1: %%%%%%%%%%%%%%%%%%%%%%%BEGIN MANUSCRIPT%%%%%%%%%%%%%%%%%%%%
2: \tolerance=10000
3: \documentclass[11pt]{article}
4: %\def\baselinestretch{1.2}
5: %\addtolength{\headheight}{-.5in}
6: \addtolength{\textwidth}{.5in}
7: \addtolength{\oddsidemargin}{-.25in}
8: %\addtolength{\footheight}{-.5in}
9: \addtolength{\textheight}{.5in}
10: \hyphenation{author another created paper re-commend-ed}
11: \usepackage{graphicx}
12: \begin{document}
13: %\preprint{SUNY-FRE-04-02}
14: %\renewcommand{\baselinestretch}{1.2}
15: \title{Critical or Tricritical Point in
16: Mixed-Action SU(2) Lattice Gauge Theory?}
17: \author{Michael Grady\\
18: Department of Physics\\ State University of New York at Fredonia\\
19: Fredonia NY 14063 USA}
20: \date{\today}
21: \maketitle
22: \thispagestyle{empty}
23: %\renewcommand{\baselinestretch}{1.2}
24: \begin{abstract}
25:
26: An analysis of scaling along the first-order bulk transition line
27: in fundamental-adjoint SU(2) lattice gauge theory
28: strongly supports the first-order endpoint being a tricritical point,
29: and is inconsistent with it being an ordinary critical point
30: as is usually assumed.
31: If tricritical, the transition must continue from the endpoint
32: further into the phase diagram as a second-order bulk transition and extend
33: to and beyond the Wilson axis. Observations
34: indicate that this is most likely the same transition that
35: has been traditionally considered a finite-temperature transition.
36: %\\ PACS:11.15.Ha, 11.30.Qc.
37: \end{abstract}
38: \section{Introduction}
39: The characterization of phase transitions has often been made clearer by
40: considering higher-dimensional coupling spaces, especially ones that
41: become more-familiar or exactly-known theories
42: at one or more edges of the phase diagram. Then
43: one can see how the various
44: phase boundaries and critical points attach to better-known transitions.
45: In SU(2) and SU(3) lattice
46: gauge theory, the fundamental-adjoint plane has provided interesting
47: insights. The SU(2) case was first studied by Bhanot and Creutz \cite{bc},
48: who found two lines
49: of first-order transitions which joined at a triple point
50: and then continued as
51: a single first-order line until ending at a
52: presumed critical point (Fig.~\ref{fig1}). They argued that
53: since the transition apparently ended,
54: the strong coupling confining phase
55: could be continued around the endpoint resulting in a
56: confining continuum limit,
57: since a connecting path which encounters no phase
58: transition could be found.
59: The fact that the Polyakov loop, an order parameter for deconfinement
60: (or disorder parameter for confinement), appeared to undergo a sudden
61: jump to non-zero values across this line would seem to be inconsistent
62: with the idea that both sides of the transition were confining,
63: however it was assumed that
64: in the limit of an infinite lattice the
65: deconfinement signal would disappear.
66: One place where it cannot disappear,
67: however is along the top line of the phase diagram
68: where $\beta _A = \infty$.
69: This is the well-know Z2 lattice gauge theory, which has a
70: bulk first-order transition at
71: $\beta _F=\frac{1}{2} \ln (1+\sqrt{2}) \approx 0.44$
72: (determined exactly from self-duality)\cite{wegner}.
73: This transition is deconfining with the Polyakov loop
74: as order parameter. Therefore the line $\overline{\rm{AB}}$ on
75: Fig.~\ref{fig1} is definitely deconfined even on the 4-d infinite lattice.
76: In the conventional interpretation of lattice gauge theory
77: the entire rest of the phase diagram
78: is confining on such a lattice.
79:
80: The situation became clearer when it was realized that finite
81: lattices were at a finite
82: 3-d (ordinary physical) temperature which increased
83: as $\beta$ increased. A deconfinement
84: transition due to physical temperature - a so called
85: finite-temperature transition - should
86: occur on a finite lattice. For SU(2) this is a
87: second-order transition and was studied
88: extensively on the Wilson axis ($\beta _A =0$).
89: In Ref.~\cite{ggm}
90: the finite-temperature transition
91: was studied on the fundamental-adjoint plane on lattices
92: with temporal extent $N_\tau = 4$.
93: Some couplings were further studied at $N_\tau = 6$ and 8 \cite{rajivt6}.
94: The rather surprising result of these studies was
95: that the line of second-order finite-temperature transitions seemed to
96: join up with the first-order bulk transition
97: at its endpoint.
98: The finite-temperature deconfinement
99: transition also became first-order at this point
100: (point D in Fig.~\ref{fig1}). If these
101: transitions truly were joined that would call
102: into question the finite-temperature interpretation of
103: the second-order transition. This is because a finite-temperature
104: transition should move all
105: the way to the right of the phase diagram as $N_\tau$ is
106: increased (similar to the hypothetical
107: dashed lines 1-4
108: on Fig.~\ref{fig1}). However this would not be possible if
109: one end was tied down at point D. In this case it
110: would be unlikely for the second-order transition to move
111: beyond line 1 (this is a line of constant
112: physics as determined by continuum two-loop perturbative
113: renormalization-group theory, as are lines
114: 2-4). Line 1 is constructed to join at the endpoint
115: of the first-order line,
116: as determined by extrapolating the latent heat (which appears to
117: vary linearly with $\beta _A$)
118: to zero using the $12^4$ lattice
119: data from the current study.
120: This is at ($\beta _F$, $\beta _A$)=(1.38$\pm 0.03$, 1.04$\pm 0.01$),
121: a somewhat
122: higher $\beta _A$ (and lower $\beta _F$) than the original
123: low-statistics Bhanot-Creutz result. This is in agreement
124: with the results of Refs. \cite{ggm,rajivt6},
125: although not with \cite{rajivsl} (more on this later). In the
126: current paper,
127: this point will be referred to as the ``first-order-endpoint" (FOE), as its
128: identification as either a critical point or a
129: tricritical point is the question
130: under consideration.
131: To the right of the phase line the coupling quickly becomes weak enough
132: for these perturbative lines of constant physics to be accurate.
133: Near the phase transition they are more hypothetical.
134: Putting aside the unreasonable possibility of a phase line
135: that curves up and then down, one can therefore
136: conclude that if the second-order
137: deconfinement transition continues to join
138: the bulk transition at point D as
139: $N_\tau \rightarrow \infty$ then,
140: at least for $\beta _A >1$, the zero-temperature continuum limit
141: (right hand side of the phase diagram) would be deconfined.
142: In other words, the entire transition would
143: be bulk. The movement of the second-order line
144: with changing lattice size, seen on the Wilson axis,
145: would be explained as an ordinary finite-size shift in the
146: critical point, perhaps with an unusually large shift-exponent or
147: following something other than a power law.
148: In this case the transition point would converge to some finite
149: value of $\beta _F$, probably in the range 3.0 to 4.0,
150: as $N_\tau \rightarrow \infty$.
151:
152: Another hypothesis consistent with the
153: original Bhanot-Creutz scenario is that the two transitions
154: join, but {\em not} at the FOE.
155: There are a few systems known in which a
156: second-order line meets a first-order line
157: at a point other than its endpoint, such as the
158: Blume-Emery-Griffiths model\cite{beg}
159: and certain metamagnets\cite[pp175-181]{CL}.
160: These systems each
161: have two different order parameters
162: and two corresponding correlation lengths. However, in these systems both
163: transitions are bulk.
164:
165: Applying this scenario to the lattice gauge theory,
166: where the second-order transition
167: is finite-temperature,
168: the point where the second-order transition joins the
169: first-order line is hypothesized to slowly move up the diagram as $N_\tau$
170: increases (lines 2-4 in Fig.~\ref{fig1}). The
171: lower part of the bulk transition would no longer be deconfining.
172: Eventually as $N_\tau$ became infinite
173: the entire phase diagram except for the line AB would be confined.
174: The bulk transition would have
175: nothing to do with confinement. The observed deconfinement across
176: the bulk line on small lattices would be due to
177: the coincident finite-temperature transition,
178: which somehow becomes first-order
179: due to the influence of the bulk transition.
180: The bulk transition would have its own order
181: parameter which was not symmetry-breaking,
182: similar to the liquid-gas transition.
183: This order parameter would have a correlation length
184: associated with it which would become infinite at the critical point
185: at the end of the first-order line.
186: This would happen at a place within the confining region,
187: where the string tension
188: (and correlation length associated with it) is finite. In other
189: words the theory would have to have two independent correlation lengths.
190: This seems a bit odd in that there
191: is no evidence for the more-complicated
192: scaling laws that would normally result from a theory
193: with more than one correlation length.
194: However, otherwise this interpretation is consistent with the
195: conventional interpretation of Lattice Gauge Theory -
196: in particular with a confining
197: zero-temperature continuum limit.
198:
199: It is difficult to distinguish these two
200: hypotheses simply by looking for a joining away from the
201: FOE, because one would have to go to rather high $N_\tau$ to get a
202: convincing separation. A small
203: separation was reported for the SU(3) case \cite{su3} which
204: has a similar phase diagram, except that both transitions
205: (bulk and deconfining) are first-order.
206: However, critical points determined on finite lattices from different
207: quantities or by different techniques
208: can differ substantially from one another.
209: This could lead to a small apparent separation of
210: the critical point and the end of the finite-temperature line,
211: since one is determined from the plaquette and
212: the other from the Polyakov loop. A more convincing demonstration
213: of separation would be the observation of uncorrelated tunneling events
214: in two different order parameters.
215:
216: There is actually a much easier way to determine which
217: of these cases is correct, based on a study of the
218: bulk transition itself. In the conventional scenario just described,
219: the line of first-order transitions
220: ends in an ordinary critical point.
221: Its order parameter therefore cannot be associated with spontaneous
222: symmetry breaking, because otherwise the
223: transition would have to continue, in order
224: to divide the plane into symmetry-broken and unbroken parts.
225: First-order transitions ending in a critical point
226: are characterized in the Landau theory as having a
227: cubic term in the free energy which explicitly breaks
228: the symmetry. Only at the critical point, where the
229: cubic term is zero, is the symmetry accidentally realized,
230: allowing for a single point of criticality.
231: The other scenario, where the second-order line joins the
232: first-order at the endpoint, is exactly what happens at
233: a {\em tricritical} point. A tricritical point is
234: associated with a symmetry-breaking phase transition.
235: Its Landau free-energy has only even-order terms
236: but must be considered out to sixth order,
237: because in part of the phase diagram the quartic term is negative.
238: Here there is a first-order transition,
239: which becomes second-order when the quartic term becomes positive.
240: The tricritical point is simply where the change in order takes place,
241: when the quartic term vanishes.
242: If the endpoint of the fundamental-adjoint
243: bulk transition is tricritical, then the transition {\em must}
244: continue as second-order,
245: and it must be symmetry breaking.
246:
247: A favorable aspect of this study is that the
248: bulk transition should not, by its very nature, depend much
249: on lattice size. Bhanot and Creutz's study was done on a
250: $5^4$ lattice, and more recent results,
251: such as those presented here
252: for $12^4$ and $20^4$ lattices, do not differ much
253: in the location of the transition
254: or other parameters such as latent heat.
255: Except for an expected reduction in variance from simple statistics, no
256: significant differences are seen between our runs
257: on $12^4$ and $20^4$ lattices(see Fig.~\ref{fig3} below).
258: Similarly, because it is a bulk quantity,
259: one would expect the Landau free energy function
260: to be accurately determined by modest lattices with
261: very minor corrections from surface effects.
262: Any results linked to the behavior of this free energy
263: are therefore unlikely to change much on larger lattices.
264:
265: The lack of finite size dependence seen in the data shown
266: below contrast with what was reported by Gavai who also studied the
267: bulk transition on symmetric lattices\cite{rajivsl}. This
268: can be traced to a difference in measurement
269: technique. Gavai found a significant decrease in latent heat as lattice
270: size was increased at $\beta _A = 1.25$.
271: The latent heat decreased by a factor of
272: three as the lattice size was varied from $6^4$ to $16^4$.
273: This information was extracted
274: from runs very close to the phase transition for each lattice,
275: from which the
276: latent heat was taken from the peak separation of the apparently
277: bimodal distribution.
278: However, this method has a particular problem when used on the symmetric
279: lattice due to the symmetry actually being (Z2)$^4$
280: rather than just Z2 (assuming periodic boundary
281: conditions in all four directions).
282: As the phase transition is approached on the finite lattice, the lattice
283: goes from having four broken directions (and no unbroken),
284: first to three broken
285: (and one unbroken), then to two broken, then one, and finally to the fully
286: unbroken symmetry case. The reason for this is simply the entropy factor
287: associated with each. There are 16 states with four broken directions,
288: $4\cdot 8=32 $ with three broken directions, 24 with two, and 8 with one and
289: only one corresponding state in the fully symmetric state. At the critical point
290: where these all have equal energy, the fully unbroken state would occur
291: only 1/81 of the time. It is easy to see how this state could be missed. If
292: one set $\beta$ so that this state occurred 50\% of the time, one would be below
293: the critical point (in $\beta$). The Boltzmann factor would then suppress the multiple-broken
294: direction cases, which might not appear at all on larger lattices where the energy
295: fluctuations are smaller. Thus, from a practical point of view, it is very difficult to
296: display the full range of symmetry-broken cases in a single simulation.
297: The multiple symmetry breakings appear
298: to be associated with
299: nearly equal jumps in the plaquette (see Fig.~\ref{fig2}).
300: On smaller lattices close to the transition,
301: tunneling will occur between all of these states, showing nearly the full
302: latent heat. However, larger lattices, with their smaller
303: energy fluctuations
304: (requiring hitting the critical point more accurately)
305: and longer tunneling times, may
306: only tunnel between two or three of the five levels in a reasonable-length
307: run, showing an apparently smaller
308: latent heat.
309: Fig.~\ref{fig2}a shows a time history on an $8^4$
310: lattice at $\beta _A$=1.25
311: and $\beta _F$=1.2185 along with the Polyakov loop histories.
312: There appear to be
313: several energy plateaus in between the upper and lower,
314: associated with only some of the Polyakov loop
315: directions breaking.
316: A similar simulation an a $12^4$ lattice with $\beta _F = 1.2183$
317: (closer to the
318: critical point) populates only four of the five levels.
319: Its plaquette histogram, shown in
320: Fig.~\ref{fig2}b, shows four peaks. The missing peak in this case is
321: from all-four
322: Polyakov loops unbroken.
323: A similar $16^4$ simulation at the same couplings populated
324: only three peaks.
325: The multimodal nature of the distribution is not always as apparent
326: as in Fig.~\ref{fig2}b.
327: If the peaks are unequally populated then only shoulders will be seen.
328: Thus it is risky to try to measure the latent heat from the widths of
329: these distributions unless all five peaks of the multimodal distribution
330: can be resolved. This multiple symmetry-breaking
331: effect could easily explain the decrease in latent heat
332: with lattice size seen in Ref.~\cite{rajivsl}.
333:
334: The data presented in the current study are from
335: hysteresis loops, which do not suffer from this
336: problem, as not much time is spent
337: at $\beta _c$. As shown below, the hysteresis
338: loops on $12^4$ and $20^4$ lattices are
339: nearly identical, suggesting almost no finite size effects.
340: Our values of latent
341: heat agree closely with the values Gavai and coworkers found on
342: {\em asymmetric} lattices,
343: for which no finite size effect was found. Asymmetric
344: lattices with one short
345: direction do not suffer
346: from the above problem, as the symmetry broken
347: at the transition in question
348: is then just
349: the single Z2 of the short direction.
350:
351:
352:
353:
354: \section{Critical vs. Tricritical}
355: One needs to find an easily-measured quantity which can distinguish
356: the critical from the tricritical cases.
357: As shown below, it turns out that the size of the hysteresis region,
358: $(T^{**} - T^*)$ is
359: a linear function of the latent heat for the critical case
360: and a quadratic function for the tricritical case.
361: Here $T^*$ is the lower (supercooling) metastability limit
362: and $T^{**}$ is the upper (superheating).
363: Between these temperatures, there are two minima
364: of the free energy and tunneling exists.
365: Outside of this region there is only one local minimum and
366: no tunneling exists.
367: This gross difference in scaling behavior follows from basic dimensional
368: analysis of the powers in the Landau free energy,
369: but a detailed derivation is also given below.
370: Both the latent heat and metastability regions are easily
371: determined from hysteresis sweeps.
372: By plotting vs. latent heat no assumptions need be made concerning the
373: possibly complex relationship between
374: ($\beta _F$, $\beta _A$) and the temperature and next higher coefficient
375: in the free energy. It is interesting that this method is able to distinguish
376: between symmetry-breaking and non-symmetry-breaking first-order transitions
377: from energetics alone, without the need to identify the symmetry or order parameter.
378:
379: In the Landau theory, the free energy is given
380: by a power series in the order parameter.
381: \begin{equation}
382: f=\frac{1}{2} r \phi ^2 - w \phi ^3 + u_4 \phi ^4 + u_6 \phi ^6 .
383: \end{equation}
384: The quantity $r$ is an increasing function of temperature,
385: which can be defined as $r=a(T-T^* )$.
386: For an ordinary 1st order transition that ends in a critical point,
387: $w>0$ and $u_4>0$ ($w$ becomes zero at the
388: critical point, whereas $u_4$ remains positive everywhere).
389: The sixth order term can be ignored.
390: The critical point occurs when $f=0$, $\partial f/ \partial \phi = 0$ for
391: the minimum away from $\phi = 0$.
392: These are easily solved for $\phi _c = w/2u_4$ and $r_c = w^2/2u_4$.
393: The latent heat can be obtained from the
394: change in entropy between phases. Taking $s=-df/dT$ and expanding
395: $f$ to lowest order in $r$ about the two minima gives\cite{CL}
396: $\Delta s=(a/2)\phi_{c}^2$. Therefore the latent heat is
397: given by
398: \begin{equation}
399: q=T\Delta s = (aT_c /2)(w/2u_4)^2 .
400: \end{equation}
401: Further details can be found in Ref.~\cite[pp.\ 168-175]{CL} from which
402: this and the following derivations are abstracted.
403: Note that $q$ is quadratic in $w$, the parameter
404: which is rapidly varying as one moves along the critical line,
405: away from the critical point (rapidly varying
406: because it must vanish at the critical point).
407: The metastability limit on superheating, $T^{**}$ occurs when the
408: local minimum away from $\phi =0$
409: becomes an inflection point instead,
410: i.e. $\partial f/\partial \phi =0$ and
411: $\partial ^2 f /\partial \phi ^2 =0$.
412: Solving these gives
413: \begin{equation}
414: r^{**}=9w^2 /16u_4 ,
415: \end{equation}
416: also quadratic
417: in $w$. Taking the usual assumption that $u_4$ is slowly varying,
418: results in the prediction that
419: \begin{equation}
420: T^{**}-T^*=a^{-1}r^{**} \propto q.
421: \end{equation}
422:
423: In contrast, for the first-order transition that ends
424: in a tricritical point,
425: $w=0$ is enforced by symmetry. The quantity $u_4<0$ and
426: one needs the positive $u_6$ term for stability. In this case {\em two}
427: additional minima away from $\phi =0$ occur, one for
428: positive and one for negative $\phi$.
429: If these dip below the minimum at $\phi =0$ a first-order phase
430: transition occurs. The tricritical point occurs when $u_4=0$.
431: Beyond this is a line of second-order
432: transitions ($u_4 >0$).
433: In this case the transition changes from first-order
434: to second-order, rather
435: than disappearing. In fact it must continue on, in order
436: to divide the entire
437: coupling plane into symmetry-broken and symmetry unbroken phases.
438: Here $u_4$ is the rapidly
439: changing parameter as one moves along the
440: transition line near the tricritical point, and
441: $u_6$ is assumed to be slowly varying.
442: Following the same procedure given above
443: results in \cite{CL}
444: \begin{equation}
445: \phi _c = \pm [ |u_4 |/2u_6 ]^{\frac{1}{2}}
446: \end{equation}
447: \begin{equation}
448: q=aT_c|u_4 |/(4u_6)
449: \end{equation}
450: (linear in $u_4$),
451: and
452: \begin{equation}
453: r^{**} =2u_4^2 /(3u_6)
454: \end{equation}
455: (quadratic in $u_4$).
456: Therefore the prediction for the
457: tricritical case is
458: \begin{equation}
459: T^{**}-T^* \propto q^2 .
460: \end{equation}
461:
462: \section{Hysteresis Loops}
463:
464: Rather slow hysteresis sweeps were performed to
465: determine $q$ and $\Delta \beta \equiv \beta _F^{**} - \beta _F^*$,
466: for various fixed $\beta _A$.
467: One has a fair degree of flexibility in deciding which
468: parameter to choose as the
469: temperature. Here $\beta _F$ is being treated as the
470: inverse (4-d) temperature in the partition function.
471: The $\beta _A$ term with $\beta_ A$ held fixed can be thought
472: of as either a temperature dependent external field
473: term (with coefficient $\beta_A / \beta _F$) or a temperature independent
474: modification to the measure (contributing to the entropy).
475: Because $\Delta \beta \ll \beta _F$, $\Delta \beta \propto T^{**} - T^*$
476: to lowest order.
477: Each run was begun with 500 equilibration sweeps,
478: followed by 2000-4000 sweeps where $\beta _F$ is changed by 0.0001
479: on each sweep. This is slow enough that no
480: hysteresis can be detected away from critical regions.
481: Measurements were performed after each sweep. Runs were
482: performed on a $12^4$ lattice,
483: except for five additional runs performed on a $20^4$
484: lattice to test for finite
485: size effects.
486: Typical results for three different $\beta _A$ are
487: shown in Fig.~\ref{fig3}. Each $12^4$ sweep was
488: performed three times to test for repeatability and
489: to estimate errors (only one is shown).
490: The latent heat was measured as the
491: jump in fundamental plaquette, $<\! p\!>$, at the
492: position of maximum vertical
493: distance between
494: the hysteresis curves (with the definition of temperature above,
495: the internal energy is given by $1-<\! p\!>$).
496: The quantity $\Delta \beta$ was measured as the maximum width of the
497: hysteresis curve.
498: One can also take $\beta _F ^*$ and $\beta _F ^{**}$ to be
499: the points of maximum slope of the
500: hysteresis curves, and compute
501: $\Delta \beta$ from the difference - the results are nearly identical.
502: Multiple runs are remarkably similar,
503: indicating modest statistical errors (detailed later).
504: A worrisome systematic error from hysteresis sweeps is
505: the possibility of premature tunneling.
506: If one sweeps too slowly, the system could tunnel to the other phase
507: before the metastability limit is reached.
508: In rare instances it could happen a considerable distance away.
509: However, the fact that the $20^4$ sweeps were nearly identical
510: to the $12^4$ at the same $\beta$'s
511: (see Figs.~\ref{fig3} and \ref{fig4}), would seem to
512: indicate this is not a problem.
513: Tunneling times on the larger lattice are much longer,
514: so if the smaller lattice
515: were tunneling prematurely by a significant amount
516: then some difference between these different-size lattice runs
517: would be expected. The main result is shown in Fig.~\ref{fig4},
518: where $\Delta \beta$ is plotted against the square
519: of the jump in average plaquette, $(\Delta \! <\! p\!>)^2$
520: (proportional to $q^2$).
521: Although three independent measurements for each point
522: are not sufficient to accurately determine individual error bars,
523: an overall error estimate for the entire dataset
524: can be made, which indicates error bars of about one-third the
525: size of plotted points vertically and twice this horizontally.
526: If one does compute individual error bars, no particular trend is
527: observed - they are consistent with approximately equal error bars
528: for all couplings.
529: A linear trend (which on these axes is a pure quadratic)
530: is observed to fit the data well.
531: Thus the data agree with the
532: prediction of the tricritical case.
533: If one tries to fit to a linear function of $\Delta \! <\! p\!>$,
534: the result is
535: not satisfactory, with $\chi ^2/\rm{d.f.} =53$,
536: whereas the pure quadratic shown in Fig.~\ref{fig4},
537: plotted as a straight line with axes given,
538: has $\chi ^2 /\rm{d.f.}= 0.6$. A linear+quadratic fit was also performed.
539: In the possible case that the trend is linear, but the region of validity of
540: the Landau theory is small, this should be able to pick up the linear term.
541: This fit, however, gives a linear coefficient of $0.008 \pm 0.016$, consistent
542: with zero. The quadratic term (with coefficient $0.839\pm 0.039$) dominates
543: already at $(\Delta \! <\! p\!>)^2 = 0.0001$. Thus the data are not consistent with
544: ``beginning linear" over any reasonable region of validity.
545: Therefore, the data appear to be strongly inconsistent with the possibility
546: of there being an ordinary critical point at
547: the FOE but entirely consistent with it being a tricritical point.
548:
549: These results rely on certain assumptions
550: inherent to the Landau theory, namely that higher order terms
551: in the free energy are slowly varying.
552: However if this assumption were not valid it would be unlikely to obtain
553: such a clean result as pure quadratic scaling.
554: Much more likely in this case would be a less conclusive result requiring a
555: multiple-term fit. Also, mean field theory is
556: much more likely to give valid results in four dimensions than
557: in three, where it still provides useful predictions in many cases.
558:
559: \section{Identification of the order parameter}
560:
561: The tricritical behavior described above requires
562: a symmetry breaking order parameter. So far, in this analysis, it has
563: not been necessary to identify this symmetry or the associated
564: order parameter. This was deliberately done in
565: order to make as few assumptions as possible.
566: However, now that the tricritical case seems to be established
567: from energetics alone, it makes sense to try to identify
568: the associated broken symmetry. There are many reasons to
569: believe this is no other than the familiar
570: Z2 Polyakov-loop symmetry. For one thing, this is the symmetry that
571: is broken in the attached Z2 lattice gauge theory at the top of
572: the phase diagram. Secondly, the Polyakov loop
573: is seen to break along the bulk line,
574: not just at the same couplings,
575: but also tunneling at the same times. Fig.~\ref{fig5}
576: shows Polyakov-loop histories for heating and
577: cooling sweeps together with plaquette
578: histories for $\beta _A = 1.7$,
579: 1.25 and 1.0 on the $20^4$ lattice. These are well above,
580: moderately above, and
581: slightly below the tricritical point. In the first-order region,
582: the metastability in Polyakov loop matches exactly
583: with that of the plaquette. However even in the
584: second-order region the small hysteresis
585: signal in the plaquette from critical
586: slowing-down appears to be associated with the
587: symmetry breaking of the Polyakov loop.
588: This, together with the correlations seen in Fig.~\ref{fig2} would seem to
589: indicate not merely coincident phase transitions, but an intimate
590: locking of order parameters as well, since the tunnelings
591: in plaquette and Polyakov loops are
592: observed to take place at the same Monte Carlo times.
593:
594: The $20^4$ data employ moving averages in the Polyakov loops to reduce the
595: variance enough to see the symmetry breaking. The Polyakov loop
596: values in the broken region on such a large lattice are tiny. The
597: individual datapoints
598: are swamped by random fluctuations. Moving averages of, say 100 points,
599: reduce these random fluctuations by a factor of 10, allowing the small
600: nonzero average value in the broken phase to show through remarkably well.
601: Tunneling times in the broken region are generally much longer than this,
602: so the average values deduced are fairly accurate .
603:
604: One important objection to the Polyakov loop being an order parameter
605: for a bulk transition is that its very definition depends on there
606: being periodic boundary conditions. A bulk transition, on the other hand,
607: should exist for any boundary conditions, such as open boundary conditions
608: for which the Polyakov loop is not defined. However, it is important to
609: realize that this same objection can be made for the Z2 lattice gauge
610: theory, for which there is no controversy about the existence of a
611: bulk deconfining transition. Therefore, at least in this theory, there must be a
612: second hidden
613: order parameter and associated broken symmetry that exists for
614: the case of open boundary conditions. It is reasonable to expect that this
615: same situation may also exist in the SU(2) case. A possible symmetry
616: and order parameter which exist for both theories have been identified.
617: If one employs a partial axial gauge fixing that leaves unbroken a global
618: gauge symmetry on each 3-d layer of the lattice
619: perpendicular to a certain direction,
620: then these remaining global gauge symmetries appear to break spontaneously at
621: weak coupling, apparently
622: in concert with the Polyakov loop\cite{hidden}. The order
623: parameters are just the average perpendicular links in each layer.
624:
625:
626: The existence of a tricritical point implies that a
627: bulk (4d zero-temperature) second-order transition exists below it,
628: down to and even below the Wilson axis (because it must divide the
629: plane into two non-connected regions with different symmetry).
630: If the Polyakov loop is the order parameter then
631: there is no finite-temperature
632: deconfinement transition. Deconfinement is a
633: bulk transition and the zero-temperature continuum theory is not confining.
634: In order for the deconfinement transition to be a
635: finite-temperature one as is usually assumed, it
636: must decouple from the bulk transition as described in the introduction.
637: However, this has apparently not happened
638: yet, even on the $20^4$ lattice, in the coupling regions studied.
639: If a separation does occur, then a {\em new} symmetry and
640: order parameter must be found for the second-order bulk transition
641: emanating from the tricritical point.
642: It will be important to locate this
643: new phase transition on the Wilson axis.
644: In order for lattice gauge results to be
645: analytically connected to the continuum limit,
646: one may only run simulations on the
647: weak-coupling side of any bulk transition.
648: If it occurs near the deconfinement transition on accessible lattices,
649: as is likely the case, then there would only be
650: a narrow region of valid couplings in which reliable confining simulations
651: using the Wilson action could be run,
652: lying between the new second-order bulk transition and the
653: previously known finite-temperature transition.
654: However, the
655: fact that no second-order bulk transition separate
656: from the deconfinement transition has ever been seen would
657: seem to make the entire scenario of two separate transitions unlikely.
658: The order parameter studied in ref. \cite{hidden} breaks both
659: the global gauge symmetry and the Z2 Polyakov loop symmetry.
660: If this is driving the bulk transition, then it will only be
661: possible for the Polyakov loop deconfinement transition to
662: split off to the strong coupling side of this transition, because
663: on the weak coupling side the Z2 symmetry will already be broken by the bulk transition.
664: In this case there would be no region of validity for the confining theory (always
665: separated from the continuum limit by the bulk transition).
666:
667: The evidence for the Polyakov-loop
668: transition being a finite-temperature one stems
669: mostly from an observed shift
670: in transition point with temporal lattice size,
671: $N_\tau$, on asymmetric lattices with $N_ \sigma > N_ \tau$. The
672: size of the shift is larger than one usually expects
673: for a bulk transition. However the possibility exists that
674: the four-dimensional non-abelian gauge theory could just
675: have an unusually large finite-size shift. Fig.~\ref{fig6}
676: shows data for the finite-temperature
677: deconfinement transition point, $\beta _c$ for the Wilson
678: action on different size asymmetric lattices
679: (data from Ref.~\cite{fhk}). Also shown is data from a
680: different action used by Gavai\cite{rajivft}, in which Z2 monopoles
681: and vortices are suppressed. This action has the same
682: $\Lambda$-parameter and therefore the same perturbative
683: scaling as the Wilson action, but the scaling is much
684: different in the region of the deconfinement transition.
685: Indeed, although the Wilson-action data appear to fit roughly to
686: the weak-coupling renormalization group scaling law (though
687: not acceptably, with a $\chi ^2 /d.f. = 28$), it does not fit at all to the
688: Z2 monopole and vortex suppressed action data ($\chi ^2 /d.f. = 100$). However,
689: eliminating strong coupling lattice artifacts would be expected to
690: improve scaling. This suggests that the rough fit of the Wilson-action data
691: may be accidental. A linear fit to $1/\ln (N_{\tau})$ also produces
692: a rough fit in the Wilson-action case (but also unacceptable considering
693: the very small errors quoted for these points, with $\chi ^2 /d.f. = 64$).
694: A linear fit fares better with the Gavai data with $\chi ^2 /d.f. = 4.6$
695: (none of the fits discussed thus far include the lowest $N_{\tau}=4$ points).
696: The apparent intersection of these lines at $N_{\tau}=\infty$ is probably
697: fortuitous, but it is interesting that they would agree on the infinite lattice
698: critical point.
699: The scaling behavior of lattices with the same $\Lambda$-parameter
700: need only match in the weak coupling region. They do not need
701: to match exactly at $\beta _{c \infty}$ (the critical coupling
702: for an infinite lattice), but since this $\beta$ is close
703: to the perturbative region they should be close.
704: Above a phase transition there
705: is no reason for them to match at all, which could explain the
706: rather different slopes. The dashed line fits include a quadratic
707: as well as a linear term (here the $N_{\tau}=4$ point was included in the Gavai-data
708: fit). This is able to accommodate the small curvature in the data rather well, and
709: still has near-agreement for the infinite lattice critical point of around $\beta = 3.9$.
710: The $\chi ^2 /d.f.$ for these fits are 3.5 and 2.3 which are coming rather close
711: to acceptability, considering the quadratic term is probably just approximating a
712: more complex non-linearity. A possible reason for approximate but not exact
713: $1/\ln (N_{\tau})$
714: scaling is as follows.
715:
716: The behavior pictured in Fig.~\ref{fig6} can be
717: understood if the transition is associated with percolation of
718: abelian-monopole loops in the maximal abelian gauge. It has
719: been shown that confinement seems to be associated with
720: the existence of a monopole loop that wraps through
721: the periodic boundary, or with the closely-related existence
722: of a percolating cluster of such loops. The deconfinement
723: transition seems to occur when this is no longer the case.
724: The probability of a monopole loop of size $l$ (for $l<N$) existing on
725: a lattice, normalized per lattice site, has been shown to be
726: proportional to $l^{-\gamma}$ where $\gamma$ is a $\beta$-dependent
727: quantity\cite{teper,myq}. Here $N$ is the linear size
728: of a symmetric lattice.
729: $\gamma$ is about 3 in the crossover region and becomes
730: equal to 5 around $\beta = 2.9$\cite{myq}. The scaling law may
731: be somewhat different for $l>N$, but the results are insensitive to this
732: so long as the exponent $\gamma \ge 1$ for all $l>N$\cite{myq}.
733: To have a wrapping loop, one requires at least one loop of size
734: of order $N^{1+\epsilon}$ or larger,
735: where $\epsilon$ is a fractal dimension
736: between 0 and 1, which has not yet been accurately measured.
737: Probably $\epsilon >0$ because the loops are
738: generally somewhat crumpled but $\epsilon = 0$ is
739: also a possibility. The probability of finding such a loop
740: on an $N^4$ lattice is $CN^{(5+\epsilon )-\gamma (1+\epsilon)}$
741: where $C$ is
742: some constant. This expression results from integrating the probability
743: over loop sizes beginning at the critical loop size for a wrapping loop.
744: Setting this probability to $\frac{1}{2}$ results in an
745: estimate for the critical
746: value of $\gamma$ for
747: that lattice, from which $\beta _c$ can be determined.
748: This therefore gives a model for the dependence of $\beta_c$ on $N$.
749: If one assumes that the asymmetric lattices, for which most of the
750: data has been taken, have a similar scaling law, and also assuming
751: that $\gamma$ can be taken to be a linear function of $\beta$ in the
752: region of interest (a simplification), then one obtains the
753: finite-lattice scaling law
754: \begin{equation}
755: \beta_c = \beta _{c\infty} -\frac{c}{ln(N)}
756: \end{equation} where $c$ is a constant.
757: This contrasts with the usual finite-lattice shift for a thermal
758: transition
759: \begin{equation}
760: \beta_c =\beta_{c \infty} -\frac{c}{N^{1/\nu}}
761: \end{equation}
762: which converges much more rapidly to $\beta _{c \infty}$ as
763: $N \rightarrow \infty$.
764: The idea that this transition may be a 4-d percolation transition
765: could explain why it has been so hard to identify, as there are not many
766: examples of such transitions. The fact that $\gamma$ is a somewhat
767: non-linear function of $\beta$ \cite{myq} could explain the need for a
768: quadratic term in the fits. Better determination of $\gamma (\beta )$ and
769: measurements on asymmetric lattices would be needed
770: to come to a definitive conclusion on the higher-order terms.
771:
772:
773: It has been shown by Gavai and Mathur that the bulk
774: first-order transitions are lattice artifacts that can be removed through
775: a judicious choice of action,
776: leaving only a second-order deconfining transition\cite{gmla}.
777: Indeed all bulk transitions are
778: lattice artifacts in that they do not affect the continuum limit.
779: However, the question still
780: exists whether this remaining second-order transition is
781: bulk or finite-temperature. By using the original Bhanot-Creutz action,
782: one can learn more about this transition by its apparent
783: connection to the first-order bulk transition at a
784: tricritical point. The existence of a tricritical point,
785: surmised above from the behavior of the free energy around the
786: first-order transition,
787: strongly implies that the second-order line is also bulk (i.e. a 4-d
788: zero-temperature transition), due to its attachment to
789: the bulk first-order endpoint. If this is true in the
790: original action, then the continuum limit is deconfined for
791: this action, and weak-coupling universality would imply
792: this is also true for all other actions. Thus this work supports
793: the hypothesis made some time ago that the continuum limit
794: of SU(2) pure-glue lattice gauge theory may not
795: be confining\cite{zp,ps2}.
796:
797: One should, of course, remember that the SU(2)
798: non-abelian gauge theory is not Quantum Chromodynamics (QCD) but an
799: approximation to it. Lack of confinement in the SU(2) theory
800: does not imply that quarks are unconfined in the real world.
801: However, it does shed important light on
802: the possible confinement mechanism.
803: Of course this result must first
804: be checked in the SU(3) case. If it holds up there,
805: then suspicion must be cast on light quarks as the source
806: of confinement in QCD, a position held by Gribov\cite{gribov,abgribov} and
807: others\cite{zp,nc,cahill}. It could be that confinement is a byproduct of
808: chiral symmetry breaking rather than the other way around
809: as sometimes stated.
810: One possibility is that strong color fields disrupt the chiral condensate,
811: creating a bag of diminished and polarized chiral condensate
812: around a hadron,
813: carrying an energy proportional to the volume of excluded condensate.
814: Supporting this conjecture is the observation in lattice simulations
815: that the strength of the chiral condensate is reduced from its vacuum value
816: in the presence
817: of a quark source\cite{markum}.
818:
819: A way to conclusively demonstrate confinement due to light
820: quarks would be to find an
821: action that completely erases the bulk confinement transition,
822: including the second-order one, but agrees with the Wilson action
823: at weak coupling. Since all bulk transitions are lattice artifacts,
824: this should be
825: possible. An action that
826: suppresses both Z2 monopoles and vortices as well as
827: another gauge-invariant topological lattice artifact,
828: the SO(3)-Z2 monopole \cite{so3z2}
829: appears promising. Simulations on lattices up to $30^4$
830: remain deconfined with this action for all couplings \cite{metoappear}.
831: If confinement were to
832: return when light quarks are added to this theory,
833: one would then have clear evidence of their essential role in
834: confinement.
835:
836: \begin{thebibliography}{99}
837:
838: \bibitem{bc} G. Bhanot and M. Creutz, Phys. Rev. D {\bf 24} (1981) 3212.
839: \bibitem{wegner}F.J. Wegner, J. Math. Phys. {\bf 12} (1971) 2259.
840: \bibitem{ggm}R.V. Gavai, M. Grady, and M. Mathur,
841: Nucl. Phys. B {\bf 423} (1994) 123.
842: \bibitem{rajivt6}M. Mathur and R.V. Gavai,
843: Nucl.Phys. {\bf B448} (1995) 399;
844: R.V. Gavai and M. Mathur, Phys. Rev. D {\bf 56} (1997) 32.
845: \bibitem{rajivsl}R.V. Gavai, Nucl. Phys. B {\bf 474} (1996) 446.
846: \bibitem{beg}M. Blume, V.J. Emery, and R.B. Griffiths,
847: Phys. Rev. A {\bf 4} (1971) 1071.
848: \bibitem{CL} P.M. Chaiken and T.C. Lubensky, {\em Principles of
849: Condensed Matter Physics}, Cambridge University Press, Cambridge, 1995.
850: \bibitem{su3} T. Blum et. al., Nucl. Phys B {\bf 442} (1995) 301.
851: \bibitem{fhk}J. Fingberg, U. Heller and F. Karsch, Nucl. Phys. B
852: {\bf 392} (1993) 493.
853: \bibitem{rajivft}R.V. Gavai, Nucl Phys. B, {\bf 586} (2000) 475.
854: \bibitem{hidden}M. Grady, report SUNY-FRE-04-03, hep-lat/0409163.
855: \bibitem{teper}A. Hart and M. Teper, Nucl. Phys. B
856: (Proc. Suppl.) {\bf 53} (1997) 497;
857: Nucl. Phys. B (Proc. Suppl.) {\bf 63} (1998) 522.
858: \bibitem{myq}M. Grady, Phys. Lett. B {\bf 455}, 239 (1999).
859: \bibitem{gmla}R. V. Gavai and M. Mathur,
860: Phys. Lett. B {\bf 458} (1999) 331.
861: %\bibitem{fisher} M.E. Fisher in {\em Proc. of Int. School of Physics
862: %Enrico Fermi Course LI, Critical Phenomena}, ed. M.S.Green,
863: %Academic Press, NY, 1971.
864: \bibitem{zp}M. Grady, Z. Phys C, {\bf 39} (1988) 125.
865: \bibitem{ps2}A. Patrascioiu, E. Seiler, and I.O. Stamatescu,
866: Nuovo Cimento D {\bf 11} (1989) 1165;
867: A. Patrascioiu, E. Seiler, V. Linke, and
868: I.O. Stamatescu, Nuovo Cim. {\bf 104B} (1989) 229.
869: \bibitem{gribov}V.N. Gribov, Phys. Scr. {\bf T15} (1987) 164;
870: Phys. Lett. B {\bf 194} (1987) 119.
871: \bibitem{abgribov} J. Nyiri, ed.,
872: {\em The Gribov Theory of Quark Confinement},
873: World Scientific, Singapore, 2001.
874: \bibitem{nc} M. Grady, Nuovo Cim. {\bf 105A} (1992) 1065.
875: \bibitem{cahill}K. Cahill and G. Herling,
876: Nucl. Phys. B Proc. Suppl. {\bf 73} (1999) 886;
877: K. Cahill, report NMCPP/99-7, hep-ph/9901285.
878: %\bibitem{cqm} Chr.V. Christov et. al., Prog. Nucl.
879: %Part. Phys. {\bf 37}, 91 (1996), and references therein.
880: \bibitem{markum}W. Feilmair, M. Faber, and H. Markum, Phys. Rev. D,
881: {\bf 39} (1989) 1409.
882: \bibitem{so3z2}M. Grady, report SUNY-FRE-98-09, hep-lat/9806024.
883: \bibitem{metoappear} M. Grady, to appear.
884:
885: \end{thebibliography}
886: \newpage
887: %\begin{center}
888: %{\Large Figure Captions}
889: %\end{center}
890: \begin{figure}[ht]
891: \includegraphics[width=5in]{FIGURE1.EPS}
892: \caption{Phase transitions on the fundamental-adjoint plane.
893: Nonlinear
894: axes have been chosen
895: so the entire coupling plane, including continuum limit at right and Z2
896: lattice gauge theory at top can be seen.
897: Scaling of couplings has been chosen to
898: reproduce ``look" of usual plot on linear axes.
899: Nonlinear scale for $\beta _A$
900: is shown at right. Diamonds are Bhanot-Creutz data \cite{bc},
901: triangles are second-order $8^3 \times 4$, and squares are
902: $8^3 \times 4$ first-order data from \cite{ggm}. Solid lines are
903: first-order, dashed second-order. Lines 1-4 are hypothetical
904: second-order lines for very large lattices, following perturbative
905: lines of constant physics. The line $\overline{\rm{AB}}$ is the deconfined
906: phase of the 4-d Z2 lattice gauge theory. }
907: \label{fig1}
908: \end{figure}
909: \newpage
910: \noindent
911: \begin{figure}[ht]
912: \includegraphics[width=2.6in]{FIGURE2A.EPS} \quad
913: \includegraphics[width=2.6in]{FIGURE2B.EPS}
914: \caption{
915: (a) Time history of an $8^4$ lattice run at $\beta _F = 1.2185$,
916: $\beta_A = 2.25$.
917: Upper trace is average plaquette, lower four traces are Ployakov loops.
918: Each successive Polyakov loop trace is offset by 0.2 for clarity.
919: Steplike
920: structure in plaquette appears associated with the
921: number of loops which show
922: spontaneous symmetry breaking at any time.
923: (b) Plaquette histogram on a $12^4$ lattice at $\beta _F = 1.2183$,
924: $\beta _A = 2.25$ showing clear multimodal distribution.}
925: \label{fig2}
926: \end{figure}
927: \newpage
928: \begin{figure}[ht]
929: \includegraphics[width=5in]{FIGURE3.EPS}
930: \caption{
931: Hysteresis sweeps at $\beta _A$ = 1.7, 1.25, and 1.05. Upper curves
932: are $12^4$ lattice, lower offset curves are $20^4$
933: lattice (scale at right).
934: $\beta _F$ is changed by 0.0001 after each Monte-Carlo sweep.
935: Near coincidence
936: of curves shows finite lattice size effects are small.}
937: \label{fig3}
938: \end{figure}
939: \newpage
940: \begin{figure}[ht]
941: \includegraphics[width=5in]{FIGURE4.EPS}
942: \caption{
943: Width of metastable region vs. latent heat squared. Diamonds are
944: for $12^4$ lattice, triangles for $20^4$ lattice. Linear fit on these axes
945: is predicted by the
946: tricritical hypothesis. The alternative hypothesis, that of an
947: ordinary critical point, predicts scaling directly with latent heat rather
948: than its square. The points comprise a
949: range in $\beta _A$ from 1.05 to 1.7.}
950: \label{fig4}
951: \end{figure}
952: \newpage
953: \begin{figure}[ht]
954: \includegraphics[width=2.6in]{FIGURE5A.EPS}\hspace*{.5in}
955: \includegraphics[width=2.6in]{FIGURE5B.EPS}\vspace*{.4in}
956: \includegraphics[width=2.6in]{FIGURE5C.EPS}
957: \caption{(a) A detailed look at $20^4$ hysteresis sweep at $\beta _A =1.7$.
958: Upper curve (right scale) is the difference in average
959: plaquettes as measured on
960: cooling vs. heating sweeps. Next four curves are Polyakov loops for heating
961: (beta decreasing) sweep. Final four curves, offset by 0.02 for clarity, are
962: Polyakov loops for cooling sweep. Polyakov loop curves are 25-point moving
963: averages to reduce variance. Spontaneous breaking of Polyakov loop is
964: clearly associated with plaquette tunneling events.
965: (b) Same at $\beta _A =1.25$.
966: (c) Same at $\beta _A = 1.0$. This lies below the tricritical point, in the
967: second-order region. Now there is no longer much separation between
968: the symmetry breakings for heating and cooling, but the
969: breaking still seems to be
970: coincident with the rising edge of the plaquette hysteresis. Here 100-point
971: moving averages are used for the Polyakov loops. Note the hysteresis
972: curve has lost the steep sides associated with
973: first-order tunneling, and its
974: magnitude is small compared to the first-order values.}
975: \label{fig5}
976: \end{figure}
977: \newpage
978: \begin{figure}[ht]
979: \includegraphics[width=5in]{FIGURE6.EPS}
980: \caption{Plot of $\beta _c$ for deconfinement transition on
981: asymmetric lattices, to test
982: possibility of $1/\ln (N_{\tau})$ scaling law.
983: Diamonds from Ref.~\cite{fhk}
984: are for Wilson action, squares from Ref.~\cite{rajivft}
985: are for Z2 monopole and vortex suppressed action.
986: Straight lines are linear fits to the data, excluding the $N_{\tau}=4$
987: points.
988: Dashed lines add a quadratic term to the fit.
989: Solid curved lines are, for comparison,
990: fits to the
991: normal two-loop scaling formula.}
992: \label{fig6}
993: \end{figure}
994: \end{document}
995: