1: %
2: %
3: %
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \chapter{The infrared behaviour of Green functions}
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %
8: %
9: %
10: \label{chapter_IR}
11:
12: There are compelling reasons to think that confinement is a property of QCD, and
13: does not result from some other theory. One of such indications
14: is a non-zero value of the string tension found in the lattice simulations.
15: As a matter of fact lattice simulations give access to many non-perturbative
16: quantities. We are particularly interested in knowing the Landau gauge Green functions at low momenta,
17: i.e. at energies of order and below than $\Lqcd$. No free quarks
18: or gluons exist at very small momenta because of confinement.
19: So, a study of gluonic correlation functions in the deep-infrared
20: domain may seem useless. However, in order to study the property of confinement
21: from the first principles one has to understand the change in behaviour
22: of Green functions found at low momenta. Knowing these functions exactly
23: would be a great support for the future development of the theory, because many
24: confinement scenarii (for example the Gribov-Zwanziger scenario) give predictions for
25: low energies momentum dependence of the Green functions in Landau gauge.
26: Lattice simulations allow to test these predictions.
27:
28: Lattice results for different Green functions of QCD have been
29: successfully tested at large momenta by perturbation theory up to
30: NNNLO (see chapter~\ref{chapter_UV_behaviour} and~\cite{Boucaud:2005np},\cite{Boucaud:2005xn}). We shall see below that
31: lattice Green functions also satisfy the complete ghost Schwi\-nger-Dyson
32: equation (see Figure~\ref{SDcheck}). Thus lattice approach
33: gives consistent results non only in the ultraviolet domain but also
34: in the infrared one. Of course numerical methods could never
35: give us complete Green functions for all possible values of momenta.
36: Nevertheless, lattice gives a \emph{quasi-unique method for testing different
37: analytical approaches}, like study of truncated system of Schwinger-Dyson equations,
38: renormalisation group flow equations and other non-perturbative
39: relations like Slavnov-Taylor identities.
40:
41: Most of analytical predictions are done for the infrared exponents $\alpha_F$ and $\alpha_G$
42: that describe power-law deviations from free propagators when $p\rightarrow 0$
43: %
44: \begin{equation}
45: \label{PowerLawParametrization}
46: \begin{array}{l}
47: p^2 G^{(2)}(p^2) \equiv G(p^2)\propto \left(\frac{p^2}{\lambda_G^2}\right)^{\alpha_G} \\
48: p^2 F^{(2)}(p^2) \equiv F(p^2)\propto \left(\frac{p^2}{\lambda_F^2}\right)^{\alpha_F} \\
49: \ldots
50: \end{array}
51: \end{equation}
52: %
53: where $\lambda_{G,F}$ are some fixed parameters of dimension one. When $p$ is large, the functions
54: $F(p^2)$ and $G(p^2)$ are logarithmic functions of momentum, and thus $\alpha_{F,G}=0$. But at low momenta
55: it may not be true. In fact, power-law behaviour is the crudest approximation, allowing to exhibit the most general features
56: of the momentum dependence of the Green functions in the infrared.
57: The real law governing the infrared gluodynamics might be much more complicated.
58: In the following section we review (very briefly) different predictions
59: for the exponents $\alpha_{G}$ and $\alpha_{F}$ of, respectively, the gluon and ghost
60: two-point functions. Next we present our analysis of the Slavnov-Taylor identities
61: imposing some limits on these exponents. After this we turn to the study of
62: the ghost Schwinger-Dyson equation and test the
63: widely accepted relation (\ref{R}) between the exponents $\alpha_{G}$ and $\alpha_{F}$.
64: Our conclusion (supported by numerical simulations) is that this relation is not valid.
65: We revisit the usual proof and conclude that either the ghost-gluon vertex behaves
66: unexpectedly in the infrared, or that $\alpha_{F}=0$. At the end of the chapter
67: we discuss the results of direct fits of two-point functions, and compare
68: our results with other lattice collaborations.
69:
70:
71: %
72: %
73: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74: \section{Review of today's analytical results}
75: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
76: %
77: %
78: \label{Review_of_today_s_analytical_results}
79:
80: In this section we quote the main analytical results regarding the infrared
81: exponents (\ref{PowerLawParametrization}). We start with the Zwanziger's
82: prediction obtained for gluonic correlation functions. Next we present the
83: results of other analytical approaches.
84:
85: %
86: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
87: \subsection{Zwanziger's prediction}
88: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89: %
90: \label{subsection_Zwanziges_prediction}
91:
92: Zwanziger suggested in~\cite{Zwanziger:1990by},\cite{Zwanziger:1991gz} that the (Landau gauge) gluon propagator
93: vanishes at zero momentum in the infinite volume limit. The argument is the following.
94: The Faddeev-Popov operator (\ref{MFPlat}) is positive definite at a local minimum of the functional
95: (\ref{LatticeGaugeMinimisingFunctional})
96: %
97: \begin{equation}
98: \label{poz}
99: (\omega,\MFPlat[U]\omega)\ge 0.
100: \end{equation}
101: %
102: Choosing a test vector
103: %
104: \begin{equation}
105: \omega^a(x)=\frac{\exp{i\frac{2\pi\emu}{L} x}}{\sqrt{V}}\chi^a,
106: \end{equation}
107: %
108: where the normalised colour vector $\chi^a$ is an eigenfunction of the ``angular momentum" operator
109: $\left(J^b\right)_{ac} = if^{abc}$, one obtains from (\ref{MFPlat}) and (\ref{poz})
110: the following limit on the mean colour spin
111: %
112: \begin{equation}
113: \label{Zwanziger_estimate}
114: \left\arrowvert\frac{1}{V}\sum_{x} A_\mu(x)\right\arrowvert\le 2\tan{\frac{\pi}{L} }.
115: \end{equation}
116: %
117: Introducing an external colour field $H^a_\mu$ source (independent of $x$),
118: one obtains from (\ref{Zwanziger_estimate}) an estimate for the generating functional
119: %
120: \begin{equation}
121: \frac{Z(H)}{Z(0)}=\frac{1}{Z(0)} \int [\mathcal{DA}] e^{-S[\mathcal{A}] + H^a_\mu \sum_{x}A^a_\mu(x)} \le e^{2V\sum_{\mu}
122: \left\arrowvert H_\mu \right\arrowvert\tan{\frac{\pi}{L}}}.
123: \end{equation}
124: %
125: The free energy density $w(H)=\frac{1}{V}\log{Z[H]}$ is convex and bounded from below ($w(0)=0$), thus
126: one has
127: %
128: \begin{equation}
129: 0\le w(H)\le2\sum_\mu\arrowvert H_\mu \arrowvert\tan{ \frac{\pi}{L} }.
130: \end{equation}
131: %
132: All connected gluonic Green functions can be obtained by calculating the variations of the free energy
133: with respect to the external sources $H_\mu^a(x)$. The last inequality suggest that in the infinite
134: volume limit all Green functions vanish at zero momentum. However, the inversion of derivation and
135: thermodynamic limit is not supported by a rigorous proof.
136:
137: %
138: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
139: \subsection{Study of truncated SD and ERG equations}
140: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
141: %
142:
143: Diverse analytical approaches (study of truncated Schwinger-Dyson equations
144: and of renormalisation group equation, see Table~\ref{Predictions_analytiques})
145: agree that the infrared divergence of the ghost propagator is enhanced,
146: i.e. $\alpha_F \leq 0$; while they predict different values for $\alpha_G$,
147: mostly around $\alpha_G\approx 1.2$. This means that the gluon
148: propagator is suppressed in the infrared. However, some groups obtain $\alpha_G\leq 1$, i.e.
149: an infrared-divergent gluon propagator.
150: Lattice simulations confirmed the prediction for the ghost propagator,
151: whereas the lattice gluon propagator seems to remain finite and non-zero
152: in the infrared, i.e. $\alpha_G = 1$. We discuss this question in details in the section~\ref{section_direct_fits}.
153: %
154: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
155: \begin{table}[!h]
156: \centering
157: \begin{small}
158: \begin{tabular}{l|lcc}
159: \hline\hline
160: {\bf Reference} &{\bf Method} & \boldmath$\alpha_G$\unboldmath & \boldmath$\alpha_F$\unboldmath
161: \\ \hline
162: Zwanziger~\cite{Zwanziger:1990by} & see subsection~\ref{subsection_Zwanziges_prediction} & $>1$ & no
163: \\ \hdashline[0.4pt/1pt]
164: Bloch~\cite{Bloch:2003yu} & SD truncation + perturbation theory & $[0.34,1.06]$ & $[-0.53,-0.17]$
165: \\
166: von Smekal et al.~\cite{vonSmekal:1997is} & SD truncation & $1.84$ & $-0.92$
167: \\
168: Zwanziger~\cite{Zwanziger:2001kw} & SD truncation + Zwanziger condition & $2$ or $1.19$ & $-1$ or $-0.595$
169: \\
170: Aguilar et al.~\cite{Aguilar:2004sw} & SD equation & $0.98$ & $-0.04$
171: \\ \hdashline[0.4pt/1pt]
172: Kato~\cite{Kato:2004ry} & ERGE & $0.292$ & $-0.146$
173: \\
174: Pawlowski et al.~\cite{Pawlowski:2003hq} & ERGE & $1.19$ & $-0.595$
175: \\
176: Fischer et al.~\cite{Fischer:2004uk} & ERGE & $1.02$ & $-0.52$
177: \\
178: \hline\hline
179: \end{tabular}
180: \caption{\footnotesize\it Summary of various analytical predictions}
181: \label{Predictions_analytiques}
182: \end{small}
183: \end{table}
184: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185: %
186:
187:
188:
189: %
190: %
191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192: \section{Constraints on the infrared exponents and the Slavnov-Taylor identity}
193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194: %
195: %
196: \label{section_Slavnov_Taylor_IR}
197:
198: Let us consider the Slavnov-Taylor identity (\ref{STid}) relating the three-gluon vertex
199: $\Gamma_{\lambda \mu \nu}$, the ghost-gluon vertex $\widetilde{\Gamma}_{\lambda\mu}(p,q;r)$
200: and the ghost and gluon propagators:
201: %
202: \begin{equation}
203: \label{ST}
204: p^\lambda\Gamma_{\lambda \mu \nu} (p, q, r) =
205: \frac{F(p^2)}{G(r^2)} (\delta_{\lambda\nu} r^2 - r_\lambda r_\nu) \widetilde{\Gamma}_{\lambda\mu}(r,p;q) -
206: \frac{F(p^2)}{G(q^2)} (\delta_{\lambda\mu} q^2 - q_\lambda q_\mu) \widetilde{\Gamma}_{\lambda\nu}(q,p;r).
207: \end{equation}
208: %
209: Taking the limit $r \rightarrow 0$ keeping $q$ and $p$ finite, and using
210: the parametrisation $G(r^2) \simeq \left( r^2\right)^{\alpha_G}$ valid for $r^2\ll \Lqcd^2$,
211: one finds the following limits on the infrared exponents
212: %
213: \begin{equation}
214: \label{STpredictions}
215: \left\{
216: \begin{array}{ll}
217: \alpha_G < 1 & \text{\small gluon propagator \emph{diverges} in the infrared, and} \\
218: \alpha_F \le 0 & \text{\small the divergence of the ghost propagator is \emph{unchanged} or \emph{enhanced} in the infrared }
219: \end{array}
220: \right.
221: \end{equation}
222: %
223:
224: \noindent Let us discuss in details the origin of the limits (\ref{STpredictions}). The ghost-gluon
225: vertex $\widetilde{\Gamma}_{\mu\nu}(p,k;q)$ may be parametrised~\cite{Ball:1980ax} in the most general way as
226: %
227: \bea
228: \widetilde{\Gamma}_{\mu}^{abc}(p,k;q) & = & f^{abc} (-i p_\nu) g_0
229: \widetilde{\Gamma}_{\nu\mu}(p,k;q) \label{vertghost} \\
230: &=& f^{abc} (-i p_\nu) g_0
231: %\nonumber \\
232: %& &
233: \cdot \left[
234: \delta_{\nu\mu} a(p,k;q) - q_\nu k_\mu b(p,k;q) + p_\nu q_\mu c(p,k;q) + \nonumber
235: \right. \\ &&\qquad\qquad\qquad\quad + q_\nu p_\mu d(p,k;q) + \left. p_\nu p_\mu e(p,k;q)\right]
236: \label{vertgluon}
237: \eea
238: %
239: We recall that in this formula $-p$ is the momentum of the outgoing ghost, $k$ is the momentum of the
240: incoming one and $q=-p-k$ the momentum of the gluon (all momenta are taken as entering).
241: For some particular kinematic configurations we use the following dense notations
242: %
243: \begin{equation}
244: \begin{array}{l}
245: a_3(p^2) = a(-p,p; 0 ) \\
246: a_1(p^2) = a(0, -p; p), \quad b_1(p^2) = b(0, -p; p), \quad d_1(p^2) = d(0, -p; p).
247: \end{array}
248: \end{equation}
249: %
250: The limit $r^2\rightarrow 0 $ leads to an asymmetric kinematic configuration for the three-gluon
251: vertex in the l.h.s. of (\ref{ST}). This particular configuration allows a general parametrisation
252: ~\cite{Chetyrkin:2000dq}
253: %
254: \begin{small}
255: \begin{equation}
256: \Gamma_{\mu\nu\rho}(p,-p,0) =
257: \left( 2\delta_{\mu\nu}p_\rho - \delta_{\mu\rho}p_\nu - \delta_{\rho\nu}p_\mu\right) T_1(p^2) -
258: \left( \delta_{\mu\nu} - \frac{p_\mu p_\nu}{p^2} \right) p_{\rho}T_2(p^2) +
259: p_\mu p_\nu p_\rho T_3 (p^2).
260: \end{equation}
261: \end{small}
262: %
263: with functions $T_{1,2,3}(p^2)$. The scalar function $T_1(p^2)$ is proportional to the
264: gauge coupling in the $\widetilde{\text{MOM}}$ renormalisation scheme.
265: Now, exhibiting the dominant part of each term in (\ref{ST}), we obtain
266: %
267: %\begin{small}
268: \begin{eqnarray}
269: \label{STlimr0}
270: & T_1(q^2)\left(q_\mu q_\nu - q^2 \delta_{\mu\nu}\right)+q^2 T_3(q^2)q_\mu q_\nu +\eta_{1\mu\nu}(q,r)= \nonumber \\
271: & \frac{F\left((q+r)^2\right)}{G(r^2)}\Big[(a_1(q^2)+r_1(q,r))\left(\delta_{\mu\nu}r^2-r_\mu r_\nu\right)
272: + \left(b_1(q^2)+r_2(q,r)\right)
273: q_\mu\left(r^2 q_\nu-(q\cdot r) r_\nu\right) + \nonumber
274: \\& +(b_1(q^2)+d_1(q^2)+r_3(q,r))r_\mu(r^2 q_\nu - (q\cdot r) r_\nu)\Big] + \nonumber
275: \\
276: & +\frac{F\left((q+r\right)^2)}{G(q^2)}\Big[a_3(q^2)(q_\mu q_\nu - q^2 \delta_{\mu\nu}) +\eta_{2\mu\nu}(q,r)\Big]
277: \end{eqnarray}
278: %\end{small}
279: %
280: with $r_{1,2,3}$ and $\eta_{1,2}$ satisfying
281: %
282: \begin{align}
283: \lim_{r\rightarrow 0 }r_1(q,r) = \lim_{r\rightarrow 0 }r_2(q,r) = \lim_{r\rightarrow 0 }r_3(q,r) = 0
284: \nonumber
285: \\
286: \lim_{r\rightarrow 0 }\eta_{1\mu\nu}(q,r) = \lim_{r\rightarrow 0 } \eta_{2\mu\nu}(q,r) = 0
287: \end{align}
288: %
289: Identifying the leading terms of the scalar factors multiplying the tensors $q_\mu q_\nu$ and
290: $\left(q_\mu q_\nu - q^2 \delta_{\mu\nu}\right)$ we obtain the usual relations (\cite{Chetyrkin:2000dq}):
291: %
292: \begin{equation}
293: \label{WIhab}
294: \begin{array}{l}
295: T_1(q^2) = \frac{F(q^2)}{G(q^2)}a_3(q^2) \\
296: T_3(q^2) = 0.
297: \end{array}
298: \end{equation}
299: %
300: Using these relations in (\ref{STlimr0}) we get
301: %
302: \begin{equation}
303: \lim_{r\rightarrow 0}\, \frac{F(p^2)}{G(r^2)}\Big[a_1(q^2)\left(r^2\delta_{\mu\nu}-r_\mu r_\nu\right)+b_1(q^2)
304: (r^2 q_\mu q_\nu -(r\cdot q) q_\mu r_\nu)\Big] = 0.
305: \end{equation}
306: %
307: Thus one sees that if
308: %
309: \begin{equation}
310: \label{cond_a1}
311: a_1(q^2)\neq 0\quad\text{or}\quad b_1 \neq 0
312: \end{equation}
313: %
314: then (\ref{ST}) can only be compatible with the parametrisation~(\ref{PowerLawParametrization}) if
315: %
316: \begin{equation}
317: \alpha_G < 1.
318: \end{equation}
319: %
320: The condition (\ref{cond_a1}) is satisfied because at large momentum one has to
321: all orders $a_1(p^2)$ = 1 ( because of the non-renormalisation
322: theorem ~\cite{Taylor:1971ff},\cite{Chetyrkin:2000dq}).
323:
324: We can also, instead of letting $r \to 0$, take the limit $p\rightarrow 0$ of (\ref{STid})
325: as is done in \cite{Chetyrkin:2000dq}. The dominant part of the l.h.s. of (\ref{STid}) is
326: %
327: \begin{equation}
328: \left(2\delta_{\mu\nu}(p\cdot q)-p_\mu q_\nu-p_\nu q_\mu \right)T_1(q^2)-
329: \left(\delta_{\mu\nu}-\frac{q_\mu q_\nu}{q^2}\right)(p\cdot q) T_2(q^2)
330: +(p\cdot q) q_\mu q_\nu T_3(q^2)
331: \end{equation}
332: %
333: The r.h.s. is the product of $F(p^2)$ with an expression of at least first order in p. $T_1$ and
334: $T_2$ being different from zero we can conclude in this case that $\alpha_F \le 0$.
335:
336: Let us repeat here that all these considerations are valid only if all scalar factors
337: of the ghost-ghost-gluon and three-gluons vertices are regular functions when one momentum
338: goes to zero while the others remain finite. Under those quite reasonable hypotheses
339: one obtains important constraints on the gluon and ghost propagators - namely that they are both divergent in
340: the zero momentum limit, and the divergence of the ghost propagator is enhanced.
341:
342: Let us stress that the limit (\ref{STpredictions}) on $\alpha_G$ disagrees
343: with many other analytical predictions quoted in the section~\ref{Review_of_today_s_analytical_results}.
344:
345: %
346: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
347: \section{Relation between the infrared exponents}
348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
349: %
350: \label{section_Relation_between_the_infrared_exponents}
351:
352: The Schwinger-Dyson equation for the two-point correlation function (and for the
353: quark propagator, but we consider only pure Yang-Mills case here) has the simplest
354: form among other non-perturbative relations between Green functions. It has
355: been used to constrain the the infrared exponents. Even more, there is a
356: a commonly accepted relation between the infrared exponents
357: %
358: \begin{equation}
359: \label{R}
360: 2\alpha_F + \alpha_G = 0.
361: \end{equation}
362: %
363: which we shall discus now. The origin of this relation is the dimensional analysis of
364: the Schwinger-Dyson equation for the ghost propagator
365: %
366: \begin{equation}
367: \label{SDghost_IR}
368: \frac{1}{F(k)} = 1 + g_0^2 N_c \int \frac{d^4 q}{(2\pi)^4}
369: \left( \rule[0cm]{0cm}{0.8cm}
370: \frac{F(q^2)G((q-k)^2)}{q^2 (q-k)^2}
371: \left[
372: \frac{(k\cdot q)^2 - k^2q^2}{k^2(q-k)^2}
373: \right]
374: \ H_1(q,k)
375: \right),
376: \end{equation}
377: %
378: where $H_1(q,k)$ is one of the scalar functions defining the ghost-gluon vertex:
379: %
380: \begin{equation}
381: \label{H12_definition}
382: q_{\nu'} \widetilde{\Gamma}_{\nu'\nu}(-q,k;q-k) = q_\nu H_1(q,k) + (q-k)_\nu H_2(q,k),
383: \end{equation}
384: %
385: where $H_{1,2}$ are functions of the factors $a,b,c,d,e$ (\ref{vertghost}).
386: The large momentum behaviour (\cite{Chetyrkin:2000dq},\cite{Taylor:1971ff})
387: of this vertex depends on the kinematic configuration:
388: %
389: \begin{equation}
390: \begin{array}{l}
391: \frac{p_{\mu}p_{\nu}}{p^2} \cdot \widetilde{\Gamma}_{\mu\nu}^{\overline{\text{\tiny MS}}}(-p,0;p) = 1\quad\text{to \emph{all} orders}
392: \\
393: \\
394: \frac{p_{\mu}p_{\nu}}{p^2} \cdot \widetilde{\Gamma}_{\mu\nu}^{\overline{\text{\tiny MS}}}(-p,p;0) = 1 + \frac{9}{16\pi}\alpha_s(p^2) + \ldots
395: \end{array}
396: \end{equation}
397: %
398: Note that in the case of the vanishing momentum of the out-going ghost (and only in this case)
399: the non-renormalisation theorem is applicable~\cite{Taylor:1971ff} and hence
400: %
401: \begin{equation}
402: \label{NonRenormalization}
403: H_1(q,0) + H_2(q,0)=1.
404: \end{equation}
405: %
406: Let us now consider two infrared scales $k_1\equiv k$ and $k_2\equiv \kappa
407: k$. Calculating the difference of the Schwinger-Dyson equation (\ref{SDghost_IR})
408: taken at scales $k_1$ and $k_2$ and supposing for the moment that $\alpha_F\neq 0$ one obtains
409: %
410: \begin{equation}
411: \label{k1k2}
412: \begin{split}
413: \frac{1}{F(k)} - \frac{1}{F(\kappa k)} & \propto
414: (1-\kappa^{-2\alpha_F} ) (k^{2})^{-\alpha_F}
415: = g_0^2 N_c \int \frac{d^4 q}{(2\pi)^4}
416: \left( \rule[0cm]{0cm}{0.8cm}
417: \frac{F(q^2)}{q^2} \left(\frac{(k\cdot q)^2}{k^2}-q^2\right) \right.
418: \times \\ & \left. \times
419: \left[ \rule[0cm]{0cm}{0.6cm}
420: \frac{G((q-k)^2)H_1(q,k)}{\left((q-k)^2\right)^2} -
421: \frac{G((q-\kappa k)^2)H_1(q,\kappa k)}{\left((q-\kappa k)^2\right)^2}
422: \rule[0cm]{0cm}{0.6cm} \right]
423: \rule[0cm]{0cm}{0.8cm} \right).
424: \end{split}
425: \end{equation}
426: %
427: This integral equation, as well as the initial equation (\ref{SDghost_IR}),
428: is written in terms of bare Green functions, and the integral may contain
429: ultraviolet divergences. It can be cast into a well-defined renormalised form by
430: multiplying (in (\ref{SDghost})) $G^{(2)}$ (resp. $F^{(2)}$) by $Z_3^{-1}$ (resp. $\widetilde Z_3^{-1}$)
431: and the bare coupling $g_0^2$ by $Z_g^{-2} = Z_3 \widetilde Z_3^2$,
432: and finally multiplying the $k^2$ term by $\widetilde Z_3$.
433: However, in the subtracted equation (\ref{k1k2}) all ultraviolet
434: divergences are cancelled, as well as the $\widetilde Z_3 k^2$ term.
435: Thus the subtracted Schwinger-Dyson equation holds both
436: in terms of bare and renormalised Green functions without
437: any explicit renormalisation factors.
438:
439: We now make the hypothesis that there exists a scale $q_0$ below which the
440: power-law parametrisation is valid
441: %
442: \begin{equation}
443: \label{def_q0}
444: G(q^2) \ \sim \ (q^2)^{\alpha_G}, \quad
445: F(q^2) \ \sim \ (q^2)^{\alpha_F} , \quad \mbox{\rm for} \quad q^2 \le q_{0}^2.
446: \end{equation}
447: %
448: The equation (\ref{NonRenormalization}) suggests that if \emph{both} functions $H_{1,2}$
449: are non-singular then one can suppose $H_1(q,k)\simeq 1$ in (\ref{k1k2}),
450: and (\ref{R}) is straightforward by a dimensional analysis. However, we have a priori no reason
451: to think that the scalar functions $H_1(q,k)$ and $H_2(q,k)$ are \emph{separately} non-singular
452: for all $q,k$. Writing for example~\footnote{In fact there are many possible parametrisation. We
453: choose (\ref{parametrization_H1}) in order to illustrate the argument that follows.}
454: %
455: \begin{equation}
456: \label{parametrization_H1}
457: H_1(q,k) \ \sim \ (q^2)^{ \alpha_\Gamma } \ h_1\left(\frac{q\cdot k}{q^2},\frac{k^2}{q^2} \right),
458: \end{equation}
459: %
460: with a non-singular function $h_1$,
461: we keep all the generality of the argument admitting a singular behaviour of
462: the scalar factor $H_1(q,k)$. Doing the dimensional analysis of the equation (\ref{k1k2})
463: \emph{without} putting $H_1(q,k)\simeq 1$, we obtain that the relation (\ref{R})
464: is satisfied if and only if the following triple condition is verified~\cite{Boucaud:2005ce}:
465: %
466: \begin{equation}
467: \label{conditions}
468: 2\alpha_F + \alpha_G = 0 \quad \longleftrightarrow \quad
469: \left\{
470: \begin{array}{l}
471: \alpha_F \ne 0 \\
472: \alpha_\Gamma=0 \\
473: \alpha_F +\alpha_G < 1
474: \end{array}
475: \right.
476: \end{equation}
477: %
478: All possible cases and limits obtained from the integral convergence
479: conditions are given in Table~\ref{tabledescas}. As we shall see the case $2$ is excluded
480: by lattice simulations. The case $4$ is particularly interesting,
481: it corresponds to the situation when the power-law infrared behaviour of the ghost propagator is the same as
482: in the free case, and no relation between the infrared exponents follows from the Schwinger - Dyson equation.
483: We shall return to this the discussion of this case in the section~\ref{section_direct_fits}.
484: %
485: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486: \begin{table}[!h]
487: \centering
488: \begin{small}
489: \begin{tabular}{l|c|c|c|c}
490: \hline\hline
491: \textbf{case} & \textbf{1} & \textbf{2} & \textbf{3} & \textbf{4} \\
492: \hdashline[0.4pt/1pt]
493: & $\alpha_F \ne 0$&$\alpha_F \ne 0$&$\alpha_F = 0$&$\alpha_F = 0$\\
494: &$\alpha_F +\alpha_G +\alpha_\Gamma < 1$&$\alpha_F +\alpha_G +\alpha_\Gamma \ge1$&$\alpha_G +\alpha_\Gamma < 1$&$\alpha_G +\alpha_\Gamma \ge1$\\
495: \hline
496: \textbf{IR} &&&&\\
497: \textbf{convergence}&$\alpha_F +\alpha_\Gamma > -2$&$\alpha_F +\alpha_\Gamma> -2$ &$\alpha_\Gamma> -2$&$ \alpha_\Gamma> -2$\\
498: \textbf{conditions}&$\alpha_G +\alpha_\Gamma> -1$&$\alpha_G +\alpha_\Gamma> -1$&$ \alpha_G +\alpha_\Gamma> -1$&$\alpha_G +\alpha_\Gamma> -1$\\
499: \hline
500: \textbf{SD}&&&&\\
501: \textbf{constraints}& $2\alpha_F + \alpha_G +\alpha_\Gamma
502: = 0$& $\alpha_F =-1$&excluded&none\\
503: \hline\hline
504: \end{tabular}
505: \caption{\footnotesize\it Summary of the various cases regarding the $\alpha$ coefficients}
506: \label{tabledescas}
507: \end{small}
508: \end{table}
509: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
510: %
511:
512: The first and the last conditions (\ref{conditions}) are compatible with limits coming from the analysis
513: of the Slavnon-Taylor identity (\ref{STpredictions}), and are also consistent with
514: lattice simulations (see section~\ref{section_direct_fits}, \cite{Boucaud:2005ce}).
515: If one of the conditions (\ref{conditions}) is not verified then, according to the Table~\ref{tabledescas},
516: (\ref{R}) should be replaced by
517: %
518: \begin{equation}
519: \label{R_Gamma}
520: 2\alpha_F + \alpha_G + \alpha_\Gamma = 0.
521: \end{equation}
522: %
523: In the following section we present the results of a numerical test of the relation (\ref{R}),
524: and thus we probe the validity of the condition on $\alpha_\Gamma$.
525:
526: One remark regarding the power-law parametrisation is in order. Suppose for the moment that this
527: parametrisation is exact below the scale $q_0$ defined in (\ref{def_q0}).
528: Then one can differentiate (\ref{k1k2}) $n$ times with respect to $\kappa$, keeping $q,k$ finite. We obtain
529: %
530: \begin{equation}
531: \left( k^2 \right)^{-2\alpha_F}\left(-2\alpha_F\right)\cdot\ldots\cdot
532: \left( -2\alpha_F -n \right) \kappa^{-2\alpha_F -n} \propto \int d^4 q \,
533: \frac{d^n}{d^n \kappa} \left( \frac{G((q-\kappa k)^2)H_1(q,\kappa k)}{\left((q-\kappa k)^2\right)^2} \right).
534: \end{equation}
535: %
536: The r.h.s of the last equation is not equal to zero for finite $k$, and thus one immediately
537: has
538: %
539: \begin{equation}
540: \alpha_F \neq -\frac{n}{2},\qquad n=1,2,\ldots.
541: \end{equation}
542: %
543: Thus any half-integer predictions for $\alpha_F$ should be considered as an indication of
544: incompleteness of the power-law parametrisation (\ref{PowerLawParametrization}).
545:
546: %
547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
548: \section{Lattice study of the ghost Schwinger-Dyson equation}
549: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
550: %
551:
552:
553: %
554: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
555: \subsection{Complete ghost Schwinger-Dyson equation in the lattice formulation}
556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
557: %
558:
559: In order to derive the discretized version of the ghost Schwinger-Dyson equation
560: we repeat the same steps as in the continuum case (\ref{SD_simple_begin} - \ref{SD_simple_end})
561: but for the lattice version of the Faddeev-Popov operator (\ref{MFPlat}).
562: We define the covariant Laplacian
563: %
564: \begin{equation}\label{DU}
565: \begin{array}{l}
566: %
567: \Delta_U^{ab} = \sum_\mu \left(G_\mu^{ab}(x)
568: \left(
569: \delta_{x,y} - \delta_{y,x+\emu}
570: \right)
571: -
572: G_\mu^{ab}(x-\emu)
573: \left(
574: \delta_{y,x-\emu} - \delta_{y,x}
575: \right)
576: \right).
577: \end{array}
578: \end{equation}
579: %
580: \noindent The appearance of $\Delta_U$ in (\ref{MFPlat}) is due to the appropriate
581: discretisation of the usual Laplacian operator $\Delta$, dictated by
582: the non-locality of derivatives in the lattice formulation, i.e.
583: replacement of the $\nabla$ operator by its covariant version.
584:
585: Multiplying (\ref{MFPlat}) by $F^{(2)}(x,y)$ from the right, one obtains
586: %
587: \begin{equation}
588: \label{SD_L1}
589: \begin{split}
590: & \frac{1}{N_c^2 -1} \Delta_U^{ab} (y,z) F_{\text{1conf}}^{(2)ba}(U;z,x) = \delta_{y,u} -
591: \\ -
592: \frac{f^{abc}}{2(N_c^2 -1)} &
593: \left[
594: A_\mu^c(y) F_{\text{1conf}}^{(2)ba}(U;y+\emu,x) - A_\mu^c(y-\emu) F_{\text{1conf}}^{(2)ba}(U;y-\emu,x)
595: \right].
596: \end{split}
597: \end{equation}
598: %
599: This is an exact mathematical identity for each gauge configuration $U$,
600: and thus the consequences that can be derived from this relation are free of
601: any ambiguity originating from the presence of Gribov copies.
602: Performing an averaging $\langle \bullet \rangle$ over the configurations $U$ one gets
603: %
604: \begin{equation}
605: \label{SD_L1_av}
606: \begin{split}
607: & \frac{1}{N_c^2 -1} \tr \left\langle\Delta_U(y,z) F_{\text{1conf}}^{(2)}(z,x)\right\rangle = \delta_{y,x} -
608: \\ & -
609: \frac{f^{abc}}{2(N_c^2 -1)}
610: \left\langle A_\mu^c(y) F_{\text{1conf}}^{(2)ba}(U,y+\emu,x) - A_\mu^c(y-\emu) F_{\text{1conf}}^{(2)ba}(U,y-\emu,x)
611: \right\rangle
612: \end{split}
613: \end{equation}
614: %
615: This averaging procedure depends on the way chosen to treat the Gribov problem:
616: the particular set of configurations over which it is performed depends on the prescription which is adopted
617: (fc/bc procedures on the lattice, restriction to the fundamental modular region;
618: see the subsection~\ref{subsection_Lattice_Green_functions_and_Gribov} for details).
619: Consequently, the Green functions may vary but they must in any case satisfy the above equation, even
620: when the volume of the lattice is finite.
621:
622: Like in the continuum case, we perform a Fourier transform and obtain:
623: %
624: \begin{small}
625: \begin{equation}
626: \label{SD_L1_av2}
627: \begin{split}
628: & \frac{1}{N_c^2 -1} \tr \sum_{x} e^{i p\cdot x} \left\langle\Delta_U(0,z) F_{\text{1conf}}^{(2)}(U,z,x)
629: \right\rangle = 1
630: - i \sin(p_\mu)
631: \frac{f^{abc}}{(N_c^2 -1)}
632: \left\langle A_\mu^c(0) \tilde{F}_{\text{1conf}}^{(2)ba}(U,p) \right\rangle
633: \end{split}
634: \end{equation}
635: \end{small}
636: %
637: Although the equations (\ref{SD_L1}) and (\ref{SD_L1_av}) have to be exactly verified by lattice data,
638: the relation (\ref{SD_L1_av2}) does only approximately (within statistical errors) since it
639: relies on translational invariance, which could be guaranteed only if we used an infinite
640: number of Monte-Carlo configurations.
641:
642: The presence of $\Delta_U$ in the last equation is due to non-zero lattice spacing effects.
643: Indeed, lattice perturbation theory possesses an infinite number of ghost-gluon vertices
644: depending on the lattice spacing $a$, giving tadpole contributions like the one presented at
645: the Figure~\ref{tadpole}.
646: %
647: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
648: \begin{figure}[!h]
649: \begin{center}
650: \includegraphics[width=0.3\linewidth]{EPS/ghost_ghost_gl_gl}
651: \end{center}
652: \caption{\footnotesize\it Example of the terms in the Schwinger-Dyson
653: equation on the lattice.}
654: \label{tadpole}
655: \end{figure}
656: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
657: %
658: Such tadpole contributions may be estimated by a mean field method~\cite{Lepage:1992xa}.
659: Using the average plaquette $\langle P \rangle$ (for $\beta=6.0$ $\langle P\rangle \simeq 0.5937$) one predicts
660: a tadpole correction factor $\propto \langle P \rangle^{-(1/4)} \simeq 1.14$.
661: These terms disappear in the continuum limit, but they do so only very slowly : the
662: tadpole corrections (1 - plaquette) vanish only as an inverse logarithm with the
663: lattice spacing. This is to be contrasted with the corrections arising in the r.h.s
664: which are expected to be of order $a^2$. Our lattice calculation~\cite{Boucaud:2005ce} gives
665: %
666: \begin{equation}
667: \Delta_U \simeq \Delta/\left(1.16 \pm 0.01 \right),
668: \end{equation}
669: %
670: almost independently of the momentum. This is in good agreement with
671: the correction factor $1.14$ quoted above.
672:
673: We see from Figure~\ref{SDcheck} that the lattice Green functions
674: match pretty well the SD equation (\ref{SD_L1_av2}) in both the ultraviolet and
675: infrared regions.
676: %
677: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
678: \begin{figure}[!h]
679: \begin{center}
680: \includegraphics[angle=-90, width=0.7\linewidth]{EPS/SD_with_DeltaU1}
681: \end{center}
682: \caption{\footnotesize\it Checking that lattice Green functions satisfy
683: the ghost SD equation (\ref{SD_L1_av2}). We plot $\frac{1}{1.16}\widetilde{F}(p^2)
684: - g_0 \frac{p_\mu}{N_c^2 -1} f^{abc}\langle A^c_\mu(0)\,\widetilde{F}_{\text{1conf}}^{(2)ba}
685: (\mathcal{A},p)\rangle $ compared to $1$.}
686: \label{SDcheck}
687: \end{figure}
688: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
689: %
690: % %
691: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692: % \begin{figure}[!h]
693: % \begin{center}
694: % \includegraphics[angle=-90, width=0.9\linewidth]{EPS/WITH_DELTA_U_BOTH.ps}
695: % \end{center}
696: % \caption{\footnotesize\it Checking that lattice Green functions satisfy the ghost SD equation (\ref{SD_L1_av2}).
697: % The l.h.s vs r.h.s of (\ref{SD_L1_av2}) is plotted at the left, and at the right we plot
698: % $\frac{1}{1.16}\widetilde{F}(p^2) - g_0 \frac{p_\mu}{N_c^2 -1} f^{abc}\langle A^c_\mu(0).
699: % \widetilde{F}_{\text{1conf}}^{(2)ba}(\mathcal{A},p)\rangle $ compared to $1$}
700: % \label{SDcheck}
701: % \end{figure}
702: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
703: % %
704: Lattice propagators were successfully checked by the perturbation theory at large
705: momentum, and they satisfy the ghost Schwinger-Dyson equation. This means that
706: lattice approach gives consistent results also in the infrared.
707:
708:
709: %
710: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
711: \subsection{Checking the validity of the tree-level approximation for the ghost-gluon vertex}
712: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
713: %
714:
715: The simplest approximation (used by many authors, see section~\ref{Review_of_today_s_analytical_results})
716: of the ghost Schwinger - Dyson equation~(\ref{SDghost_IR}) corresponds to the case
717: %
718: \begin{equation}
719: H_1(q,k)=1\quad \forall q,k,
720: \end{equation}
721: %
722: an approximation motivated by the non-renormalisation theorem (\ref{NonRenormalization})
723: valid for the sum $H_1(q,k)+H_2(q,k)$ when $k=0$. This gives
724: %
725: \begin{equation}
726: \label{trunc_SD_cont}
727: \frac{1}{F(k)} = 1 + \frac{g_0^2 N_c}{k^2} \int \frac{d^4 q}{(2\pi)^4}
728: \Big(\frac{F(q^2)G((k-q)^2)}{q^2 (k-q)^2}
729: \frac{(k\cdot q)^2 - k^2 q^2}{(q-k)^2}
730: \,\cdot 1 \, \Big).
731: \end{equation}
732: %
733: Strictly speaking this equation, written in this way, is meaningless since it involves
734: UV-divergent quantities. However it is well defined at fixed ultraviolet cut-off.
735:
736: We want to check whether lattice propagators satisfy it. According to perturbation theory,
737: it should be approximately true at large $k$. Lattice propagators are discrete functions of momentum and
738: thus one has to handle the problem of the numerical evaluation of the loop
739: integral $I$ in (\ref{trunc_SD_cont}). Let us express the integrand solely
740: in terms of $q^2$ and $(k-q)^2$
741: %
742: \begin{equation}
743: \label{I_is}
744: I =\int \frac{d^4 q}{(2\pi)^4} \,
745: \frac{F^2(q)G(k-q)}{q^2 (k-q)^2 }
746: \Big[ \frac{(k-q)^2}{4} + \frac{(k^2)^2+(q^2)^2 - 2k^2 q^2}{4(k-q)^2}
747: -\frac{q^2+k^2}{2} \Big].
748: \end{equation}
749: %
750: Then we write
751: \begin{equation}
752: I=I_1+I_2+I_3+I_4+I_5+I_6,
753: \end{equation}
754: %
755: each $I_{i}$ corresponds to one term in (\ref{I_is}). All these integrals have the form
756: %
757: \begin{equation}
758: I_i = C_i(k) \int \frac{d^4 q}{(2\pi)^4} f_i (q) h_i(k-q).
759: \end{equation}
760: %
761: The convolution in the r.h.s. is just the Fourier transform of the product at the same point
762: in configuration space:
763: %
764: \begin{equation}
765: \int \frac{d^4 q}{(2\pi)^4} \, f_i (q) h_i(k-q) = \text{F}_{+}
766: \Big( \text{F}_{-}(f_i)[x] \text{F}_{-}(h_i)[x] \Big)(k),
767: \end{equation}
768: %
769: where $\text{F}_{-}(\hat{f})(x)$ is an inverse and $\text{F}_{+}(f)(k)$ a direct Fourier transform.
770: Thus, in order to calculate the integral $I$ from discrete lattice propagators one proceeds as follows:
771: %
772: \begin{enumerate}
773: \item calculate $\{f_i\}(p)$ and $\{h_i\}(p)$ as functions of $F(p), G(p), p^2$ for all $i$
774: \item apply the inverse Fourier transform $\text{F}_{-}$ to all these functions and get $f_i(x)$ and $h_i(x)$
775: \item compute the product at the same point $f_i(x)\cdot h_i(x)$
776: \item apply the direct Fourier transform $\text{F}_{+}$ to $f_i(x)\cdot h_i(x)$
777: \end{enumerate}
778: %
779: The integrands in (\ref{I_is}) depend only on the squared norms $q^2$ and $(k-q)^2$,
780: and thus the angular part may be integrated out, giving the four-dimensional
781: Hankel transformation
782: %
783: \begin{align}
784: & \widehat{f(\arrowvert x \arrowvert)}[p]=
785: \frac{1}{\arrowvert p \arrowvert}
786: \int_{0}^{\infty}
787: J_1(\arrowvert p \arrowvert r)r^{2}
788: f(r)dr\nonumber
789: \\ &
790: f(r)=\frac{1}{(2\pi)^2}
791: \frac{1}{r}
792: \int_{0}^{\infty}
793: J_{1}(\rho r)\rho^{2}
794: \widehat{f(\arrowvert x \arrowvert)}[\rho] d\rho.
795: \end{align}
796: %
797: These integrals are evaluated numerically by means of the Riemann sum
798: %
799: \begin{equation}
800: \label{HT}
801: f(r) = (2\pi)^{-2} \arrowvert r \arrowvert^{-1}
802: \sum_{i=1}^{N} J_{1}(r \rho_i ) \rho_i^{2} \frac{\hat{f}[\rho_i] + \hat{f}[\rho_{i-1}]}{2} (\rho_i-\rho_{i-1}),\qquad \rho_0 = 0,
803: \end{equation}
804: %
805: where $N$ is the number of data points. The inverse transformation is
806: done in the similar way. In practice, because of the lattice artifacts
807: (see subsection~\ref{subsection_discretisation_errors}) which become
808: important at large $\rho$ the summation has to be
809: restricted to $\rho < \rho_{max} \simeq 2.2$ instead of the maximal value $ 2 \pi$.
810:
811: Now we are ready to check the approximate equation~(\ref{trunc_SD_cont}) on the lattice.
812: We still have to face the same problem we have already encountered in the previous subsection,
813: namely that the lattice Faddeev-Popov operator involves the non trivial discretisation $\Delta_U$
814: of the Laplacian operator. This is taken into account by means of the substitution of $\widetilde{\Delta_U}(p^2)/p^2$ to the ``1'' term in the l.h.s of equation~(\ref{trunc_SD_cont})
815: We present on Figure~\ref{figure_SDcheckH1} the result of the numerical integration
816: described above. We have chosen for this purpose the data set from the simulation
817: with the gauge group $SU(3)$ at $\beta=6.4, V=32^4, a^{-1}\approx 3.6\text{ GeV}$.
818: One sees that the equality is achieved at large momenta, but in the infrared
819: the naive approximation of the ghost Schwinger-Dyson equation fails.
820: %
821: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
822: \begin{figure}[!htb]
823: \begin{center}
824: \includegraphics[angle=0, width=0.8\linewidth]{EPS/truncSD2}
825: \end{center}
826: \caption{\footnotesize\it Checking whether lattice Green functions satisfy the ghost SD equation (\ref{SDghost_IR}) with
827: an assumption $H_1(q,k)=1$. The upper line(circles) correspond to the loop integral in (\ref{SDghost_IR}),
828: and the down line (triangles) corresponds to $1/F(p^2)-1$. In this plot $a^{-1}\approx 3.6$ GeV. }
829: \label{figure_SDcheckH1}
830: \end{figure}
831: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
832: %
833: The errors on Figure~\ref{figure_SDcheckH1} include statistical Monte-Carlo errors
834: for $F(q^2)$ and $G(q^2)$ and the bias coming from the UV cut-off of the integral $I$.
835:
836: We see that at small momenta (below $\approx3$ GeV) the ghost Schwinger-Dyson equation
837: with the assumption $H_1(q,k)=1$ is not satisfied. However, it is quite difficult to establish
838: whether this disagreement is due to the infrared or ultraviolet dependencies of
839: $H_1(q,k)$. To check this one has to know $H_1(q,k)$ for all values of $q,k$.
840: Unfortunately this information is not available. Thus the main conclusion of the
841: present subsection is that the scalar function $H_1(q,k)$ plays an important
842: role in the infrared gluodynamics, and it cannot be set to one.
843:
844: %
845: %
846: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
847: \section{Direct fits of infrared exponents}
848: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
849: %
850: %
851: \label{section_direct_fits}
852:
853: We have seen in the previous section that lattice simulations give consistent
854: results for the Green functions at all momenta. Another interesting feature
855: that we have established is the important role
856: of the scalar factor $H_1(q,k)$ coming from the complete ghost-gluon
857: vertex (\ref{H12_definition}). In this section we present numerical
858: results allowing to check the relation (\ref{R}). After this we
859: present our results for direct fits of the exponents $\alpha_F$ and $\alpha_G$,
860: and compare them to the results of other lattice collaborations.
861: %
862: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
863: \begin{figure}[!thb]
864: \vspace*{0.02\paperheight}
865: \centering
866: \hspace*{-0.1cm}
867: \begin{tabular}{lr}
868: \includegraphics[width=0.45\linewidth]{EPS/SU3_F2G}&\includegraphics[width=0.45\linewidth]{EPS/SU2_F2G}
869: \end{tabular}
870: \caption{\footnotesize\it Direct test of the relation $2\alpha_F + \alpha_G=0$. If the last is true $F^2G$ has to be
871: constant in the infrared. We see that it is clearly not the case. In these plots $a^{-1}\approx 1.2$ GeV, so the peak is located at $\approx 600$ MeV.}
872: \label{F2G}
873: \end{figure}
874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
875: %
876:
877:
878: %
879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
880: \subsection{Testing the relation $2\alpha_F + \alpha_G =0$.}
881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
882: %
883:
884: In order to test the relation (\ref{R}) we plot at Figure~\ref{F2G} the quantity $F^2(p^2) G(p^2)$.
885: If all the conditions (\ref{conditions}) are satisfied this quantity should be constant
886: in the infrared (or slightly varying). We see from Figure~\ref{F2G} that in the infrared (below $\approx 600$ MeV)
887: the quantity $F^2 G$ is not constant, and thus one of the conditions (\ref{conditions})
888: is not verified. We have seen that the conditions $\alpha_F \neq 0 $ and $\alpha_F+\alpha_G < 1$
889: are consistent with the limits (\ref{STpredictions}) from the Slavnov-Taylor
890: identity (\ref{ST}). We have also seen (cf. Figures~\ref{SDcheck} and \ref{figure_SDcheckH1})
891: that neglecting the momentum dependence of the vertex is not possible in the infrared, because in this case
892: the ghost Schwinger-Dyson equation is no longer satisfied by lattice propagators.
893: Thus the only possibility is to admit that $H_1(q,k)$ plays an important role,
894: and that the relation (\ref{R}) is not verified. If $\alpha_F\neq 0$ then the
895: modified form (\ref{R_Gamma}) that takes in account the singularity of $H_1(q,k)$ should be
896: considered (according to Fig.\ref{figure_SDcheckH1}), with $\alpha_\Gamma < 0$
897: in our parametrisation. This singularity is probably related to the non-perturbative
898: power corrections to the vertex discussed in the subsection~\ref{subsection_on_Wilson_coeff_from_ST}.
899: %
900: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
901: \begin{figure}[!h]
902: \vspace*{0.02\paperheight}
903: \begin{center}
904: \includegraphics[width=0.6\linewidth]{EPS/G2_SU2_48_32}
905: % \begin{tabular}{lcr}
906: % \includegraphics[width=0.45\linewidth]{EPS/G2_SU2_48_32}&&\includegraphics[width=0.45\linewidth]{EPS/G2_SU3}
907: % \end{tabular}
908: \end{center}
909: \caption{\footnotesize\it $G^{(2)}(p^2)$ from lattice simulation for $SU(2)$ (left).
910: $\beta_{\text{SU(2)}}=2.3$ and $\beta_{\text{SU(3)}}=5.75$.
911: The volumes are $32^4$ and $48^4$ for $SU(2)$.
912: In these plots $a^{-1}\approx 1.2$ GeV.}
913: \label{plotSU2}
914: \end{figure}
915: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
916: %
917:
918: Another reason to think that the relation (\ref{R}) is not exact
919: is the dependence of $\alpha_F$ and $\alpha_G$ on the
920: choice of the Gribov copy. We have seen in the section~\ref{section_Gribov_copies_sur_reseau}
921: that the low-momentum dependence of the gluon propagator
922: is not sensitive to the bc/fc choice while the infrared behaviour
923: of the ghost propagator depends on it. But the Schwinger - Dyson equation
924: for the ghost propagator is independent of the choice of the copy, because it is valid
925: exactly for every gauge configuration, even on a finite lattice
926: (see equations (\ref{SD_simple_end}) and (\ref{SD_L1_av})). Hence if there is a relation
927: between the infrared exponents $\alpha_F$ and $\alpha_G$
928: resulting from the ghost Schwinger - Dyson equation then
929: it could not depend on the choice of the copy. Thus it is
930: not possible to have a relation with $\alpha_F$ and $\alpha_G$ alone.
931: This above argument is not directly applicable in the case $\alpha_F = 0$.
932:
933: According to the analysis performed in the section~\ref{section_Relation_between_the_infrared_exponents},
934: and given (see next subsection) that the case $2$ of the Table~\ref{tabledescas} is excluded by lattice simulations
935: the following explanations of the non-validity of the relation $2\alpha_F+\alpha_G=0$ are possible:
936: %
937: \begin{enumerate}
938: \item The ghost-gluon vertex contains scalar factors that are singular in the infrared, i.e. $\alpha_\Gamma\neq 0$ in
939: the equation \ref{parametrization_H1}.
940:
941: \item The case $4$ of the Table~\ref{tabledescas} is realised~\cite{Boucaud:2006if} and hence there exists \emph{no} relation between the infrared exponents.
942: Let us recall that in the above case one has $\alpha_F=0$ and $\alpha_G+\alpha_\Gamma\geq 1$. If the ghost-gluon vertex is regular in the
943: infrared then one has
944: %
945: \begin{equation}
946: \label{case_4_vertex_regulier}
947: \left\{
948: \begin{array}{l}
949: \alpha_F=0 \\
950: \alpha_G \geq 1.
951: \end{array}
952: \right.
953: \end{equation}
954: %
955: \end{enumerate}
956: %
957:
958: %
959: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
960: \subsection{Lattice fits for $\alpha_F$ and $\alpha_G$.}
961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
962: %
963:
964: Let us now discuss the direct fits for the infrared exponents $\alpha_G$ and $\alpha_F$.
965: The examples of such fits of lattice data are presented on Figure~\ref{plotSU2}.
966: The errors are quite large, leading to an instability in the fit results.
967: That is why we fit both propagators in the infrared to the formula
968: %
969: \begin{equation}
970: (q^2)^\alpha (\lambda + \mu q^2)
971: \end{equation}
972: %
973: where we added an additional term of the form $\mu q^2$ in order to describe a situation
974: like the one at Figure~\ref{plotSU2}(left) where $G^{(2)}(p^2)$ seems
975: to go to a finite limit when $p$ goes to zero.
976: The obtained values for $\alpha_{F,G}$ are summarised
977: in Table~\ref{tablalpha}.
978: %
979: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
980: \begin{table}[h]
981: \begin{center}
982: \begin{tabular}{c|c|c|cc}
983: \hline\hline
984: \text{Group}&\text{Volume}&$\beta$&$\alpha_G$&$\alpha_F$\\
985: \hline
986: $SU(2)$ & $48^4$ & $2.3$ & $1.004(15)$ &$-0.087(15)$\\
987: \hdashline[0.4pt/1pt]
988: $SU(2)$ & $32^4$ & $2.3$ & $0.968(11)$ &$-0.109(14)$\\
989: %\hline
990: %$SU(3)$ & $32^4$ & $5.75$ & $0.864(16)$ &$-0.153(22)$\\
991: \hline\hline
992: \end{tabular}
993: \caption{Summary of the fit results for the $F$ and $G$ functions}
994: \label{tablalpha}
995: \end{center}
996: \label{default}
997: \end{table}%
998: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
999: %
1000: For $SU(2)$ and the larger lattice volume the value obtained for $\alpha_G$ is compatible
1001: with 1.% The situation is less clear for $SU(3)$, but in this case data with the
1002: %larger volume ($48^4$) are lacking.
1003: We also take into account our experience from previous studies of the
1004: gluon propagator where we have always observed that the gluon
1005: propagator goes continuously to a finite limit in the infrared region (see Figure~\ref{GdPAll}).
1006: %
1007: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1008: \begin{figure}[!h]
1009: %\vspace*{0.01\paperheight}
1010: \begin{center}
1011: \includegraphics[ angle=-90 ,width=0.75\linewidth]{EPS/GdpAll}
1012: \end{center}
1013: \caption{\footnotesize\it The continuity of the lattice gluon propagator in the infrared.}
1014: \label{GdPAll}
1015: \end{figure}
1016: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1017: %
1018: However, the fits are quite instable, and depend a lot on the choice of the fit formula
1019: that can considerably change the result. The main problem is the lack of data points at low momenta.
1020:
1021: Regarding the gluon propagator another strategy may be taken. It consist in
1022: extrapolating the available data to the infinite volume limit.
1023: A very detailed study of the gluon dressing function and specially of
1024: its volume dependence at $k=0$ has already been performed in~\cite{Bonnet:2001uh}.
1025: This study shows that a value $\alpha_G = 1$ is compatible with the
1026: data (the dressing function shows no signal of discontinuity in the neighbourhood
1027: of zero) and that no pathology shows up as the volume goes to infinity.
1028: Let us compare all available lattice results for the point $G^{(2)}(0)$
1029: and check whether there is an agreement between the data.
1030: Following~\cite{Bonnet:2001uh} we renormalise the gluon propagator in the MOM scheme
1031: at $4~\text{GeV}$ and use the suggested extrapolation formula
1032: %
1033: \begin{equation}
1034: G_R^{(2)}(0,\mu=4 \,{\rm GeV}) = G_{R\,\infty}^{(2)}(0,\mu=4 \,{\rm GeV})
1035: + \frac{c}{V}.
1036: \end{equation}
1037: %
1038: We compare the results of~\cite{Bonnet:2001uh},\cite{Oliveira:2005hg} and our data from the
1039: Table~\ref{Gzero_OLD_data}. The results for the fit parameters $G_{R\,\infty}^{(2)}(0)$ and $c$
1040: are presented in the Table~\ref{Gde0}.
1041: %
1042: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%55
1043: \begin{table}[h]
1044: \centering
1045: \begin{tabular}{c;{0.4pt/1pt}c;{0.4pt/1pt}c;{0.4pt/1pt}c}
1046: \hline\hline
1047: $\beta$ & $V$ \text{in units of} $a$
1048: &\text{bare propagator} $G^{(2)}(p)$ & $1/aL$ \text{in GeV}
1049: \\ \hline
1050: 5.7 &$16^4 $&$16.81\pm0.13$ & 0.0672 \\
1051: 5.7 &$24^4 $ &$15.06\pm0.29$ & 0.0448 \\
1052: 5.8 &$16^4 $ &$19.12\pm0.16$ & 0.0841 \\
1053: 5.9 &$24^4 $ &$18.12\pm0.30$ & 0.0685 \\
1054: 6.0 &$32^4 $ &$17.70\pm0.59$ &0.0615 \\
1055: 6.0 &$24^4$ &$19.67\pm0.35$ &0.0821 \\
1056: \hline\hline
1057: \end{tabular}
1058: \caption{\footnotesize\it Physical lattice sizes and raw data for the gluon propagator at zero momentum
1059: $G^{(2)}(p)$ from our old data.}
1060: \label{Gzero_OLD_data}
1061: \end{table}
1062: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1063: %
1064: %
1065: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1066: \begin{table}[h]
1067: \centering
1068: \begin{tabular}{c|c|c|c}
1069: \hline\hline
1070: \text{reference} &
1071: $G_R^{(2)}(0,\mu=4 \,{\text{GeV}})$ \text{in GeV}$^{-2}$ & $c$ \text{in GeV}$^{-2}$\text{ fm}$^4$
1072: & \text{max vol in fm}$^4$
1073: \\ \hline \cite{Bonnet:2001uh} &
1074: $7.95\pm 0.13$ & $245\pm 22$ & 2000
1075: \\ Table \ref{Gzero_OLD_data} &
1076: $9.1\pm 0.3$ & $140\pm 50 $ & 90
1077: \\ \cite{Oliveira:2005hg} &
1078: $10.9-11.3$ & $47-65$ & 110s
1079: \\ \hline\hline
1080: \end{tabular}
1081: \caption{\footnotesize\it Summary of the infinite volume zero momentum
1082: propagator and its slope in terms of $1/V$ for three different simulations.
1083: The largest volume used in the fit is also indicated. The statistical error is not
1084: quoted in~\cite{Oliveira:2005hg}.}
1085: \label{Gde0}
1086: \end{table}
1087: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1088: %
1089: We are aware that not all systematic errors are taken into account:
1090: $O(a)$ effects, effect due to different lattice shape, insufficiently large volumes
1091: (for the second and third lines), uncertainty in the estimate of
1092: the lattice spacing in physical units, etc. However, it seems that not only
1093: there is a clear indication in favour of a finite non vanishing zero momentum
1094: gluon propagator, but that different lattice collaborations agree on the value.
1095: Of course a more extensive study is necessary to check this statement.
1096: The other free parameter of the fit - the slope $c$ - is clearly different, but
1097: still all the values are in agreement in the order of magnitude.
1098:
1099: We conclude~\cite{Boucaud:2006pc} thus that all available numerical results point towards a
1100: finite non-vanishing and zero momentum renormalised lattice gluon propagator in
1101: the infinite volume limit. This suggests that $\alpha_G = 1$.
1102: An additional study at much larger lattices is needed to get a reliable
1103: result for this infrared exponent.
1104:
1105: This last result is in conflict with the limits found from the study of the
1106: Slavnov - Taylor identity (\ref{STpredictions}), and contradicts the Zwanzigers's
1107: prediction that the gluon propagator is infrared suppressed. However, it is very close to
1108: the results presented in~\cite{Aguilar:2004sw}.
1109:
1110: Let us finally discuss the gauge-dependence of the parameter $\alpha_G$. This is still an open
1111: question. However, the results of works~\cite{Giusti:2000yc},\cite{Giusti:2001kr} suggest
1112: that the value of $\alpha_G$ does not change drastically (see Figure~\ref{gluon_gauge_xi}) when
1113: changing the gauge parameter $\xi$. This question deserves a separate study.
1114: %
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: \begin{figure}[!h]
1117: \begin{center}
1118: \includegraphics[width=0.55\linewidth]{EPS/ghost_IR_log_fit}
1119: \end{center}
1120: \caption{\footnotesize\it The fit of the ghost scalar factor $F(p)$ to the formula $A+B\log{p}$. It suggests that
1121: the infrared divergence of $F(p)$ is very slow. Hence $\alpha_F$ is close to zero~\cite{Boucaud:2006if}. The simulations was performed on
1122: a $V=32^4$ lattice at $\beta=5.8$. The Landau gauge was fixed using the f.c. choice for the Gribov copies.}
1123: \label{ghost_log_fit}
1124: \end{figure}
1125: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1126: %
1127:
1128: To finish this chapter, let us summarise the lattice results. We have found
1129: that $\alpha_F$ is very close to 0 (see Figure~\ref{ghost_log_fit}), $\alpha_G$ is close to 1 and the widely
1130: used relation $2\alpha_F + \alpha_G = 0$ is not true. Going back to the possibilities
1131: given in the Table~\ref{tabledescas} we find that the cases $2$ and $3$ are not realised.
1132: We are left with the cases $1$ and $4$, and for the moment lattice simulations cannot
1133: say which possibility is true. However, all numerical results are better explained
1134: by the possibility (\ref{case_4_vertex_regulier}) corresponding to the
1135: case $4$ of the Table~\ref{tabledescas} supplied with an hypothesis of the regularity of the
1136: scalar factors entering the ghost-gluon vertex. We recall that in this case one has:
1137: %
1138: \begin{equation}
1139: \left\{
1140: \begin{array}{l}
1141: \alpha_F=0 \\
1142: \alpha_G \geq 1. \\
1143: % \text{no relation between $\alpha_F$ and $\alpha_G$}
1144: % \\
1145: % \text{follows from the ghost Schwinger - Dyson equation}.
1146: \text{no relation between $\alpha_F$ and $\alpha_G$ follows from the ghost SD equation}.
1147: \end{array}
1148: \right.
1149: \end{equation}
1150: %
1151: %
1152: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1153: \begin{figure}[!h]
1154: %\vspace*{0.01\paperheight}
1155: \begin{center}
1156: \includegraphics[width=0.75\linewidth]{EPS/fit04}
1157: \end{center}
1158: \caption{\footnotesize\it Transverse part of the gluon propagator $p^2 G^{(2)}(p)$ in
1159: covariant gauges as a function of $p$. The two sets of data refer to ($\xi=\lambda$) $\xi=0$ (Landau gauge) and $\xi=8$,
1160: 221 thermalized $SU(3)$ configurations at $\beta=6.0$ with a volume $V\times T=16^3\times 32$.
1161: Extracted from~\cite{Giusti:2000yc}.}
1162: \label{gluon_gauge_xi}
1163: \end{figure}
1164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1165: %
1166: Note that it is still in conflict with the constraints coming from the Slavnov - Taylor identity (\ref{STpredictions}).
1167: Thus the essential question today is to understand whether $\alpha_F=0$ or not~\cite{Boucaud:2006if}. And, of course, a study on
1168: larger lattices is necessary to perform better fits of the infrared exponents.