0709.0669/g2.tex
1: \documentclass{JHEP3} % 10pt is ignored!
2: \usepackage{epsfig}
3: %
4: \title {$G_2$ gauge theory at finite temperature}
5: \author{Guido Cossu\\
6:         Scuola Normale Superiore, Piazza dei Cavalieri 27, 56126 Pisa\\
7:         and INFN Pisa, Largo B. Contecorvo 3 Ed.~C, 56127 Pisa, Italy\\
8:         E-mail: \email{g.cossu@sns.it}}
9: \author{Massimo D'Elia\\
10:         Dipartimento di Fisica \& INFN Genova, Via Dodecaneso 33, 16146 Genova, Italy\\ 
11:         E-mail: \email{delia@ge.infn.it}}
12: \author{Adriano Di Giacomo\\
13:         Dipartimento di Fisica \& INFN Pisa, Largo B. Contecorvo 3 Ed.~C, 56127 Pisa, Italy\\
14:         E-mail: \email{adriano.digiacomo@df.unipi.it}}
15: \author{Biagio Lucini\\
16:         Department of Physics, Swansea University, Singleton Park, Swansea SA2 8PP, UK\\
17:         E-mail: \email{b.lucini@swansea.ac.uk}}
18: \author{Claudio Pica\\
19:         Physics Department, Brookhaven National Laboratory, Upton, NY 11973, USA\\
20:         E-mail: \email{pica@bnl.gov}}
21: \abstract
22: {The gauge group being centreless, $G_2$ gauge theory is a good laboratory for studying the role of the centre of the group for colour confinement in Yang-Mills gauge theories. In this paper, we investigate $G_2$ pure gauge theory at finite temperature on the lattice. By studying the finite size scaling of the plaquette, the Polyakov loop and their susceptibilities, we show that a deconfinement phase transition takes place. The analysis of the pseudocritical exponents give strong evidence of the deconfinement transition being first order. Implications of our findings for scenarios of colour confinement are discussed.}
23: %
24: \keywords{Lattice Gauge Field Theories, Confinement}
25: %
26: \preprint{IFUP-TH/2007-22, BNL-NT-07/36}
27: %
28: \begin{document}
29: \section{Introduction}
30: Confinement is one of the most elusive problems in QCD. There is strong experimental evidence that quarks and gluons, which are the fundamental degrees of freedom of the theory, never appear as final states of strong interactions. It is still a challenge to understand how confinement is encoded in the QCD Lagrangian.\\
31: Following the large number of colours idea~\cite{'tHooft:1973jz}, it is reasonable to conjecture that confinement is a property of the gauge sector of the theory. Hence, it should be possible to solve the problem by looking at the pure gauge theory, and the solution should not be specific to a given number of colours $N$. For the pure gauge theory at finite temperature, it has been shown that confinement is lost at some critical temperature $T_c$~\cite{Borgs:1983wb}. The deconfinement phase transition in SU($N$) gauge theories can be understood in terms of the centre of the gauge group, which is $\mathbb{Z}_N$. An order parameter for the phase transition is the Polyakov loop
32: \begin{equation}
33: L(\vec{x},T) = \frac{1}{n} \mbox{Tr} \exp\left( i g \int_0^{1/T} A_0 \mbox{d} t \right) \ ,
34: \end{equation}
35: where $A_0$ is the gauge field in the compact direction, naturally associated
36: to the temperature $T$, whose length is $1/T$, $g$ is the gauge coupling and
37: $n$ the dimension of the fundamental representation (in SU($N$), $n =
38: N$). Since one dimension is compact, gauge transformations which are
39: continuous modulo $2 \pi/g$ are acceptable in the theory. Under those
40: transformations, $L(\vec{x},T) \to z L(\vec{x},T)$, where $z$ is an element of
41: $\mathbb{Z}_N$. If the centre symmetry is not broken, $\langle L \rangle =
42: (1/V)\int L(\vec{x},T) d^3 x = 0$ in the thermodynamic limit $V \to \infty$,
43: $V$ being the volume of the system. Conversely, a value of $\langle L \rangle$
44: different from zero implies breaking of the centre symmetry. It is possible to
45: show that at low temperatures $\langle L \rangle = 0$, while at high
46: temperatures $\langle L \rangle \ne 0$. Hence, a centre symmetry breaking phase
47: transition must take place. The expectation value of the Polyakov loop can be
48: related to the free energy $F$ of a static quark as
49: \begin{eqnarray}
50: L \propto e^{- \beta F} \ .
51: \end{eqnarray} 
52: It is then natural to identify the centre symmetry breaking phase transition with the deconfinement phase transition. In a famous paper~\cite{Svetitsky:1982gs}, Svetitsky  and Yaffe conjectured that the universality class of the deconfinement phase transition for SU($N$) gauge theory in D=d+1 dimensions is that of a d-dimensional $\mathbb{Z}_N$ Potts model, provided that the latter has a second order phase transition.  The Svetinsky-Yaffe conjecture has been verified numerically in 3+1 and 2+1 dimensions (see~\cite{Lucini:2003zr,Liddle:2005qb} for recent lattice calculations). It is interesting to remark that whenever the underlying spin model has a first order phase transition, so does the SU($N$) gauge theory.\\
53: This analysis hints toward the relevance of the centre for confinement. An
54: independent way to relate centre symmetry and confinement is presented
55: in~\cite{'tHooft:1977hy}, where confinement is described in terms of
56: condensation of vortices carrying magnetic flux. The allowed $N$ magnetic
57: fluxes are in one to one correspondence with the centre elements of the
58: group. Condensation of vortices in the confined phase means that the area
59: spanned by a Wilson loop randomly intersect vortex worldsheets. The resulting
60: cancelations determine the so-called area law for the Wilson loop, which is
61: one of the accepted criteria for colour confinement. Numerical works have
62: confirmed the vortex scenario~\cite{DelDebbio:1996mh}. To characterise the
63: transition in terms of a symmetry, the 't Hooft loop operator can be
64: introduced~\cite{'tHooft:1977hy}, which is expected to have a non-zero vacuum
65: expectation value in the confined phase and to be zero in average in the
66: deconfined phase. This behaviour has been checked numerically
67: in~\cite{deForcrand:2000fi,DelDebbio:2000cx,DelDebbio:2000cb, Kovacs:2000sy}.\\
68: While this scenario for colour confinement is perfectly consistent, the centre symmetry is lost when dynamical fermions are added to the action. Hence, either one gives up the idea that confinement in the pure Yang-Mills theory and in the full theory is basically the same phenomenon or we must assume that the centre is just a useful way to look at confinement, but does not embody any fundamental physics in relation to it. One possible way to look at this issue is to study the deconfinement phase transition in other gauge groups that have a different centre pattern. The physics of the phenomenon being inherently non-perturbative, lattice calculations are well suited for those investigations. In this context, SO($3$)$\equiv$ SU(2)/$\mathbb{Z}_2$ would be an ideal candidate: it is expected to confine (like SU(2), since the two groups share the same algebra), but has a trivial centre. Recent results suggest that a deconfinement phase transition takes place, but the presence of lattice artifacts ({\em centre monopoles}) makes it difficult to extract a reliable continuum limit~\cite{Barresi:2003jq}. Moreover, the centre structure of the underlying universal covering group (SU(2)) reflects in the existence of twist sectors, which might imply that the centre still plays a role, despite the group being centreless.\\
69: A different way to approach the problem is to use a fundamental group that is genuinely centreless\footnote{We use the word centreless to refer to a group whose centre is given only by the unity element.}. The simplest group in this category is the exceptional group $G_2$. There are other properties that make $G_2$ interesting for QCD: it contains SU(3) as a subgroup and (as in full QCD) an asymptotic string tension does not exist, since the colour charge carried by a quark can be completely screened by gluons~\cite{Holland:2003jy}. The existence of two phases has been proved in~\cite{Holland:2003jy}. However, this does not exclude that, instead of a real phase transition, a crossover separates the two phases. Were this the case, the physics of deconfinement in $G_2$ would be noticeably different from that of SU($N$) gauge theories, and this would cast serious doubts about what we can learn from $G_2$ for confinement in more physical gauge theories. While data reported in~\cite{Pepe:2005sz,Pepe:2006er} are compatible with a first order phase transition taking place, no exhaustive and detailed study of deconfinement has been performed so far. In this paper, we shall fill this gap by studying the finite size scaling behaviour of the plaquette, of the Polyakov loop and of their susceptibilities, from which we extract the critical exponents for the transition. We will then be able to show that a real transition takes place and that this transition is first order.\\
70: This work is organised as follows. In Sect.~\ref{g2} we will review the basic properties of the exceptional group $G_2$. Details of our lattice simulations are presented in Sect.~\ref{details}. Sect.~\ref{thermodynamics} contains our results and provides evidence for a first order deconfinement phase transitions occurring in $G_2$ at finite temperature. The implications of our findings for possible mechanisms of colour confinement are discussed in Sect.~\ref{discussions}. Finally, in Sect.~\ref{conclusions} we summarise the main points of our investigation. 
71: %
72: \section{Basic properties of the exceptional group $G_2$}
73: \label{g2}
74: We begin by summarising some basic properties of the Lie Group $G_2$. In mathematical terms this is the group of automorphisms of the octonions and it can be naturally constructed as a subgroup of the real group $SO(7)$ - which has 21 generators and rank 3. Besides the usual properties of $SO(7)$ matrices
75: \begin{equation}
76:   \det \Omega = 1 \qquad \Omega^{-1} = \Omega^{T}
77: \end{equation}
78: we have in addition another constraint
79: \begin{equation}
80:   T_{abc} = T_{def} \Omega_{da} \Omega_{eb} \Omega_{fc}
81:   \label{constraint}
82: \end{equation}
83: where $T_{abc}$ is a totally antisymmetric tensor whose nonzero elements are (using the octonion basis given by \cite{Cacciatori:2005yb})
84: \begin{equation}
85:   T_{123} = T_{176} = T_{145} = T_{257} = T_{246} = T_{347} = T_{365} = 1.
86: \end{equation}
87: Equations~(\ref{constraint}) are 7 independent relations reducing the numbers of generators to 14. The fundamental representation of $G_2$ is 7 dimensional. Using the algebra representation of \cite{Cacciatori:2005yb} (we refer to appendix \ref{G2algebra} for details) we can clearly identify an $SU(3)$ subgroup and several $SU(2)$ subgroups,  6 of which are sufficient to cover the whole group, a useful property for MC simulations. The first three $SU(2)$ subgroups are in the $4 \times 4$ real representation of the group while the remaining three are a mixture of the $4 \times 4$ and the $3 \times 3$ representations and are extremely difficult to simulate with standard heat-bath techniques. See the next section for details on simulations.\\
88: The following relations hold:
89: \begin{equation}
90:   SU(3) \subset G_2  \Rightarrow \mathcal C(G_2) \subset \mathcal {\rm{Centr}}(SU(3)) = \mathbb Z_3
91: \end{equation}
92: in which ${\rm{Centr}}(SU(3))$ is the centralizer of $SU(3)$ (i.e. the
93: matrices in $G_2$ that commute with every element in $SU(3)$). Intersections of
94: centralizers of different $SU(3)$ subgroups give
95: \begin{equation}
96:  \mathcal C(G_2)= \{1\}
97: \end{equation}
98:  i.e. a trivial centre.\\
99: The Lie group $G_2$ has rank 2, like $SU(3)$. This implies that the residual
100: symmetry after an Abelian projection is $U(1)^2$, its Cartan subgroup. Stable monopole solutions are classified according to the homotopy group\footnote{The first equality follows from $\pi_1(G_2) = 0$. See for example~\cite{Weinberg:1996kr}.}:
101: \begin{equation}
102:   \pi_2(G_2/U(1)^2) = \pi_1(U(1) \times U(1)) = \mathbb Z \times \mathbb Z 
103: \end{equation}
104: i.e. we have two distinct species of monopoles, classified by elements of the discrete group $\mathbb Z_2$, as for $SU(3)$. 
105: An extension of the 't Hooft tensor - the gauge invariant field of monopoles - can be written for the $G_2$ gauge group so Abelian monopole solutions are really possible in this theory.\\
106: Another interesting homotopy group shows that centre vortices are absent in the theory:
107: \begin{equation}
108:   \pi_1(G_2/\mathcal C(G_2)) = \pi_1(G_2) = 0
109: \end{equation}
110: while for $SU(3)$ for example
111: \begin{equation}
112:   \pi_1(SU(3)/\mathbb Z_3) = \mathbb Z_3
113: \end{equation}
114: and 
115: \begin{equation}
116:   \pi_1(SO(3)/\{ 1\}) = \pi_1(SO(3)) = \mathbb Z_2 \neq 0
117: \end{equation}
118: as stated before. So $G_2$ is a good playground to study the dual superconductor picture in a theory without centre vortices, thus isolating monopole contribution in confinement.
119: %
120: \section{Simulations of $G_2$ Lattice Gauge Theory}
121: \label{details}
122: In this work we are going to investigate the thermodynamical properties of the gauge group $G_2$ (see also \cite{Holland:2003jy, Pepe:2005sz}). To simulate the pure gauge theory
123: \begin{equation}
124: \mathcal L = \frac{1}{7g^2} {\rm Tr} \, F_{\mu\nu}F_{\mu\nu}
125: \end{equation}
126: with the Wilson action, we used a simple Cabibbo-Marinari update (heat-bath + overrelaxation in a tunable ratio, for every step) for the first three $SU(2)$ subgroups ($4 \times 4$ representation, set 1,3 and 4 in appendix A) spanning the $SU(3) \subset G_2$. This simple setting cannot be used for the remaining three subgroups because the integration measure is not as simple. We make a random gauge transformation every $n$ updates (tipically 1 or 2) to guarantee the ergodicity of the algorithm\footnote{The matrices for random gauge transformation are regenerated every step by a random algorithm to assure that no periodicities or orbits in phase space can arise.}.
127: To study the thermodynamical properties we simulated several
128: asymmetric lattices $N_t \times N_s^3$ of spatial dimensions $N_s =
129: 12, 14, 16, 18, 20, 24, 32$ and temporal dimension $N_t = 6$ ($N_t =
130: 4$ only for the smallest lattice). An average of 20 $\beta$s per lattice have
131: been simulated. The temperature of the system is
132: given by $T = (a(\beta) N_t)^{-1}$, where $a(\beta)$ is the lattice spacing
133: as a function of $\beta = 1/7 g^2$. The critical behaviour of the
134: system has been extracted by applying the theory of finite size
135: scaling (FSS), which has been used to extrapolate the behaviour of the
136: observables we have studied to the thermodynamic limit ($N_s \to \infty$). We
137: needed histories of order $10^5$ updates near the transition (1 week on a
138: 1.5GHz Opteron processor for a medium lattice).\\
139: The code is highly optimized and very fast (using only real algebra), is
140: written using explicitly assembler SSE2 instructions in single precision for
141: the matrix-multiplication core and run on an Opteron farm in the computer
142: facilities of the Physics Department of the University of Pisa. 
143: 
144: The observables we have measured are the standard plaquette and the Polyakov loop. A clarification is in order here. While one should expect to be able to characterise the critical behaviour of a system by looking at the plaquette, doubts could be cast into the usefulness of the Polyakov loop: since $G_2$ is centreless, the Polyakov loop is not an order parameter for a possible deconfining phase transition. In principle, phase transitions can be reliably investigated only by using an order parameter field, whose critical behaviour characterises the transition itself. However, in order to prove that a transition takes place and to determine the critical indices, a non-trivial overlap on the order parameter is the only property we need\footnote{The reverse of this sentence is not true: no conclus§ion can be drawn from the absense of critical behaviour in a non-order parameter field.}. Hence, if we can observe a divergence in the peak of the Polyakov loop susceptibility (and of the specific heat, whose reliability is hard to question) we can safely conclude that a phase transition takes place.\\  
145: The theory of FSS predicts that as a function of the volume the
146: maximum of susceptibilities scale in the following way:
147: \begin{equation}
148: \label{fss}
149: \chi \sim a \cdot L^{\frac{\gamma}{\nu}} + b \ ,
150: \end{equation}
151: where $\gamma$ is the critical exponent of the generating quantity (in
152: our case either the plaquette or the Polyakov loop) and $\nu$ is the
153: critical exponent related to the divergence of the correlation
154: length. 
155: The position of the maximum scales as
156: \begin{eqnarray}
157: \label{betac}
158: \beta_c(L) = \beta_c(\infty) + c L^{-1/\nu} \ ,
159: \end{eqnarray}
160: where $\beta_c(L)$ is the pseudocritical $\beta$ for size $L$ and
161: $\beta_c(\infty)$ is the critical value of $\beta$.
162: This analysis also applies to first order phase transition,
163: whose signature is given by $\gamma = 1$ and $\nu = 1/d$, with $d$ the
164: dimension of the system. \\
165: 
166: %
167: \section{Thermodynamics of $G_2$ gauge theory}
168: \label{thermodynamics}
169: We studied the thermodynamics of this theory using the typical observables, the plaquettes 
170: \begin{equation} 
171: P_s = \frac{1}{3 \cdot 7 N_s^3 N_t} \sum_{\square_s} {\rm Tr} U_{\square_s} \qquad  P_t = \frac{1}{3 \cdot 7 N_s^3 N_t} \sum_{\square_t} {\rm Tr} U_{\square_t}
172: \end{equation}
173: where the two sums are  on space-space and space-time plaquettes respectively. The peak of the susceptibility
174: \begin{equation}
175: \chi_P = N_s^3 (\langle P^2 \rangle - \langle P \rangle^2) \qquad P = (P_s + P_t)/2
176: \end{equation}
177: signals the phase transition point. This quantity (often referred to in the literature as the "lattice specific heat") is only part of the (physical) specific heat, whose complete reconstruction requires various correlators weighted with different coefficients; nonetheless, this is a singular piece from which the critical scaling behaviour can be inferred.
178: 
179: %This is only
180: %part of the specific heat, whose complete reconstruction deserves different
181: %coefficients for the several correlators, nonetheless it is a singular piece
182: %from which the critical scaling behaviour can be inferred. 
183: 
184: %This is not exactly the specific heat,
185: %which deserves different coefficients for the several correlators, 
186: %nonetheless it has the correct scaling behaviour. 
187: We also measured the Polyakov loop and its susceptibility:
188: \begin{equation}
189: L = \frac {1}{N_s^3}\sum_{\vec x}\Bigl (\frac {1}{7}\prod_{t=0}^{N_t-1} U_4 (\vec x) \Bigr ) \qquad \chi_L = N_s^3 (\langle L^2 \rangle - \langle L \rangle^2).
190: \end{equation}
191: \FIGURE[ht]{
192:   \epsfig{file=./FIGS/Plaquette_susceptibility.eps,width=0.85\linewidth, clip=}
193:   \caption{Plaquette susceptibility plotted against $\beta$. The peak signals
194:   the bulk transition while the peak corresponding to the physical transition
195:   for $N_t = 6$ is shown in the inset. We also show results from a simulation at T=0 on a $16^4$ lattice (black triangle points).}
196:   \label{susc-plaq}
197: }
198: \FIGURE[ht]{
199:   \epsfig{file=./FIGS/Plaquette_bulk_transition.eps,width=0.85\linewidth,clip=}
200:   \caption{Comparison of finite and zero temperature simulations. In the box:    magnification of the physical transition region (reweighted curves).}
201:   \label{Comparison}
202: }
203: \FIGURE[ht]{
204: \centering
205: \begin{tabular}{cc}
206:   \epsfig{file=./FIGS/Plaquette_Peak_Scaling.eps,width=0.45\linewidth,clip=}&
207:   \epsfig{file=./FIGS/Plaquette_susceptibility_reweighted_rescaled.eps,width=0.48\linewidth,clip=}
208: \end{tabular}
209:   \caption{Left: scaling of the peak of plaquette susceptibility with the
210:   volume. The continuous line is a linear fit to the data, as
211:   explained in the text. Right: FSS of the plaquette susceptibility assuming a
212:   first order transition. For this plot, we have used the value $\beta_c = 1.395$, obtained from the fit to the position of the maximum according to~(\ref{betac}).} 
213:     \label{Plaq_scaling}
214: }
215: \FIGURE[ht]{
216:   \epsfig{file=./FIGS/Polyakov_Loop_Density_12.eps,width=0.85\linewidth, clip=}
217:   \vspace{20pt}
218:   \\
219:   \epsfig{file=./FIGS/Polyakov_Loop_Density_16.eps,width=0.85\linewidth, clip=}
220:   \caption{Normalized densities of the Polyakov Loop in a semilog plot for
221:     $\beta$ varying in the range from 1.35, the critical coupling of the bulk
222:     transition ``$\beta_{bulk}$'', to 1.401, in the deconfined phase (data
223:     from the $6\times14^3$ lattice for the upper graph and from $6\times16^3$
224:     for the other - same scales and limits for both axes are used for better
225:     comparison). As an aside we notice that far in the confined phase,
226:     $\beta_c < 1.395$, the Polyakov loop is zero within errors and this feature can not be explained on the ground of any manifest symmetry of the system. Continues on next page}
227:   \label{PolDensity1}
228: }
229: \FIGURE[th]{
230:   \epsfig{file=./FIGS/Polyakov_Loop_Density_20.eps,width=0.85\linewidth, clip=}
231:   \caption{Continues from last page ($6\times20^3$ lattice).}
232:   \label{PolDensity2}
233: }
234: The lattices considered for the scaling analysis are only the $N_s = 12, 14, 16, 18,
235: 20$ times $N_t = 6$ for the following reasons. The computational cost of
236: locating the transition grows exponentially fast with the volume; anticipating here a first
237: order transition, the intrinsic problem is that two (or more) phases
238: coexist. The simulated system tunnels between pure phases by building an
239: interface of size $N_s$. The free-energy cost of such a mixed configuration is
240: $\sigma N_s^{D-1}$ ($\sigma$ being the surface tension), the interface is
241: built with probability $\exp(-\sigma N_s^{D-1})$ and the natural time scale
242: for the simulation grows with $N_s$ as $\exp(\sigma N_s^{D-1})$. This is
243: called exponential critical slowing down and makes simulations impractical for
244: lattices with $N_s > 20$ for a reliable estimate of susceptibilities. Looking
245: at Figs. \ref{PolDensity1}, \ref{PolDensity2} and comparing the densities in the tunneling region
246: for the three different lattices gives an idea of the problem, common to all
247: systems exhibiting a first order transition. Multicanonical methods
248: \cite{Janke:1998} will be needed for feasible simulations on such large
249: lattices. The other reason concerns the number of time slices and is related
250: to the presence of an unphysical bulk transition that we shall explain below (see also Fig. \ref{susc-plaq}). Being very close to the bulk transition, the
251: physical deconfinement transition for $N_t = 4$ is extremely difficult to
252: detect, the signal being highly contaminated by the ``noise'' coming from the
253: bulk. $N_t = 6$ is needed to be sufficiently away from the bulk. By increasing furtherly $N_t$, one can move
254: the physical transition far away from the bulk transition point. Hence,
255: choosing a larger $N_t$ will clean the signal from the bulk ``noise''. To
256: investigate this possibility, we performed some simulations at $N_t = 8$,
257: which confirmed the general features of the $N_t = 6$ simulation. The
258: displacement of the critical $\beta$ was clearly visible but not sufficient to
259: bring any practical advantage over the $N_t = 6$ calculation, while the simulation time increased considerably. For this reason, we sticked to the $N_t = 6$ calculation, giving up the possibility of performing a continuous limit extrapolation of the critical temperature. However, our pilot study at $N_t = 8$ suggests that there is no reason to doubt that such a continuous limit exists.\\
260: \FIGURE[ht]{ 
261:   \centering
262:   \begin{tabular}{cc}
263:     \epsfig{file=./FIGS/Plaquette_History_4x12_1.3594.eps,width=0.45\linewidth,
264:       clip=}&
265:     \epsfig{file=./FIGS/Polyakov_Loop_History_6x20_1.395.eps,width=0.45\linewidth,clip=}
266:   \end{tabular}
267:   \caption{Left: MC history of the plaquette ($\beta = 1.3594,
268:       12^3\times4$). Right: A typical Monte Carlo history of the Polyakov loop (data from
269:       $\beta = 1.395, 20^3\times6$). }
270:   \label{Histories}
271: }
272: \FIGURE[th]{
273:   \epsfig{file=./FIGS/Polyakov_Loop_Rescaling.eps,width=0.85\linewidth, clip=}
274:   \caption{Scaling of the Polyakov loop assuming first order. For the smallest
275:     lattice $12^3\times 6$ corrections to the scaling are evident (even the
276:     lattice $14^3\times 6$ is not big enough but corrections are reduced); $\beta_c = 1.395$ as explained in the text.}
277: \label{RescalingPol_up}
278: }
279: \FIGURE[b]{
280:   \epsfig{file=./FIGS/Polyakov_Loop_Peak_Scaling.eps,width=0.85\linewidth, clip=}
281:   \caption{Scaling of the peak of $\chi_L$. The solid line is a linear
282:   fit to the data.} 
283:   \label{RescalingPol_down}
284: }
285: In a finite volume no divergences can arise, since the partition function is analytical. Nevertheless critical indices can be measured by looking at the scaling with the volume of the plaquette susceptibility (related to the specific heat $C_V$). The height of the peak for a first-order transition scales with the volume $V$ and the width and the displacement from the real critical point of the peak position scales as $1/V$ (plus corrections to this leading behaviour).\\
286: A pronounced peak is present at any volume and $N_t$ and always at the same
287: $\beta \sim 1.35$. There is no scaling with volume and no movement toward the
288: weak coupling region passing from $N_t = 4$ to $N_t = 6$ as we would expect
289: for a physical transition. This transition is the equivalent of the bulk phase
290: transition in SU($N$) gauge theories, and separates the (physical) weak
291: coupling region from the (unphysical) strong coupling one. The bulk peak
292: almost completely overshadows the real physical transition, a smaller peak in
293: the weak coupling region at $\beta \sim 1.395$ for $N_t = 6$. This peak scales
294: with the volume, provided that the bulk contribution has been subtracted. This
295: subtraction procedure is needed in order to disentangle the physics from the
296: discretisation artifacts. To estimate the bulk background, we simulated the
297: system also at zero temperature on $16^4$ and $20^4$ lattices (to control
298: systematic errors). The bulk contribution has
299: to be subtracted from the plaquette susceptibility for a correct finite
300: scaling analysis. This procedure could be seen as a normalisation of the free
301: energy following the request that this quantity be zero at zero
302: temperature. The influence of the bulk transition on the plaquette
303: susceptibility is shown in Fig. \ref{susc-plaq}. The nature of the two transitions manifests itself comparing finite temperature and zero temperature simulations in Fig. \ref{Comparison}. The integral of the difference between the two curves is the free energy density:
304: \begin{equation}
305:   \frac{f}{T^4}\Bigr|_{\beta_0}^{\beta} = - N_\tau^4 \int_{\beta_0}^\beta d\beta^\prime (P_0 - P_T)
306: \end{equation}
307: in which $P_0$ and $P_T$ are the mean plaquettes at zero and finite temperature
308: respectively.  At the bulk transition $f$ is zero within errors and
309: develops a value different from zero at the physical transition.\\
310: The MC time history of the plaquette is
311: displayed in Fig.~\ref{Histories} (left), and shows a two-phase
312: structure typical of first order phase transitions.
313: The extracted maxima of the plaquette susceptibility ($\propto C_V$) using the reweighted data are shown in
314: Fig. \ref{Plaq_scaling}. Maxima and their errors are estimated by a simple
315: inspection of the reweighting output. A linear fit of the form $y = a \cdot x + b$
316: (see Eq.~\ref{fss}) gives $a = 0.00079(14)\cdot10^{-3},\, b = 0.98(62)\cdot10^{-3},\,
317: \chi^2_{\rm red} = 1.35$, providing good evidence for a first order
318: phase transition. A fit according to Eq.~(\ref{betac}) gives $\beta_c(\infty) = 1.3950(4)$.\\
319: The Polyakov loop is insensitive to the bulk transition so we used it
320: to detect the position of the physical one, even if, strictly
321: speaking, this quantity is not an order parameter. The Polyakov loop develops
322: an evident double peak structure typical of a first order transition (see
323: Figs.~\ref{PolDensity1},\ref{PolDensity2}). In this semilog plot is also clear, by looking
324: at the relative ratio of peaks height and valley height near the
325: transition point, the exponential decreasing of tunneling
326: probability with the volume. In Fig.~\ref{Histories} we show the typical
327: Monte Carlo history of the Polyakov loop. Once again, a clean two-state
328: signal appears. This reflects in a double-peak structure of the
329: observable shown e.g. in Fig.~\ref{PolDensity2}. The same FSS analysis as for the specific heat
330: again gives evidence of a first order transition, with a good $\chi_{\rm red}^2$ in the
331: linear fits of peak heights (Figs.~\ref{RescalingPol_up}~and~\ref{RescalingPol_down}). The parameters of the linear fit of the peak
332: heights $y = a \cdot x + b$ are $a = 0.1183(2),\, b = 60(5),\,
333: \chi^2_{\rm red} = 0.61$. A subtraction of the background is
334: understood. The background is assumed to be weakly dependent on coupling
335: $\beta$. This is an educated guess suggested by the zero temperature
336: simulations. The background is estimated by mean of a linear fit of the tails
337: of the peak and being an ultraviolet effect, it is assumed to be the same for
338: all volumes. In practice we took the smallest lattices $6 \times 12^3$, $6
339: \times 14^3$ and some
340: of the extremal points in tails for the fit. The number of points is
341: unessential giving practically the same parameters and a good $\chi_{\rm red}^2$. The Polyakov loop susceptibility can be also used to determine $\beta_c(\infty)$. Using formula~(\ref{betac}), we get $\beta_c(\infty) = 1.3951(2)$, which is compatible with the result obtained from the susceptibility of the plaquette.
342: %
343: \section{Discussion}
344: \label{discussions}
345: As we have stated in the introduction, an asymptotic string tension in
346: $G_2$ does not exist. Hence, one can question whether this group is
347: confining. This is mostly a semantic
348: problem. In~\cite{Greensite:2006sm} it is argued that because of the
349: absence of the asymptotic string, $G_2$ gauge theory is not
350: confining. This would fit the idea of confinement as related to centre
351: vortices randomly piercing the Wilson loop. Sharing this view means to
352: accept the logical conclusion that full QCD (in which an asymptotic
353: string tension does not exist because of quark pair production) is not
354: a confining theory. Since it is common understanding that QCD confines, the
355: essence of confinement must be found in some other property of the
356: theory. In our opinion, this property is a low-energy dynamics
357: dominated by glueballs and mesons (which are colour-singlet
358: states). Colour-singlet states are also present in $G_2$ at zero
359: temperature. At high temperature the dynamics is instead dominated by
360: a gluon plasma. In this sense, despite the absense of an asymptotic
361: string tension, $G_2$ gauge theory is a confining theory. Accepting this statement means to infer that centre degrees of freedom are not related to confinement (unless one want to put all the weight of the centre on the trivial element, see~\cite{Greensite:2006sm}). Hence, the degrees of freedom responsible for colour confinement must be searched for in other properties of the gauge group.\\
362: Like SU(3), $G_2$ is a rank two group, i.e. it has two Cartan generators\footnote{A Cartan generator is a generator which commutes with all the others.}. It is then an attractive possibility that like in SU($N$) pure gauge theories~\cite{DiGiacomo:1999fa,DiGiacomo:1999fb,Carmona:2001ja} and in full QCD~\cite{Carmona:2002ty,D'Elia:2005ta} the mechanism for colour confinement is related to the condensation of magnetic monopoles, as it seems to be the case also for the SO($3$) gauge theory~\cite{Barresi:2004qa}. An investigation in this direction is currently in progress, and will be reported elsewhere.
363: %
364: \section{Conclusions}
365: \label{conclusions}
366: We studied the thermodynamics of the Yang-Mills theory with gauge
367: group $G_2$. The presence of an unphysical transition (most probably
368: due to the choice of the discretised action used in simulations) makes
369: the problem harder. Nevertheless a physical transition is found by
370: looking at plaquette and Polyakov loop susceptibilities. Time
371: histories of the Polyakov group and the plaquette show double peaks
372: typical of first order transitions. A detailed FSS
373: analysis agrees with the first order hypothesis.
374: Hence, we can conclude that $G_2$ gauge theory has two distinct phases
375: separated by a jump in the free energy. Those phases are immediately
376: identified with the confined (low temperature) and deconfined (high
377: temperature) phase. The same dynamics characterises SU($N$) Yang-Mills
378: theories at finite temperature. Since $G_2$ does not have a
379: (non-trivial) centre, our findings suggest that the dynamics of colour
380: confinement cannot be directly related to the centre of the gauge
381: group, as it has been inferred from previous works on SU($N$) gauge
382: theories. At this stage, the possibility that dual superconductivity
383: of the vacuum explains colour confinement is still open. The next step
384: of our study is to investigate the FSS of the monopole creation
385: operator, to test if the dual superconductor picture of confinement
386: works also for $G_2$ gauge theory.
387: 
388: \acknowledgments
389: \label{acknowledgments}
390: We would thank M.~Pepe for various useful discussions on the topic. The work
391: of C.P. has been supported in part by contract DE-AC02-98CH1-886 with the
392: U.S. Department of Energy and B.L. is supported by the Royal Society.
393: 
394: 
395: %
396: \appendix
397: \section{$G_2$ algebra representation}
398: \label{G2algebra}
399: In this appendix we simply report a representation of the 14 generators of the $G_2$ group \cite{Cacciatori:2005yb}. They are normalized such that $\rm{tr} (C_i C_j) = - \delta_{ij}$. The first 8 matrices generate the $SU(3) \subset G_2$. Here is also a list of 6 $SU(2)$ subroups that cover the entire group (useful for the Cabibbo-Marinari update):
400: \begin{enumerate}
401: \item $C_1, C_2, C_3$
402: \item $\sqrt{3} C_8, \sqrt{3}C_9, \sqrt{3}C_{10}$
403: \item $C_4, C_5, \frac{(C_3 + \sqrt{3} C_8)}{2}$
404: \item $C_6, C_7, \frac{(C_3 - \sqrt{3} C_8)}{2}$
405: \item $\frac{(3 C_3 - \sqrt{3} C_8)}{2}, \sqrt{3}C_{11}, \sqrt{3} C_{12}$
406: \item $\frac{(3 C_3 + \sqrt{3} C_8)}{2}, \sqrt{3}C_{13}, \sqrt{3} C_{14}$
407: \end{enumerate}
408: 
409: \section{Algebra}
410: \small
411: $$
412: C_1 =\frac{1}{2}\left(
413: \begin{array}{ccccccc}
414: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
415: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
416: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
417: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
418: 0 & 0 & 0 & 0 & 0 & -1 & 0 \\
419: 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
420: 0 & 0 & 0 & 1 & 0 & 0 & 0
421: \end{array}
422: \right)
423: \qquad
424:  C_2 =\frac{1}{2}\left(
425: \begin{array}{ccccccc}
426: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
427: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
428: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
429: 0 & 0 & 0 & 0 & 0 & 1 & 0 \\
430: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
431: 0 & 0 & 0 & -1 & 0 & 0 & 0 \\
432: 0 & 0 & 0 & 0 & 1 & 0 & 0
433: \end{array}
434: \right)
435: $$
436: $$
437: C_3 =\frac{1}{2}\left(
438: \begin{array}{ccccccc}
439: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
440: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
441: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
442: 0 & 0 & 0 & 0 & -1 & 0 & 0 \\
443: 0 & 0 & 0 & 1 & 0 & 0 & 0 \\
444: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
445: 0 & 0 & 0 & 0 & 0 & 1 & 0
446: \end{array}
447: \right)
448: \qquad
449: C_4 =\frac{1}{2}\left(
450: \begin{array}{ccccccc}
451: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
452: 0 & 0 & 0 & 0 & 0 & 0 & 1 \\
453: 0 & 0 & 0 & 0 & 0 & 1 & 0 \\
454: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
455: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
456: 0 & 0 & -1 & 0 & 0 & 0 & 0 \\
457: 0 & -1 & 0 & 0 & 0 & 0 & 0
458: \end{array}
459: \right)
460: $$
461: $$
462: C_5 =\frac{1}{2}\left(
463: \begin{array}{ccccccc}
464: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
465: 0 & 0 & 0 & 0 & 0 & -1 & 0 \\
466: 0 & 0 & 0 & 0 & 0 & 0 & 1 \\
467: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
468: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
469: 0 & 1 & 0 & 0 & 0 & 0 & 0 \\
470: 0 & 0 & -1 & 0 & 0 & 0 & 0
471: \end{array}
472: \right)
473: \qquad
474: C_6 =\frac{1}{2}\left(
475: \begin{array}{ccccccc}
476: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
477: 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
478: 0 & 0 & 0 & -1 & 0 & 0 & 0 \\
479: 0 & 0 & 1 & 0 & 0 & 0 & 0 \\
480: 0 & -1 & 0 & 0 & 0 & 0 & 0 \\
481: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
482: 0 & 0 & 0 & 0 & 0 & 0 & 0
483: \end{array}
484: \right)
485: $$
486: $$
487: C_7 =\frac{1}{2}\left(
488: \begin{array}{ccccccc}
489: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
490: 0 & 0 & 0 & -1 & 0 & 0 & 0 \\
491: 0 & 0 & 0 & 0 & -1 & 0 & 0 \\
492: 0 & 1 & 0 & 0 & 0 & 0 & 0 \\
493: 0 & 0 & 1 & 0 & 0 & 0 & 0 \\
494: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
495: 0 & 0 & 0 & 0 & 0 & 0 & 0
496: \end{array}
497: \right)
498: \qquad
499: C_8 =\frac 1{2\sqrt 3} \left(
500: \begin{array}{ccccccc}
501: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
502: 0 & 0 & -2 & 0 & 0 & 0 & 0 \\
503: 0 & 2 & 0 & 0 & 0 & 0 & 0 \\
504: 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
505: 0 & 0 & 0 & -1 & 0 & 0 & 0 \\
506: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
507: 0 & 0 & 0 & 0 & 0 & 1 & 0
508: \end{array}
509: \right)
510: $$
511: $$
512: C_9 =\frac 1{2\sqrt 3} \left(
513: \begin{array}{ccccccc}
514: 0 & -2 & 0 & 0 & 0 & 0 & 0 \\
515: 2 & 0 & 0 & 0 & 0 & 0 & 0 \\
516: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
517: 0 & 0 & 0 & 0 & 0 & 0 & 1 \\
518: 0 & 0 & 0 & 0 & 0 & -1 & 0 \\
519: 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
520: 0 & 0 & 0 & -1 & 0 & 0 & 0
521: \end{array}
522: \right)
523: \qquad
524: C_{10} =\frac 1{2\sqrt 3} \left(
525: \begin{array}{ccccccc}
526: 0 & 0 & -2 & 0 & 0 & 0 & 0 \\
527: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
528: 2 & 0 & 0 & 0 & 0 & 0 & 0 \\
529: 0 & 0 & 0 & 0 & 0 & -1 & 0 \\
530: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
531: 0 & 0 & 0 & 1 & 0 & 0 & 0 \\
532: 0 & 0 & 0 & 0 & 1 & 0 & 0
533: \end{array}
534: \right)
535: $$
536: $$
537:  C_{11} = \frac 1{2\sqrt 3} \left(
538: \begin{array}{ccccccc}
539: 0 & 0 & 0 & -2 & 0 & 0 & 0 \\
540: 0 & 0 & 0 & 0 & 0 & 0 & -1 \\
541: 0 & 0 & 0 & 0 & 0 & 1 & 0 \\
542: 2 & 0 & 0 & 0 & 0 & 0 & 0 \\
543: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
544: 0 & 0 & -1 & 0 & 0 & 0 & 0 \\
545: 0 & 1 & 0 & 0 & 0 & 0 & 0
546: \end{array}
547: \right)
548: \qquad
549: C_{12} = \frac 1{2\sqrt 3} \left(
550: \begin{array}{ccccccc}
551: 0 & 0 & 0 & 0 & -2 & 0 & 0 \\
552: 0 & 0 & 0 & 0 & 0 & 1 & 0 \\
553: 0 & 0 & 0 & 0 & 0 & 0 & 1 \\
554: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
555: 2 & 0 & 0 & 0 & 0 & 0 & 0 \\
556: 0 & -1 & 0 & 0 & 0 & 0 & 0 \\
557: 0 & 0 & -1 & 0 & 0 & 0 & 0
558: \end{array}
559: \right)
560: $$
561: $$
562: C_{13} = \frac 1{2\sqrt 3} \left(
563: \begin{array}{ccccccc}
564: 0 & 0 & 0 & 0 & 0 & -2 & 0 \\
565: 0 & 0 & 0 & 0 & -1 & 0 & 0 \\
566: 0 & 0 & 0 & -1 & 0 & 0 & 0 \\
567: 0 & 0 & 1 & 0 & 0 & 0 & 0 \\
568: 0 & 1 & 0 & 0 & 0 & 0 & 0 \\
569: 2 & 0 & 0 & 0 & 0 & 0 & 0 \\
570: 0 & 0 & 0 & 0 & 0 & 0 & 0
571: \end{array}
572: \right)
573: \qquad
574: C_{14} = \frac 1{2\sqrt 3} \left(
575: \begin{array}{ccccccc}
576: 0 & 0 & 0 & 0 & 0 & 0 & -2 \\
577: 0 & 0 & 0 & 1 & 0 & 0 & 0 \\
578: 0 & 0 & 0 & 0 & -1 & 0 & 0 \\
579: 0 & -1 & 0 & 0 & 0 & 0 & 0 \\
580: 0 & 0 & 1 & 0 & 0 & 0 & 0 \\
581: 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
582: 2 & 0 & 0 & 0 & 0 & 0 & 0
583: \end{array}
584: \right)
585: $$
586: %
587: \bibliographystyle{JHEP}
588: \bibliography {g2}
589: %
590: \end{document}