ below.) Since the two TLH modes
5346: do not interact and thus are stable (no decay
5347: into light fermions is possible because SU(2)$_{\tiny\mbox{CMB}}$
5348: is not in its center phase yet) they yield a tiny contribution to clustering dark matter.
5349:
5350: Finally, we would like to make a remark concerning the observed
5351: large-angle anomaly in the temperature-(electric)polarization
5352: cross correlation seen by WMAP \cite{WMAP2003}. The standard explanation
5353: is that this effect is generated by CMB photons scattering off electric charges which are
5354: released by an early re-ionization of the interstellar medium at
5355: redshift $z\sim 10-20$. We would like to propose that CMB photons scatter off electrically
5356: charged and dilute monopoles at temperatures
5357: $>T_{\tiny\mbox{CMB}}$ which are condensed at $T_{\tiny\mbox{CMB}}$.
5358: Since static magnetic (electric with respect to SU(2))
5359: fields are completely screened in the
5360: photon propagator deep inside the electric phase of SU(2)$_{\tiny\mbox{CMB}}$, see Eq.\,(\ref{Pi00limit}),
5361: the effect should be weaker in the temperature-(magnetic)polarization
5362: cross correlation.
5363: \vspace{0.1cm}\\
5364: \noindent\underline{SU(2)$_e \times$ SU(2)$_\mu\times$ SU(2)$_{\tau}$:}\vspace{0.1cm}\\
5365: To relate the existence and the interactions of leptons, as they are described by the
5366: electroweak sector of the Standard Model (SM), of which Quantum Electrodynamics (QED)
5367: is an integral part, with
5368: pure SU(2) Yang-Mills dynamics is motivated by the following observations.
5369: (i) The masses of charged leptons are unexplained parameters in the SM.
5370: In particular, their small values on the scale of their apparent pointlikeness is
5371: unexplained. (ii) No deeper explanation
5372: for the value of the magnetic dipole moment of a charged
5373: lepton other than that following from the Dirac equation and small radiative
5374: corrections is given. (iii) There are experimental indications
5375: that scattering processes involving the electron or the positron do not obey the
5376: QED predictions if the momentum transfer is close to the mass of a
5377: charged lepton. (iv) No Higgs particle has been observed experimentally up to a hypothetical
5378: Higgs mass of $\sim 115$\,GeV suggesting that electroweak
5379: symmetry breaking takes place by a different mechanism than assumed in the SM.
5380: (v) The naive ground state of the SM generates a
5381: cosmological constant which is many orders of magnitude
5382: larger than the observed value.
5383:
5384: Points (i), (ii), (iv), and (v) are undisputed facts. To see that there is some truth to point (iii)
5385: we present experimental results. In Fig.\,\ref{Fig45} a plot of the ratio of experiment
5386: to theory of the wide-angle $e^+e^-$ pair-creation cross section through $\gamma$ scattering off of the
5387: field of a carbon nucleus is shown as a function
5388: of the invariant mass $M$ of the created lepton pair.
5389: %***********************
5390: \begin{figure}
5391: \begin{center}
5392: \leavevmode
5393: %\epsfxsize=9.cm
5394: \leavevmode
5395: %\epsffile[80 25 534 344]{}
5396: \vspace{6.4cm}
5397: \special{psfile=Fig-34.ps angle=0 voffset=-190
5398: hoffset=-120 hscale=50 vscale=50}
5399: \end{center}
5400: \caption{Ratio of experiment
5401: to theory of the wide-angle $e^+e^-$ pair-creation cross section by $\gamma$ scattering off the
5402: field of a carbon nucleus as a function of the invariant mass $M$ of
5403: the lepton pair. Plot taken from \protect\cite{Alvensleben1968}.\label{Fig45}}
5404: \end{figure}
5405: %************************
5406: Notice the substantial deviation from unity for
5407: $M\sim 2\,m_\mu\cdots 4\,m_\mu\sim (210\cdots 420)$\,MeV. Notice also that for $M>500\,$MeV theory
5408: and experiment do agree. A much more drastic deviation within
5409: the same kinematic regime was seen earlier in \cite{Blumenthal1966}. This, however, was not
5410: confirmed in \cite{Alvensleben1968}.
5411:
5412: In Fig.\,\ref{Fig50}
5413: a plot of the differential cross section for $e^-e^-$ scattering (electron incident on an atomic
5414: electron of a target) at a fixed fraction $\nu=0.5$ of the incident kinetic energy $E_{\tiny\mbox{kin}}$
5415: transferred in the collision is shown for $0.6\,\mbox{MeV}\le E_{\tiny\mbox{kin}}\le 1.7\,\mbox{MeV}$ taken with
5416: a $270^o$ apparatus (left panel) and for $0.6\,\mbox{MeV}\le E_{\tiny\mbox{kin}}\le 1.2\,\mbox{MeV}$ taken with
5417: a $180^o$ apparatus (right panel) \cite{Ashkin1953}. The solid line indicates the theoretical result
5418: obtained by using the M\o ller formula.
5419: %***********************
5420: \begin{figure}
5421: \begin{center}
5422: \leavevmode
5423: %\epsfxsize=9.cm
5424: \leavevmode
5425: %\epsffile[80 25 534 344]{}
5426: \vspace{5.4cm}
5427: \special{psfile=Fig-35.ps angle=0 voffset=-150
5428: hoffset=-220 hscale=60 vscale=60}
5429: \end{center}
5430: \caption{M\o ller scattering of electrons, for an explanation see text. Taken from \protect\cite{Ashkin1953}.\label{Fig50}}
5431: \end{figure}
5432: %************************
5433: Notice the agreement with QED for large values of $E_{\tiny\mbox{kin}}$ (in particular in the
5434: left panel of Fig.\,\ref{Fig50}). In \cite{Scott1951} the differential cross
5435: section for M\o ller scattering was measured for $E_{\tiny\mbox{kin}}=15.7\,$MeV
5436: as a function of the scattering angle and found to be in agreement with QED on the
5437: 0.4\% error level. The disagreement at the lowest
5438: value $E_{\tiny\mbox{kin}}=0.6\,$MeV in Fig.\,\ref{Fig50}, which corresponds to a center-of-mass
5439: energy of about $2.5\,m_e$, is conspicuous.
5440:
5441: If the Yang-Mills scales of the factors SU(2)$_e$, SU(2)$_\mu$, and SU(2)$_\tau$ are about $m_e$,
5442: $m_\mu$, and $m_\tau$, respectively, then the masses
5443: of the charge-one states in the center phase of each theory, see Fig.\,\ref{intersect}, are determined to
5444: be these values. Since $T_{\tiny\mbox{CMB}}\ll m_e, m_\mu, m_\tau$ these theories
5445: are in their center phases. By looking at Fig.\,\ref{intersect} a $g$-factor of two
5446: is imperative by the asignment of angular momentum one half in the presence of one unit of (electric) center
5447: flux in the vortex loop. The latter generates the lowest nonvanishing quantum of
5448: magnetic moment (Bohr magneton). Notice that Fig.\,\ref{intersect} provides for an
5449: intuitive manifestation of the concept of spin-1/2: Inside the intersection core the center flux
5450: is diverted to the right above and to the left below such that an eddy is generated. The latter carries the electric
5451: charge of the soliton.
5452:
5453: Let us now give some qualitative arguments why the gauge group
5454: SU(2)$_{\tiny\mbox{CMB}}\times$SU(2)$_e\times$SU(2)$_\mu\times$SU(2)$_\tau$
5455: together with a Planck-scale axion may be a viable candidate to describe the phenomenology
5456: of electroweak interactions. Why do the photon (SU(2)$_{\tiny\mbox{CMB}}$) and the massive intermediate
5457: vector bosons (magnetic and electric phase of SU(2)$_e$) couple
5458: to the charge and/or the magnetic moment of charge-one and charge-zero states?
5459: The answer to this question is rooted in the symmetry
5460: breakdown at $T\sim M_P$ where the factors SU(2)$_{\tiny\mbox{CMB}}$,
5461: SU(2)$_e$, SU(2)$_\mu$, and SU(2)$_\tau$ were generated out
5462: of one large gauge group whose gauge bosons where mixing. (In the SM the mixing of the 'photon' of U(1)$_Y$ with that
5463: of SU(2)$_W$ is parametrized by the Weinberg
5464: angle $\theta$ with $\sin^2\theta=0.23$.) The interaction of the photon of
5465: SU(2)$_{\tiny\mbox{CMB}}$ with the electrically charged soliton of, say,
5466: SU(2)$_e$ thus is furnished by an adiabatic rotation into the
5467: (massive) photon of SU(2)$_e$ when approaching the charge of the latter and
5468: an adiabatic back-rotation into the photon of
5469: SU(2)$_{\tiny\mbox{CMB}}$ after the interaction has taken place. Where are the higher
5470: charge states? These states are instable by repulsion mediated by the photon of
5471: SU(2)$_{\tiny\mbox{CMB}}$, and thus they are very broad. The density of these states, however,
5472: is over-exponentially rising. Why do we only see the structure of a charged lepton
5473: in scattering experiments with a center-of-mass energy $\sqrt{s}$ comparable to
5474: the mass of the lepton, see Figs.\,\ref{Fig45},\ref{Fig50}? Radial excitations of a 't
5475: Hooft monopole have been investigated in \cite{ForgasVolkov2003}. The first
5476: excited level is comparable to twice the mass of the monopole ground state.
5477: This must semi-quantitatively also hold for a $Z_2$ monopole (self-intersection of a center-vortex loop).
5478: For $\sqrt{s}\ll m_e, m_\mu, m_\tau$ the $Z_2$ monopole is not excitable, a QED
5479: point-particle description holds. For $m_Z\gg\sqrt{s}\gg m_e, m_\mu, m_\tau$,
5480: where $m_Z\sim 90\,$GeV is the mass
5481: of the $Z$ boson, the energy deposited
5482: into the vertex is converted into a large entropy carried by the Hagedorn spectrum
5483: of instable states. The latter protect the $Z_2$ monopole against radial excitations, a QED
5484: point-particle description again holds. For $\sqrt{s}\sim 2\,m_e, 2\,m_\mu, 2\,m_\tau$ the $Z_2$
5485: monopole is excited radially: A QED point-particle description fails.
5486: For $\sqrt{s}\gg m_Z$ the Hagedorn phase boundary of SU(2)$_e$ is locally
5487: overcome (the $Z$ boson is interpreted as the decoupled dual gauge mode on the magnetic
5488: side of the magnetic-center phase boundary, the $W^\pm$ bosons as the decoupled gauge modes on
5489: the electric side of the electric-magnetic phase boundary): The multiplicity of final states should be in stark
5490: contradiction to the SM prediction (we expect charge
5491: nonconservation in such processes). Where does the parity violation come from? This is an intermediate
5492: consequence of the existence of a Planck-scale axion. What is the nature of the charge-zero state
5493: (single center-vortex loop)? The mass of this state for SU(2)$_e$
5494: is roughly given by $m_\nu\sim \frac{m_e}{g_{dec}}$, compare with Eq.\,(\ref{EANOvortex}).
5495: Moreover, the mass of the $Z$ boson is given as
5496: $m_Z\sim g_{dec}\,m_e$. From the experimentally known values $m_e\sim 5\times 10^5$\,eV and
5497: $m_Z\sim 9\times 10^{10}$\,eV we thus have $g_{dec}\sim \frac{9}{5}\times 10^5$ and therefore
5498: $m_\nu\sim \frac{25}{9}\,$eV. This is close to the upper bound for the mass of the
5499: electron neutrino obtained from a tritium $\beta$ decay
5500: experiment: $m_\nu<2.3\,$eV \cite{Krauss2004}. Thus a single
5501: center-vortex loop of SU(2)$_e$ viably is a candidate for the electron neutrino. Notice that
5502: this soliton has no antiparticle: Neutrinos need to be of
5503: the Majorana type in accord with the successful search for
5504: neutrinoless double beta decay \cite{Klapdor2004}. A similar
5505: situation holds for SU(2)$_\mu$ and SU(2)$_\tau$. We expect the masses of their
5506: intermediary vector bosons $m_{Z^\prime}, m_{W^{\prime,\pm}}$ and
5507: $m_{Z^{\prime\prime}}, m_{W^{\prime\prime,\pm}}$ to scale with
5508: their Yang-Mills scales $m_\mu$ and $m_\tau$ and large values of the
5509: gauge couplings at the respective phase boundaries.
5510: Thus there are very weak and very, very weak interactions in addition
5511: to the weak interactions which, however, will be very hard to detect experimentally.
5512: (To detect, say, $m_{Z^\prime}$ directly would need a
5513: center-of-mass energy in $e^+e^-$ annihilation which should at least be
5514: two-hundred times $m_Z$.) A remark concerning point (v) is in order:
5515: Since SU(2)$_e \times$ SU(2)$_\mu\times$ SU(2)$_{\tau}$ are in their
5516: center phases at present their contribution
5517: to the ground-state energy density and pressure of the Universe is nil. Above, we have computed
5518: the contribution arising from SU(2)$_{\tiny\mbox{CMB}}$ when assuming the
5519: Planck-scale axion to be absent.
5520:
5521: Let us make a short remark on the solar wind. This particle flux is mainly composed
5522: of protons (about $3\times10^{43}$ protons depart
5523: annually from the solar surface \cite{Manuel2004}). If the conservation of
5524: electric charge, which is a built-in feature of the SM,
5525: would hold then the sun would continuously acquire
5526: negative charge: A disastrous implication
5527: for earth's orbit would arise. The problem is resolved by the
5528: observation that in the solar core temperatures are
5529: greater than $m_e$. Electronic charge, however, is absent in the magnetic
5530: or electric phase of SU(2)$_e$. According to the phase diagram of
5531: SU(2)$_e$ the solar core contains a superconducting mantel
5532: (magnetic phase, Bose condensate of electric monopole-antimonopole pairs)
5533: whose negative pressure together
5534: with gravity balances the positive thermal pressure of the
5535: innermost core (electric phase) where fusion takes place.
5536: Within the core region there is a clear dominance of
5537: positive charge which the sun deposes off by means
5538: of the solar wind. A superconducting core of the sun is also
5539: demanded in \cite{Manuel2004} for other reasons.
5540:
5541: We conclude this section by stressing that no fundamental Higgs field is needed
5542: to break the weak symmetry SU(2)$_W$ or SU(2)$_e$. In contrast to the SM,
5543: where this symmetry breaking is complete by a nonvanishing expectation of a
5544: fundamental Higgs field, the dynamical breakdown of SU(2)$_e$ proceeds in a two-stage, Higgs-particle free
5545: way: SU(2)$_e\to$U(1) (electric phase; adjoint nonfluctuating Higgs field)
5546: and U(1)$_D\to 1$ (magnetic phase; complex nonfluctuating Higgs field). Moreover,
5547: the electron and its neutrino are stable solitons in the center phase of SU(2)$_e$.
5548:
5549:
5550: \subsection{Quantum chromodynamics: Quark confinement and fractional quantum Hall
5551: effect\label{QCD}}
5552:
5553: In this section we pursue an admittedly speculative and not very matured idea
5554: about the nature of quarks and their interactions.
5555:
5556: Quantum chromodynamics (QCD) is an integral part of the SM. QCD is the gauge
5557: theory of strong interactions: Pointlike current quarks, which are spin-1/2 fermions of to-be-measured
5558: masses, are fundamentally charged under the gauge group SU(3)$_C$ and
5559: interact by the exchanges of massless gluons,
5560: the gauge bosons of SU(3)$_C$. The latter interact with one another according
5561: to a pure Yang-Mills action. The electric charges of
5562: quarks are 2/3 or $-1/3$. Let us only discuss the three quark flavors
5563: of lowest mass $m_u=(3\cdots 5)\,$MeV (charge 2/3), $m_d=(5\cdots 7)\,$MeV (charge $-1/3$),
5564: and $m_s=(100\cdots 140)\sim \,$MeV (charge $-1/3$) (all $\overline{\mbox{MS}}$ scheme, results depend on
5565: the renormalization point).
5566:
5567: Since leptons are likely
5568: to be the stable solitons in the center phases of pure SU(2) Yang-Mills theories
5569: it is tempting to speculate that quarks are related to the
5570: charge-one solitons (center-vortex loops with one self-intersection,
5571: spin-1/2 fermions) in the center phases of
5572: pure SU(3) Yang-Mills theories. If we assign an SU(3)$_u$ and SU(3)$_d$ theory of Yang-Mills
5573: scale $\sim m_u$ and $m_d$ to the quark flavors $u$ and $d$, respectively, then we need
5574: to understand why these quarks are confined and why the
5575: electric charge appears to be $2/3$ or $-1/3$.
5576: A plausible way of generating quark confinement would be to
5577: add an additional SU(3) Yang-Mills theory of scale, say $\Lambda=140\,$MeV
5578: which, however, is a magnetic dual to the other SU(3) theories.
5579: A center-vortex condensate of this theory constrains
5580: the (color)electric flux between a $u$ or $d$ quark and a $u$ or $d$
5581: antiquark into a tube and thus confines. The additional SU(3)
5582: theory also has charge-one solitons in its center phase
5583: which are confined by the center-vortex condensates of the
5584: theories SU(3)$_u$ and SU(3)$_d$.
5585: It is tempting to interpret these solitons as
5586: strange quarks $s$ and thus to invoke
5587: the label SU(3)$_s$.
5588:
5589: What about the electric quark charges? Due to confinement
5590: the trajectories of each quark flavor are forced
5591: onto a more or less two-dimensional
5592: spherical surface if the ground state
5593: of a given hadron is considered. A flux dual
5594: to the flux in the confining tube is readily available in the
5595: center-vortex condensate being responsible for confinement. This is the set-up for the occurrence of the
5596: fractional quantum Hall effect: Quarks that would be integer
5597: charged spin-1/2 fermions in the absence of the dual fluxes form
5598: bound states with these fluxes. Bound states with three dual flux
5599: quanta are bosons, and thus they condense. Excitations above this condensate
5600: have fractional electric charge. For a thorough discussion
5601: of this phenomenon, see \cite{Laughlin1998}.
5602:
5603: Again, all of what was said in this section is preliminary. We would like to stress though that
5604: the above scenario has the potential to explain why the equation of state in hydrodynamical simulations
5605: of the elliptic flow measured in ultra-relativistic heavy-ion collisions at RHIC seems to be
5606: so close to the free-gas limit despite the fact
5607: that strong correlations thermalize the system very rapidly
5608: \cite{Shuryak2004}. (At $T_c\sim 170\,$MeV the two theories SU(3)$_u$ and SU(3)$_d$ are deep
5609: inside their electric phases while SU(3)$_s$ is just above the
5610: electric-magnetic transition.)
5611:
5612:
5613: \section{Conclusions\label{CO}}
5614:
5615: We have developed a nonperturbative approach to SU(2) and SU(3) Yang-Mills
5616: thermodynamics. The formation of a macroscopic, adjoint, and nonfluctuating
5617: Higgs field in the deconfining (electric) phase of each theory, which involves the (admissible part of the)
5618: moduli space of a caloron-anticaloron system, the Bose
5619: condensation of thermalized magnetic monopole-antimonopole systems into a macroscopic,
5620: nonfluctuating complex field in a preconfining (magnetic) phase, and the
5621: nonthermal condensation of systems composed of a center-vortex and an
5622: anti-center-vortex in a confining (center) phase were shown. A change of
5623: the statistics of the excitations from bosonic to fermionic was observed across
5624: the last phase transition. The degeneracy of the ground state with
5625: respect to a (global) electric $Z_2$ (SU(2)) and $Z_3$ (SU(3))
5626: symmetry was observed in the electric phase, and the uniqueness
5627: of the ground state with respect to these symmetries was derived in
5628: the magnetic phase. Moreover, the nature of the phase transitions,
5629: electric-magnetic and magnetic-center, was clarified. The evolution of
5630: thermodynamical quantities with temperature
5631: was computed within the electric phase and the magnetic phase, and the
5632: density of fermionic states was estimated in the center phase.
5633:
5634: It did not escape the author's attention that the results obtained may resolve a number of
5635: long-standing problems in particle physics and cosmology.
5636:
5637:
5638:
5639: \section*{Acknowledgments}
5640: The author would like to thank B. Garbrecht, H. Gies, Th. Konstandin, T. Prokopec, H. Rothe, K. Rothe,
5641: M. Schmidt, I.-O. Stamatescu, and W. Wetzel for very helpful, continuing discussions.
5642: Important support for numerical calculations was provided by J. Rohrer and is
5643: thankfully acknowledged. The author acknowledges vivid discussion with Francesco Giacosa
5644: and Markus Schwarz. In particular, I am grateful to Francesco Giacosa for pointing out the need for
5645: a modification of the evolution equations for the effective gauge couplings which
5646: was overlooked in the pervious version. Very useful discussions with P. van Baal,
5647: E. Gubankova, J. Moffat, J. Polonyi, D. Rischke, and F. Wilczek and
5648: illuminating conversations with D. B\"odeker, R. Brandenberger, G. Dunne, Ph. de Forcrand, A. Guth,
5649: F. Karsch, A. Kovner, M. Laine, H. Liu, C. Nunez, J. Pawlowski, R. D. Pisarski, K. Rajagopal, K. Redlich, E. Shuryak,
5650: D. T. Son, A. Vainshtein, J. Verbaarschot, and F. Wilczek are gratefully acknowledged.
5651: A very useful correspondence with O. Manuel
5652: on solar models is thankfully acknowledged. The warm hospitality of the Center for
5653: Theoretical Physics at M.I.T, where part of this research was carried out
5654: (sponsored by Deutsche Forschungsgemeinschaft), is thankfully acknowledged.\\
5655: This paper is dedicated to my family and in particular to my
5656: wife Karin Thier. Thank you for your continuing encouragement, your persistent
5657: moral support, and your unconditional love.
5658:
5659:
5660: \bibliographystyle{prsty}
5661:
5662: \begin{thebibliography}{10}
5663:
5664: \bibitem{Higgs2000}
5665: G. Gomez-Ceballos {\sl et al.}, Int. J. Mod. Phys. A {\bf 16S1B}, 839 (2001).
5666:
5667: \bibitem{RHIC2003}
5668: U. W. Heinz and P. F. Kolb, Nucl. Phys. A {\bf 702}, 269 (2002).
5669:
5670: \bibitem{ShuryakTeaney2001}
5671: D. Teaney, J. Lauret, and E. V. Shuryak, nucl-th/0110037.
5672:
5673: \bibitem{Cobe}
5674: C. L. Bennett {\sl et al.}, Astrophys. J. {\bf 464}, L1 (1996).\\
5675: K. M. Gorski {\sl et al.}, Astrophys. J. {\bf 464}, L11 (1996).\\
5676: G. Hinshaw {\sl et al.}, Astrophys. J. {\bf 464}, L17 (1996).
5677:
5678: \bibitem{Perlmutter1998}
5679: S. Perlmutter {\sl et al.}, Astrophys. J. {\bf 517}, 565 (1999).\\
5680: S. Perlmutter {\sl et al.}, Bull. Am. Astron. Soc. {\bf 29}, 1351 (1997).
5681:
5682: \bibitem{Schmidt1998}
5683: A. G. Riess {\sl et al.}, Astron. J. {\bf 116}, 1009 (1998).
5684:
5685: \bibitem{WMAP2003I}
5686: D. N. Spergel {\sl et al.}, Astrophys. J. Suppl. {\bf 148}, 175 (2003).
5687:
5688: \bibitem{Linde1982}
5689: A. D. Linde, Phys. Lett. B {\bf 108}, 389 (1982).
5690:
5691: \bibitem{Guth1982}
5692: A. H. Guth and S. Y. Pi, Phys. Rev. Lett. {\bf 49}, 1110 (1982).
5693:
5694: \bibitem{Dymnikova}
5695: E. B. Gliner and I.G. Dymnikova, Sov. Astron. Lett. {\bf 1} (1975) 93.\\
5696: I. G. Dymnikova, Sov. Phys. JETP {\bf 63} (1986) 1111.
5697:
5698: \bibitem{Starobinsky1982}
5699: A. A. Starobinsky, Phys. Lett. B {\bf 117}, 175 (1982).
5700:
5701: \bibitem{WMAPPol}
5702: A. Kogut {\sl et al.}, Astrophys. J. Suppl. {\bf 148}, 161 (2003).
5703:
5704: \bibitem{Dai2002}
5705: Z. G. Dai {\sl et al.}, Astrophys. J. {\bf 580}, L7 (2002).
5706:
5707: \bibitem{IMF}
5708: J. Bagchi {\sl et al.}, New Astron. {\bf 7}, 249 (2002).
5709:
5710: \bibitem{Hagedorn1965}
5711: R. Hagedorn, Nuovo Cim. Suppl. {\bf 3}, 147 (1965).
5712:
5713: \bibitem{GrossWilczek1973}
5714: D. J. Gross and F. Wilczek, Phys. Rev. Lett. {\bf 30}, 1343 (1973).\\
5715: D. J. Gross and F. Wilczek, Phys. Rev. D {\bf 8}, 3633 (1973).
5716:
5717: \bibitem{Politzer1973}
5718: H. D. Politzer, Phys. Rev. Lett. {\bf 30}, 1346 (1973).
5719:
5720: \bibitem{Linde1980}
5721: A. D. Linde, Phys. Lett. B {\bf 96}, 289 (1980).
5722:
5723: \bibitem{BraatenPisarski1990}
5724: J. C. Taylor and S. M. H. Wong, Nucl. Phys. B {\bf 346}, 115 (1990).\\
5725: E. Braaten and R. D. Pisarski, Nucl. Phys. B {\bf 337}, 569.
5726:
5727: \bibitem{Bodeker1998}
5728: D. B\"odeker, Phys. Lett. B {\bf 426}, 351 (1998).
5729:
5730: \bibitem{Bodapps1998}
5731: P. Arnold, D. T. Son, and L. G. Yaffe, Phys. Rev. D {\bf 59}, 105020 (1999).\\
5732: J.-P. Blaizot and E. Iancu, Nucl. Phys. B {\bf 557}, 183 (1999).\\
5733: D. F. Litim and C. Manuel, Phys. Rev. Lett. {\bf 82}, 4981 (1999).
5734:
5735: \bibitem{Kajantie2002}
5736: K. Kajantie, M. Laine, K. Rummukainen, and Y. Schroder, Phys. Rev. D {\bf 67}, 105008 (2003).\\
5737: K. Kajantie, M. Laine, K. Rummukainen, and Y. Schroder, Phys. Rev. Lett. {\bf 86}, 10 (2001).
5738:
5739: \bibitem{Blaizot2003}
5740: J.-P. Blaizot, E. Iancu, and A. Rebhan, hep-ph/0303185.
5741:
5742: \bibitem{Hofmann2000t2003}
5743: R. Hofmann, Phys. Rev. D {\bf 62}, 065012 (2000).\\
5744: R. Hofmann, Phys. Rev. D {\bf 62}, 105021 (2000).\\
5745: R. Hofmann, Phys. Rev. D {\bf 65}, 125025 (2002).\\
5746: R. Hofmann, Phys. Rev. D {\bf 68}, 065015 (2003).\\
5747: R. Hofmann, hep-ph/0312046.
5748:
5749: \bibitem{GinzburgLandau1950}
5750: V. L. Ginzburg and L. D. Landau, JETP {\bf 20}, 1064 (1950).
5751:
5752: \bibitem{Abrikosov1957}
5753: A. A. Abrikosov, Sov. Phys. JETP {\bf 5}, 1174 (1957).
5754:
5755: \bibitem{LitimPawlowski1999}
5756: D. F. Litim and J. M. Pawlowski, publ. in Faro 1998, {\sl The exact renormalization group}, 168,
5757: hep-th/9901063.
5758:
5759: \bibitem{BPST}
5760: A. Belavin, A. Polyakov, A. Schwartz, and Yu. Tyupkin, Phys. Lett. {\bf 59}, 85 (1975).
5761:
5762: \bibitem{HarrigtonShepard1977}
5763: B. J. Harrington and H. K. Shepard, Phys. Rev. D {\bf 17}, 105007 (1978).
5764:
5765: \bibitem{PrasadSommerfield1975}
5766: M. K. Prasad and C. M. Sommerfield, Phys. Rev. Lett. {\bf 35 }, 760 (1975).\\
5767: E. B. Bogomolnyi, Sov. J. Nucl. Phys. {\bf 24}, 449 (1976).
5768:
5769: \bibitem{Nahm1984}
5770: W. Nahm, Phys. Lett. B {\bf 90}, 413 (1980).\\
5771: W. Nahm, Lect. Notes in Physics. 201, eds. G. Denaro, e.a. (1984) p. 189.
5772:
5773: \bibitem{KraanVanBaalNPB1998}
5774: T. C. Kraan and P. van Baal, Nucl. Phys. B {\bf 533}, 627 (1998).
5775:
5776: \bibitem{vanBaalKraalPLB1998}
5777: T. C. Kraan and P. van Baal, Phys. Lett. B {\bf 428}, 268 (1998).\\
5778: T. C. Kraan and P. van Baal, Phys. Lett. B {\bf 435}, 389 (1998).
5779:
5780: \bibitem{LeeLu1998}
5781: K.-M. Lee and C.-H. Lu, Phys. Rev. D {\bf 58}, 025011 (1998).
5782:
5783: \bibitem{Brower1998}
5784: R. C. Brower, D. Chen, J. Negele, K. Orginos, and C-I Tan, Nucl. Phys. Proc. Suppl. {\bf 73}, 557 (1999).
5785:
5786: \bibitem{GrossPisarskiYaffe1981}
5787: D. J. Gross, R. D. Pisarski, and L. G. Yaffe, Rev. Mod. Phys. {\bf 53}, 43 (1981).
5788:
5789: \bibitem{BrownCarlitzLee1977}
5790: L. Brown, R. Carlitz, and C. Lee, Phys. Rev. D {\bf 16}, 417 (1977).
5791:
5792: \bibitem{BrownCarlitzCreamerLee1978}
5793: L. Brown, R. Carlitz, D. Creamer, and C. Lee, Phys. Rev. D {\bf 17}, 1583 (1978).
5794:
5795: \bibitem{BrownCreamer1978}
5796: L. Brown and D. Creamer, Phys. Rev. D {\bf 18}, 3695 (1978).
5797:
5798: \bibitem{Hooft1976}
5799: G. 't Hooft, Phys. Rev. D {\bf 14}, 3432 (1976), Erratum-ibid. D {\bf 18}, 2199 (1978).
5800:
5801: \bibitem{Diakonov2004}
5802: D. Diakonov, N. Gromov, V. Petrov, and S. Slizovskiy, Phys. Rev. D {\bf 70}, 036003 (2004).
5803:
5804: \bibitem{SchaferShuryak1998}
5805: Th. Schafer and E. V. Shuryak, Rev. Mod. Phys. {\bf 70}, 323 (1998).
5806:
5807: \bibitem{Actor1983}
5808: A. Actor, Ann. Phys. (N. Y.) {\bf 148}, 32 (1983).
5809:
5810: \bibitem{Chakrabarti1987}
5811: A. Chakrabarti, Phys. Rev. D {\bf 35}, 696 (1987).
5812:
5813: \bibitem{Atiyah1978}
5814: G. 't Hooft, Phys. Rev. D {\bf 14}, 3432 (1976), D {\bf 18}, 2199 (E) (1976).\\
5815: G. 't Hooft, unpublished.\\
5816: M. Atiyah, W. Drinfeld, N. Hitchin, and Yu. Manin, Phys. Lett. {\bf A}, 65 (1978).
5817:
5818: \bibitem{HerbstHofmann2004}
5819: U. Herbst and R. Hofmann, hep-th/0411214.
5820:
5821: \bibitem{Hosotani1983}
5822: Y. Hosotani, Phys. Lett. B {\bf 126}, 309 (1983).
5823:
5824: \bibitem{HerbstHofmannRohrer2004}
5825: U. Herbst, R. Hofmann, and J. Rohrer, Acta Phys. Pol. B {\bf 36}, 881 (2005).
5826:
5827: \bibitem{SHG2006}
5828: M. Schwarz, R. Hofmann, and F. Giacosa, hep-th/0603078.
5829:
5830: \bibitem{HofmannSept2006}
5831: R. Hofmann, hep-th/0609033.
5832:
5833: \bibitem{Gorenstein1995}
5834: M. I. Gorenstein and S. N. Yang, Phys. Rev. D {\bf 52}, 5206 (1995).
5835:
5836: \bibitem{RohrerHofmann2005}
5837: J. Rohrer and R. Hofmann, work in progress.
5838:
5839: \bibitem{KorthalsAltes}
5840: C. P. Korthals Altes, hep-ph/0406138, hep-ph/0408301.\\
5841: C. P. Korthals Altes, Acta Phys. Polon. B {\bf 34}, 5825 (2003).\\
5842: P. Giovannangeli and C. P. Korthals Altes, Nucl. Phys. B {\bf 608}, 203 (2001).\\
5843: C. Korthals-Altes, A. Kovner, and M. A. Stephanov, Phys. Lett. B {\bf 469}, 205 (1999).
5844:
5845: \bibitem{HoelbingRebbiRubakov2001}
5846: Ch. Hoelbing, C. Rebbi, and V. A. Rubakov, Phys. Rev. D {\bf 63}, 034506 (2001).
5847:
5848: \bibitem{'tHooft1974}
5849: G. 't Hooft, Nucl. Phys. B {\bf 79}, 276 (1974).
5850:
5851: \bibitem{Polayakov1974}
5852: A. M. Polyakov, JETP Lett. {\bf 20}, 194 (1974).
5853:
5854: \bibitem{PrasadSommerfield1974}
5855: M. K. Prasad and C. M. Sommerfield, Phys. Rev. Lett. {\bf 35}, 760 (1975).
5856:
5857: \bibitem{JuliaZee1975}
5858: B. Julia and A. Zee, Phys. Rev. D {\bf 11}, 2227 (1975).
5859:
5860: \bibitem{Aguilar1999}
5861: M. A. Aguilar and M. Socolovsky, Int. J. Theor. Phys. {\bf 38}, 2485 (1999).
5862:
5863: \bibitem{Steenrod1951}
5864: N. Steenrod, {\sl The topology of Fibre Bundles},
5865: Princeton University Press, Princeton, New Jersey.
5866:
5867: \bibitem{'tHooft1976}
5868: G. 't Hooft, Phys. Rev. D {\bf 14}, 3432 (1976).
5869:
5870: \bibitem{LandsmanWeert1987}
5871: N. P Landsman and C. G. van Weert, Phys. Rep. {\bf 145}, 141 (1987)
5872:
5873: \bibitem{ForgasVolkov2003}
5874: P. Forgacs and M. Volkov, Phys. Rev. Lett. {\bf 92}, 151802.
5875:
5876: \bibitem{Kaluza1921}
5877: T. Kaluza, Sitzungsber. Preuss. Akad. Wiss. Berlin (Math.Phys.) {\bf 1921}, 966 (1921).
5878:
5879: \bibitem{Klein1926}
5880: O. Klein, Z. Phys. {\bf 37}, 895 (1926).\\
5881: O. Klein, Surveys High Energ.Phys. {\bf 5}, 241 (1986).
5882:
5883: \bibitem{Olejnik1997}
5884: L. Del Debbio {\sl et al.}, proc.
5885: NATO Adv. Res. Workshop on Theor.
5886: Phys., Zakopane (1997) [hep-lat/9708023].
5887:
5888: \bibitem{NielsenOlesen1973}
5889: H. B. Nielsen and P. Olesen, Nucl. Phys. B {\bf 61}, 45 (1973).
5890:
5891: \bibitem{'tHooft1978}
5892: G. 't Hooft, Nucl. Phys. B {\bf 138}, 1 (1978).
5893:
5894: \bibitem{SvetitskyYaffe1982-1}
5895: B. Svetitsky and L. G. Yaffe, Nucl. Phys. B {\bf 210}, 423 (1982).
5896:
5897: \bibitem{SvetitskyYaffe1982-2}
5898: B. Svetitsky and L. G. Yaffe, Phys. Rev. D {\bf 26}, 963 (1982).
5899:
5900: \bibitem{Reinhardt2001}
5901: H. Reinhardt, Nucl. Phys. B {\bf 628}, 133 (2002).
5902:
5903: \bibitem{HofmannSchefflerStamatescu2005}
5904: R. Hofmann, S. Scheffler, and I.-O. Stamatescu, work in progress.
5905:
5906: \bibitem{BenderWu1969}
5907: C. M. Bender and T. T. Wu, Phys. Rev. {\bf 184}, 1231 (1969).
5908:
5909: \bibitem{nonequ}
5910: J. H. Traschen and R. H. Brandenberger, Phys. Rev. D {\bf 42}, 2491 (1990).\\
5911: L. Kofman, A. D. Linde and A. A. Starobinsky, Phys. Rev. Lett. {\bf 73}, 3195 (1994).\\
5912: J. Berges and J. Serreau, Phys. Rev. Lett. {\bf 91}, 111601 (2003).
5913:
5914: \bibitem{EnKaSaMo1982}
5915: J. Engels {\sl et} al., Nucl. Phys. B {\bf 205}[FS5], 545 (1982).
5916:
5917: \bibitem{Bielfeld1996}
5918: G. Boyd {\sl et} al., Phys. Rev. Lett. {\bf 75}, 4169 (1995).\\
5919: G. Boyd {\sl et} al., Nucl. Phys. {\bf B469}, 419 (1996).
5920:
5921: \bibitem{EngelsKarschScheideler1999}
5922: J. Engels, F. Karsch, T. Scheideler, Nucl. Phys. B {\bf 564}, 302 (1999).
5923:
5924: \bibitem{Brown1988}
5925: F. R. Brown {\sl et al.}, Phys. Rev. Lett. {\bf 61}, 2058 (1988).
5926:
5927: \bibitem{Deng1988}
5928: Y. Deng, in BATAVIA 1988, proc. LATTICE 88, 334.
5929:
5930: \bibitem{EngelsFingberg1990}
5931: J. Engels {\sl et al.}, Phys. Lett. B {\bf 252}, 625 (1990).
5932:
5933: \bibitem{GH2005}
5934: F. Giacosa and R. Hofmann, hep-th/0512184.
5935:
5936: \bibitem{LuciniTeperWenger2003}
5937: B. Lucini, M. Teper, and U. Wenger, JHEP {\bf 0401}, 061 (2004).
5938:
5939: \bibitem{LuciniTeperWenger2005}
5940: B. Lucini, M. Teper, and U. Wenger, JHEP {\bf 0502}, 033 (2005).
5941:
5942: \bibitem{Blum2004}
5943: Th. Blum, private conversation.
5944:
5945: \bibitem{Klassen1998}
5946: T. R. Klassen, Nucl. Phys. B {\bf 533}, 557 (1998).
5947:
5948: \bibitem{Sen2000}
5949: A. Sen and B. Zwiebach, JHEP03 {\bf 002} (2000).
5950:
5951: \bibitem{Prokopec2005}
5952: T. Prokopec, astro-ph/0503289.
5953:
5954: \bibitem{AtiyahSinger1984}
5955: M. F. Atiyah and I. M. Singer, Proc. Nat. Acad. Sci. {\bf 81}, 2597 (1984).
5956:
5957: \bibitem{BellJackiw1969}
5958: J. S. Bell and R. Jackiw, Nuovo Cim. A {\bf 60}, 47 (1969).
5959:
5960: \bibitem{Fujikawa1979}
5961: K. Fujikawa, Phys. Rev. Lett. {\bf 42}, 1195 (1979).
5962:
5963: \bibitem{PecceiQuinn1977}
5964: R. D. Peccei and H. R. Quinn, Phys. Rev. D {\bf 16}, 1791 (1977).\\
5965: R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. {\bf 38}, 1440 (1977).
5966:
5967: \bibitem{Sakharov1967}
5968: A. D. Sakharov, Pisma Zh. Eksp. Teor. Fiz. {\bf 5}, 32 (1967); JETP Lett. {\bf 5}, 24 (1967);
5969: Sov. Phys. Usp. {\bf 34}, 392 (1991); Usp. Fiz. Nauk {\bf 161}, 61 (1991).
5970:
5971: \bibitem{Wilczek2004}
5972: F. Wilczek, hep-ph/0408167.
5973:
5974: \bibitem{Williams1971}
5975: E. R. Williams, J. E. Faller, and H. A. Hill, Phys. Rev. Lett. {\bf 26}, 721 (1971).
5976:
5977: \bibitem{Dvali2003}
5978: E. Adelberger, G. Dvali, and A. Gruzinov, hep-ph/0306245.
5979:
5980: \bibitem{WMAP2003}
5981: D. N. Spergel {\sl et al.}, Astrophys. J. Suppl. {\bf 148}, 175 (2003).
5982:
5983: \bibitem{Alvensleben1968}
5984: H. Alvensleben {\sl et al.}, Phys. Rev. Lett. {\bf 21}, 1501 (1968).
5985:
5986: \bibitem{Blumenthal1966}
5987: R. B. Blumenthal {\sl et al.}, Phys. Rev. {144}, 1199 (1966).
5988:
5989: \bibitem{Ashkin1953}
5990: A. Ashkin, L. A. Page, and W. M. Woodard, Phys. Rev. {\bf 94}, 357 (1953).
5991:
5992: \bibitem{Scott1951}
5993: M. B. Scott, A. O. Hanson, and E. M. Lyman, Phys. Rev. {\bf 84}, 638 (1951).
5994:
5995: \bibitem{Krauss2004}
5996: Ch. Kraus {\sl et al.}, hep-ex/0412056.
5997:
5998: \bibitem{Klapdor2004}
5999: H. V. Klapdor-Kleingrothaus, I. V. Krivosheina, A. Dietz, and
6000: O. Chkvorets, Phys. Lett. B {\bf 586}, 198 (2004).
6001:
6002: \bibitem{Manuel2004}
6003: O. Manuel, in proc. Fourth Intern. Conf. on Beyond Standard Model Physics 2003, astro-ph/0411658.
6004:
6005: \bibitem{Laughlin1998}
6006: R. B. Laughlin, Nobel lecture publ. in NOBEL LECTURES IN PHYSICS (1996-2000),
6007: World Scientific Publishing Co.
6008:
6009: \bibitem{Shuryak2004}
6010: E. Shuryak, J. Phys. G {\bf 30}, S1221 (2004).
6011:
6012: \baselineskip25pt
6013:
6014: \end{thebibliography}
6015:
6016:
6017: \end{document}
6018:
6019: