1: \documentstyle[eqsecnum,psfig,aps,12pt]{revtex}
2: \setlength{\evensidemargin}{-0.0cm}
3: \setlength{\oddsidemargin}{-0.0cm}
4: \setlength{\topmargin}{-1.8cm}
5: \setlength{\baselineskip}{20pt}
6: \setlength{\textwidth}{16.4cm}
7: \setlength{\textheight}{22.5cm}
8: \renewcommand{\baselinestretch}{1.1}
9: %\renewcommand{\baselinestretch}{1.0}
10:
11: %%%%%%%%%%%% Definitionen %%%%%%%%%%
12:
13: \def\beq{\begin{equation}}
14: \def\eeq{\end{equation}}
15:
16: \def\bea{\arraycolsep .1em \begin{eqnarray}}
17: \def\eea{\end{eqnarray}}
18: \def\Tr{{\rm Tr}}
19: \def\step{\\[-1.5ex]}
20:
21: \def\dq{{q \llap{/}}\,}
22: \def\dk{{k \llap{/}}\,}
23:
24: \def\de{\delta}
25: \def\Ga{\Gamma}
26: \def\eps{\epsilon}
27: \def\om{\omega}
28:
29: \def\bpi{\mbox{\boldmath$\pi$}}
30: \def\bphi{\mbox{\boldmath$\phi$}}
31:
32: \def\eq#1{(\ref{#1})}
33: \def\Es#1{Eqs.~(\ref{#1})}
34: \def\Eq#1{Eq.~(\ref{#1})}
35: \def\fig#1{fig.~(\ref{#1})}
36:
37: \def\s0#1#2{\mbox{\small{$ \frac{#1}{#2} $}}}
38: \def\0#1#2{\frac{#1}{#2}}
39:
40: \def\llangle{\left\langle}
41: \def\rrangle{\right\rangle}
42:
43: %%%%%%%%%%%% Ende Definitionen %%%%%%%%%%
44:
45:
46: \makeatletter
47: %\renewcommand\section{\@startsection{section}{1}{\z@}%
48: % {-3.25ex\@plus -1ex \@minus -.2ex}%
49: % {1.5ex \@plus .2ex}%
50: % {\normalfont\normalsize\bfseries}}
51: %\renewcommand\subsection{\@startsection{subsection}{2}{\z@}%
52: % {-3.25ex\@plus -1ex \@minus -.2ex}%
53: % {1.5ex \@plus .2ex}%
54: % {\normalfont\normalsize\itshape}}
55:
56: \renewenvironment{thebibliography}[1]
57: {\section*{References}\frenchspacing\small
58: \begin{list}{[\arabic{enumi}]}
59: {\usecounter{enumi}\parsep=2pt\topsep 0pt
60: \settowidth{\labelwidth}{[#1]}
61: \leftmargin=\labelwidth\advance\leftmargin\labelsep
62: \rightmargin=0pt\itemsep=0pt\sloppy}}{\end{list}}
63:
64: \makeatother
65:
66:
67:
68:
69: \begin{document}
70: \begin{center}
71:
72: \thispagestyle{empty}
73:
74: {\normalsize\begin{flushright}CERN-TH-2002-023 \\[12ex]
75: \end{flushright}}
76:
77:
78: \mbox{\large \bf Critical exponents from optimised
79: renormalisation group flows}\\[6ex]
80:
81: {Daniel F. Litim}
82: \\[2ex]
83: {\it Theory Division, CERN, CH -- 1211 Geneva 23.}
84: \\
85: {\footnotesize (Daniel.Litim@cern.ch)}
86: \\[10ex]
87:
88: {\small \bf Abstract}\\[2ex]
89: \begin{minipage}{14cm}{\small
90: Within the exact renormalisation group, the scaling solutions for
91: O($N$) symmetric scalar field theories are studied to leading
92: order in the derivative expansion. The Gaussian fixed point is
93: examined for $d>2$ dimensions and arbitrary infrared
94: regularisation. The Wilson-Fisher fixed point in $d=3$ is studied
95: using an optimised flow. We compute critical exponents and
96: subleading corrections-to-scaling to high accuracy from the
97: eigenvalues of the stability matrix at criticality for all $N$.
98: We establish that the optimisation is responsible for the rapid
99: convergence of the flow and polynomial truncations thereof. The
100: scheme dependence of the leading critical exponent is analysed.
101: For all $N\ge 0$, it is found that the leading exponent is
102: bounded. The upper boundary is achieved for a Callan-Symanzik
103: flow and corresponds, for all $N$, to the large-$N$ limit. The
104: lower boundary is achieved by the optimised flow and is closest to
105: the physical value. We show the reliability of polynomial
106: approximations, even to low orders, if they are accompanied by an
107: appropriate choice for the regulator. Possible applications to
108: other theories are outlined.}
109: \end{minipage}
110: \end{center}
111:
112: \newpage
113: \pagestyle{plain}
114: \setcounter{page}{1}
115: \noindent
116: %********|*********|*********|*********|*********|*********|*********|****
117: \section{Introduction}\label{Introduction}
118: %********|*********|*********|*********|*********|*********|*********|****
119:
120:
121: Renormalisation group techniques are important tools to describe how
122: classical physics is modified by quantum fluctuations. Integrating-out
123: all quantum fluctuations provides the link between the classical
124: theory and the full quantum effective theory. Universality implies
125: that the details of the underlying classical theory -- other than the
126: global symmetries, long- or short-range interactions, and the
127: dimensionality -- are irrelevant for the characteristics of the quantum
128: effective theory. For this reason, universal properties of phase
129: transitions in numerous physical systems (entangled polymers,
130: liquid-vapour transition, superfluid transition in ${}^4$He,
131: ferromagnetic transitions, QCD phase transition with two massless
132: quark flavours) can be addressed based on simple scalar field theories
133: \cite{Zinn-Justin:1989mi}. \step
134:
135: A useful method is given by the Exact Renormalisation Group (ERG)
136: \cite{Polchinski,continuum,Ellwanger:1994mw,Morris:1994qb,Bagnuls:2000ae},
137: which is based on the Wilsonian idea of integrating-out infinitesimal
138: momentum shells. The corresponding flow, which has a simple one-loop
139: structure, is very flexible concerning approximations, and its domain
140: of applicability is not tied to weak coupling. Recently, it has been
141: shown that ERG flows can be optimised, thereby providing improved
142: results already to low orders within a given approximation
143: \cite{Litim:2000ci,Litim:2001up,Litim:2001fd,Litim:2001dt}. In the
144: present paper, we apply this idea to the universality class of $O(N)$
145: symmetric scalar theories in three dimensions and compute critical
146: exponents and subleading corrections to scaling. We expect that
147: insights gained from this investigation will also prove useful for
148: applications to gauge theories \cite{Litim:1998nf} or gravity
149: \cite{QuantumGravity1}, which are more difficult to handle.\step
150:
151: Universal critical exponents have been computed previously using
152: either polynomial truncations of exact renormalisation group flows, or
153: the derivative expansion to leading and subleading order
154: \cite{Litim:2001dt,Hasenfratz:1986dm,Margaritis:1988hv,Tetradis:1994ts,Morris:1994ie,Ball:1995ji,Litim:1995ex,Tetradis:1996br,Comellas:1997tf,Morris:1998xj,Liao:2000sh,VonGersdorff:2000kp,Litim:2001hk},
155: and in \cite{Litim:2001hk,Bohr:2001gp} based on the proper-time
156: renormalisation group \cite{Liao:1996fp}. All results are affected by
157: the underlying approximations which induce a spurious dependence on
158: the regularisation
159: \cite{Litim:2001dt,Ball:1995ji,Liao:2000sh,Litim:2001ky,Litim:1997nw,Freire:2001sx}.
160: This is somewhat similar to the scheme dependence within perturbative
161: QCD, or within truncated solutions of Schwinger-Dyson equations. While
162: this scheme dependence should vanish at sufficiently high order in the
163: expansion, practical applications are always bound to a finite order,
164: and hence to a non-vanishing scheme dependence. In some cases, it has
165: even been observed that higher order results happen to be worse than
166: lower order ones \cite{Morris:1998xj}. In consequence, one should gain
167: some understanding of the spurious scheme dependence. Without this,
168: it is difficult to decide which of the different scheme-dependent
169: results within a fixed truncation could be considered as trustworthy.
170: \step
171:
172: A partial understanding of the interplay of approximations and scheme
173: dependence has been achieved previously. For scalar QED
174: \cite{ScalarQED}, the scheme dependence in the region of first order
175: phase transition has been studied in
176: \cite{Litim:1997nw,Freire:2001sx}. For $3d$ scalar theories, the
177: interplay between the smoothness of the regulator and the resulting
178: critical exponents has been addressed in \cite{Liao:2000sh} using a
179: minimum sensitivity condition. For Einstein quantum gravity, where a
180: new UV fixed point has been found recently to low orders in a
181: polynomial truncation, the corresponding analysis has been given in
182: \cite{QuantumGravity2}. The weak scheme dependence found in these
183: cases suggests that higher order corrections remain small, thereby
184: strengthening the results existing so far. The evidence created this
185: way is partly circumstantial, because the range over which a physical
186: observable varies with the scheme depends on the class of regulators.
187: \step
188:
189: Here, we address the problem from a different perspective. The main
190: ingredient in our analysis is the concept of optimisation
191: \cite{Litim:2000ci,Litim:2001up,Litim:2001fd,Litim:2001dt}. For a
192: given physical problem, it should be possible to identify specific
193: regulators which lead to a better convergence behaviour of the flow.
194: This strategy is based only on the ERG flow itself and has lead to a
195: simple optimisation criterion for flows \cite{Litim:2000ci}. Optimised
196: flows have a number of interesting properties \cite{Litim:2001up}.
197: They lead to a fast decoupling of heavy modes, they disentangle
198: quantum and thermal fluctuations along the flow, and they lead to a
199: smooth approach towards a convex effective potential for theories in a
200: phase with spontaneous symmetry breaking. The optimisation is closely
201: linked to a minimum sensitivity condition \cite{Litim:2001fd} in a
202: sense which will be made transparent below. Furthermore, optimised
203: flows have been shown to improve the convergence of the derivative
204: expansion \cite{Litim:2001dt}. Thus, optimised flows are promising
205: candidates for high precision computations within this formalism.
206: Here, we do so within a local potential approximation. \step
207:
208: The second new ingredient of our analysis consists in a study of the
209: largest possible range of flows, and the corresponding critical
210: exponents. We find that the range is larger than
211: previously assumed. Furthermore, the results from optimised flows are
212: located at the (lower) boundary and happen to be closest to the
213: physical values. \step
214:
215: For the numerical analysis, and apart from the local potential
216: approximation, we employ a polynomial approximation. This additional
217: approximation is reliable if it converges reasonably fast towards the
218: full solution. However, it has been criticised previously in the
219: literature. For a sharp cut-off, it has lead to spurious solutions
220: \cite{Margaritis:1988hv}, and its convergence was found to be poor
221: \cite{Morris:1994ki}, which has lead to strong doubts concerning its
222: reliability (see also \cite{Aoki:1998um}). In contrast to these
223: findings, we show that the poor convergence is an artifact of the
224: sharp cut-off regulator, rather than an artifact of the polynomial
225: approximation. Using an optimised flow, we find that the polynomial
226: approximation is stable and that it converges rapidly for all
227: technical purposes. \step
228:
229:
230: The format of the paper is as follows. We review the basic ingredients
231: of the formalism and introduce the optimisation ideas
232: (Sect.~\ref{RGFlow}). Then, we introduce our numerical method and
233: study the non-trivial scaling solution in $3d$
234: (Sect.~\ref{FixedPoints}). We compute the eigenvalues at criticality
235: to high accuracy from an optimised flow. The convergence and stability
236: of the flow, and of the polynomial truncation, are established
237: (Sect.~\ref{CriticalExponents}). We study the scheme dependence of the
238: critical index $\nu$ (Sect.~\ref{Bounds}). Finally, we discuss the
239: main results of the paper with particular emphasis on the predictive
240: power, on the convergence properties of flows, and on implications for
241: other theories (Sect.~\ref{Discussion}). Three appendices contain the
242: study of the Gaussian fixed point for arbitrary regularisation and
243: $d>2$ (App.~\ref{Gauss}), and technical details for specific classes
244: of regulators, including a computation of the corresponding flows and
245: critical exponents (Apps.~\ref{VariantsOpt} and~\ref{SlidingOpt}).
246:
247: %********|*********|*********|*********|*********|*********|*********|****
248: \section{RG flow for $O(N)$ symmetric scalar theories}\label{RGFlow}
249: %********|*********|*********|*********|*********|*********|*********|****
250:
251: Here, we briefly review some basic ingredients of the ERG formalism,
252: and its approximation to leading order in the derivative expansion. We
253: also discuss important aspects of the regularisation, and its
254: optimisation, which is employed in the following sections.
255:
256:
257: %********|*********|*********|*********|*********|*********|*********|****
258: \subsection{Renormalisation group flows}
259: %********|*********|*********|*********|*********|*********|*********|****
260:
261: Exact renormalisation group equations are based on the Wilsonian idea
262: of integrating out momentum modes within a path integral
263: representation of quantum field theory \cite{Polchinski}. In its
264: modern form, the ERG flow for an effective action $\Gamma_k$ for
265: bosonic fields $\phi$ is given by the simple one-loop expression
266: \cite{continuum,Ellwanger:1994mw,Morris:1994qb,Bagnuls:2000ae}
267: %
268: \beq \label{ERG}
269: \partial_t\Gamma_k[\phi] =
270: \frac{1}{2}\Tr\left( \frac{\de^2\Gamma_k}{\delta \phi\,\delta\phi}
271: + R_k\right)^{-1} \partial_t R_k
272: \eeq
273: %
274: Here, $t\equiv\ln k$ is the logarithmic scale parameter,
275: and $R_k(q^2)$ is an infrared (IR) regulator at the momentum scale
276: $k$. From now on, we suppress the index $k$ on $R$. The flow
277: trajectory of \Eq{ERG} in the space of action functional depends on
278: the IR regulator function $R$. $R$ obeys a few restrictions, which
279: ensure that the flow equation is well-defined, thereby interpolating
280: between an initial action in the UV and the full quantum effective
281: action in the IR. We require that
282: %
283: \bea
284: \label{I} \lim_{q^2/k^2\to 0}R(q^2) > 0 \,, \\
285: \label{II} \lim_{k^2/q^2\to 0}R(q^2) \to 0 \,, \\
286: \label{III} \lim_{k\to\Lambda}R(q^2) \to \infty\,.
287: \eea
288: %
289: Equation \eq{I} ensures that the effective propagator at vanishing
290: field remains finite in the infrared limit $q^2\to 0$, and no infrared
291: divergences are encountered in the presence of massless modes.
292: Equation \eq{II} guarantees that the regulator function is removed in
293: the physical limit, where $\Ga\equiv\lim_{k\to 0}\Ga_k$. Equation
294: \eq{III} ensures that $\Ga_k$ approaches the microscopic action
295: $S=\lim_{k\to \Lambda}\Ga_k$ in the UV limit $k\to \Lambda$. We put
296: $\Lambda=\infty$ in the sequel. For later use, we introduce a
297: dimensionless regulator function $r(y)$ as
298: %
299: \beq\label{r} R(q^2) = q^2\, r(q^2/k^2)\,. \eeq
300: %
301: Now we turn to the flow equation for an $O(N)$ symmetric scalar field
302: theory in $d$ dimensions to leading order in the derivative expansion,
303: the so-called local potential approximation \cite{Golner:1986}. This
304: approximation amounts to the Ansatz
305: %
306: \beq\label{AnsatzGamma}
307: \Gamma_k=\int d^dx \left(U_k(\bar\rho)
308: + \012 \partial_\mu \phi^a\partial_\mu \phi_a
309: \right)\ .
310: \eeq
311: %
312: for the functional $\Gamma_k$. The Ansatz neglects higher order
313: corrections proportional to the anomalous dimension $\eta$ of the
314: fields. The latter are of the order of a few percent for the
315: physically interesting universality classes $N=0,\cdots, 4$. Hence, we
316: expect that a derivative expansion is sensible, and that the result of
317: a leading order computation is correct up to corrections of the order
318: of $\eta$. Inserting this Ansatz into \eq{ERG} and evaluating it for
319: constant fields leads to the flow for $U_k$. We rewrite this flow
320: equation in dimensionless variables $u(\rho)=U_k/k^d$ and $\rho=\s012
321: \phi^a\phi_a k^{2-d}$. In addition, the angular integration of the
322: momentum trace is performed to give \cite{Wetterich:1993be}
323: %
324: \beq\label{FlowPotential}
325: \partial_t u+du-(d-2) \rho u'
326: =2v_d(N-1)\ell(u')+2v_d \ell(2\rho u'')\,
327: \eeq
328: %
329: with $v^{-1}_d=2^{d+1}\pi^{d/2}\Gamma(\s0d2)$. The function
330: $\ell(\omega)$ are given by
331: %
332: \beq\label{Id}
333: \ell(\omega)
334: =\s012\int^\infty_0dy y^{d/2} \0{\partial_t r(y)}{y(1+r)+\omega}\,
335: \eeq
336: %
337: with $y\equiv q^2/k^2$ and $\partial_t r(y) = -2 y r'(y)$. The flow
338: \eq{FlowPotential} is a second order non-linear partial differential
339: equation. All non-trivial information regarding the renormalisation
340: flow and the regularisation scheme (RS) are encoded in the function
341: \eq{Id}. The momentum integration is peaked and regularised: for large
342: momenta due to the regulator term $\partial_t r(y)$, and for small
343: momenta due to $r(y)$ in the numerator. All terms on the left-hand
344: side of \Eq{FlowPotential} do not depend explicitly on the RS. They
345: simply display the intrinsic scaling of the variables which we have
346: chosen for our parametrisation of the flow.
347:
348:
349:
350:
351: %********|*********|*********|*********|*********|*********|*********|****
352: \subsection{Optimisation}\label{Optimisation}
353: %********|*********|*********|*********|*********|*********|*********|****
354:
355: A good choice for the regulator is most important for a rapid
356: convergence and the stability of an approximated flow towards the
357: physical theory. Recently, is has been argued that such (optimised)
358: choices of the IR regularisation are indeed available
359: \cite{Litim:2000ci,Litim:2001up,Litim:2001fd}. The main observation is
360: that the flow trajectory of \eq{ERG} depends on the regularisation.
361: This observation is most important for approximated flows: typically,
362: their endpoint also depends spuriously on the regularisation. The
363: dependence is absent for the full integrated flow. Hence, in order to
364: provide reliable physical predictions, it is important to seek for
365: regularisations for which the main physical informations are already
366: contained within a few leading order terms of an approximation. This
367: issue is intimately linked to the stability of flows. \step
368:
369: The main ingredient in \eq{ERG} is the full inverse propagator. Due
370: to the IR regularisation, the full inverse regularised propagator
371: displays a gap as a function of momenta
372: \cite{Litim:2001up,Litim:2001fd},
373: %
374: \beq\label{Effective1}
375: \min_{q^2\ge 0}
376: \left(
377: \left.
378: \0{\delta^{2}\Gamma_k[\phi]}{\delta \phi(q)\delta\phi(-q)}
379: \right|_{\phi=\phi_0}
380: +R_k(q^2)\right) = C\, k^2 >0\,.
381: \eeq
382: %
383: The functional derivative is evaluated at a properly chosen expansion
384: point $\phi_0$. The existence of the gap $C>0$ implies an IR
385: regularisation, and is a prerequisite for the ERG formalism.
386: Elsewise, \eq{ERG} becomes singular at points where the full inverse
387: effective propagator develops zero modes. It is expected that an
388: approximated ERG flow in the space of all action functionals is most
389: stable if the regularised full propagator is most regular along the
390: flow. This reasoning corresponds to maximising the gap. \step
391:
392: To leading order in the derivative expansion, and dropping irrelevant
393: momentum-independent terms, the gap is given by
394: %
395: \beq\label{C}
396: C=\min_{y\ge 0} y(1+r(y))\,.
397: \eeq
398: %
399: Within the local potential approximation, the gap $C$ in \eq{C}
400: depends on the regularisation, but not on the specific theory
401: considered. A comparison of the gap of different regulators requires
402: an appropriate normalisation of $r$. In order to make the flow
403: \eq{ERG} more stable, we require that the gap, as a function of the
404: RS, is maximal in the momentum regime where the flow receives its main
405: contributions. This is the optimisation criterion of
406: \cite{Litim:2000ci,Litim:2001up,Litim:2001fd}. It corresponds to an
407: optimisation of the radius of convergence of amplitude expansions and
408: the derivative expansion. It can also be shown that the optimisation
409: is closely linked to a minimum sensitivity condition. Optimised
410: regulators are those for which the maximum in \eq{C} is attained. In
411: Ref.~\cite{Litim:2001up}, these considerations have lead to a very
412: specific solution to the optimisation condition, given by
413: %
414: \beq\label{Ropt} R_{\rm opt}(q^2) = (k^2-q^2)\,\theta (k^2-q^2)\,. \eeq
415: %
416: It has a number of interesting properties. For large momenta
417: $q^2>k^2$, the propagation of modes is fully suppressed since
418: $R\equiv 0$. For small momenta $q^2<k^2$, all modes propagate with
419: an effective mass term given by the IR scale, $q^2+R(q^2)=k^2$.
420: Based on the discussion in
421: Refs.~\cite{Litim:2000ci,Litim:2001up,Litim:2001fd}, we expect that
422: \Eq{Ropt} leads to an improved convergence and hence better physical
423: predictions already to leading order in the derivative expansion.
424: Rewriting \eq{Ropt} in the form of \eq{r} leads to
425: %
426: \beq\label{ropt} r_{\rm opt}(y) = (\s01y-1)\,\theta (1-y) \eeq
427: %
428: Below, we mainly employ this regulator, and variants thereof (cf.
429: Appendices~\ref{VariantsOpt} and \ref{SlidingOpt}). When expressed in
430: terms of the optimised regulator \eq{Ropt}, the flow equation
431: \eq{FlowPotential} becomes
432: %
433: \beq\label{FlowPotentialOpt}
434: \partial_t u= -d u
435: +(d-2) \rho u'
436: +\04d v_d \0{N-1}{1+u'}
437: +\04d v_d \0{1}{1+u'+2\rho u''}\,.
438: \eeq
439: %
440: Notice that the numerical factors $(4v_d)/d$ can be absorbed into the
441: potential and the fields by an appropriate rescaling.
442:
443: %********|*********|*********|*********|*********|*********|*********|****
444: \section{Fixed points}\label{FixedPoints}
445: %********|*********|*********|*********|*********|*********|*********|****
446:
447: The flow equation \eq{FlowPotentialOpt} is known to exhibit two
448: scaling solutions in $2<d<4$, which correspond to the Gaussian and the
449: Wilson-Fisher fixed point, respectively. The Gaussian fixed point is
450: given by the trivial solution $u_\star=$ const, and is discussed in
451: the Appendix~\ref{Gauss} for arbitrary regularisation and $2<d<4$.
452: In this section, we study the non-trivial fixed point $u_\star\neq$
453: const.~in $3d$ based on an optimised flow. We introduce the numerical
454: method, and discuss the scaling form of the fixed point solution.
455: Universal critical exponents are computed in the next section.
456:
457: %********|*********|*********|*********|*********|*********|*********|****
458: \subsection{Numerics}
459: %********|*********|*********|*********|*********|*********|*********|****
460:
461: Numerical methods for solving partial differential equations are
462: well-known. Here, we employ a polynomial truncation of the scaling
463: potential, retaining vertex functions $\phi^{2 n}$ up to a maximum
464: number $n_{\rm trunc}$. Polynomial approximations are reliable if they
465: depend only very weakly on higher order operators beyond some finite
466: order of the truncation. (In the following section, we explicitly
467: confirm that this is indeed the case.) Two different polynomial
468: expansions of the potential are used: expansion I corresponds to
469: %
470: \beq\label{PolyAnsatz0}
471: u(\rho)=\sum^{n_{\rm trunc}}_{n=1} \01{n!} \lambda_n \rho^n\,.
472: \eeq
473: %
474: In \Eq{PolyAnsatz0}, $\lambda_1$ denotes the (dimensionless) mass term
475: at the origin. We have normalised the potential as $u(\rho=0)=0$. All
476: higher order coefficients $\lambda_n$ denote the $n$-th order coupling
477: at vanishing field. Expansion II, alternatively, approximates the
478: potential about the potential minimum $\rho=\rho_0$ as
479: %
480: \beq\label{PolyAnsatz}
481: u(\rho)=\sum^{n_{\rm trunc}}_{n=2} \01{n!} \lambda_n (\rho-\lambda_1)^n\,.
482: \eeq
483: %
484: In \Eq{PolyAnsatz}, $\lambda_1$ denotes the location of the potential
485: minimum defined by $u'(\rho=\lambda_1)=0$. We have normalised the
486: potential as $u(\lambda_1)=0$. All higher order coefficients
487: $\lambda_n$ denote the $n$-th order coupling at the potential minimum.
488: Both expansions \eq{PolyAnsatz0} and \eq{PolyAnsatz} are symmetric
489: under $\phi\to -\phi$, and approximate the potential as an even
490: polynomial in $\phi$. The number of independent operators contained in
491: \Eq{PolyAnsatz0} or \Eq{PolyAnsatz} is $n_{\rm trunc}$. The
492: expansions transform the partial differential equation
493: \eq{FlowPotential} into $n_{\rm trunc}$ coupled ordinary differential
494: equations $\partial_t \lambda_i \equiv \beta_i(\{\lambda_n\})$ for the
495: set of couplings
496: %
497: \beq
498: \{\lambda_n,1\le n\le n_{\rm trunc}\}\,.
499: \eeq
500: %
501: For the numerical study, we use mostly expansion II, which is known to
502: have superior convergence properties in comparison to expansion I
503: \cite{Aoki:1998um,Litim:2000ci}.
504:
505: %********|*********|*********|*********|*********|*********|*********|****
506: \subsection{Wilson-Fisher fixed point}
507: %********|*********|*********|*********|*********|*********|*********|****
508:
509: The Wilson-Fisher fixed point corresponds to the non-trivial scaling
510: solution of \eq{FlowPotentialOpt}. Here, we restrict ourselves to the
511: case $d=3$, and to the optimised regulator as introduced above. The
512: scaling potential obeys the differential equation
513: %
514: \beq\label{FlowPotential3}
515: 0 = -3 u _\star
516: + \rho u'_\star
517: +\01{6\pi^2} \0{N-1}{1+u'_\star}
518: +\01{6\pi^2} \0{1}{1+u'_\star+2\rho u''_\star}+{\rm const.}\,
519: \eeq
520: %
521: and $u _\star(\rho)\neq{\rm const}$. The constant in
522: \Eq{FlowPotential3} can be fixed freely, and we chose it such that the
523: scaling potential vanishes at its minimum, $u_\star=0$ at the point
524: $\rho_0$ with $u'_\star=0$. An analytical solution of
525: \Eq{FlowPotential3} has been given in Ref.~\cite{Litim:1995ex} for the
526: limit $N=\infty$. For $N\neq\infty$, and in the vicinity of $\rho=0$,
527: the scaling solution can be obtained analytically as a Taylor
528: expansion in the field. For $d=3$ and $N=1$, and inserting the
529: expansion \eq{PolyAnsatz0} into \eq{FlowPotential3} with const.~$=0$,
530: we find
531: %
532: \begin{mathletters}\label{solutionExpansion}
533: \bea
534: \lambda_0&=&\ \ \s01{18}\pi^{-2}(1 + \lambda_1)^{-1}\\
535: \lambda_2&=&-4\pi^2\,\lambda_1\,{\left( 1 + \lambda_1 \right) }^2 \\
536: \lambda_3&=&\ \
537: \s0{72}{15}\pi^4\,\lambda_1\, {\left( 1 + \lambda_1 \right) }^3\,
538: \left( 1 + 13\,\lambda_1 \right) \\
539: \lambda_4&=&
540: -\s0{1728}{7}\pi^6\,\lambda_1^2\,{\left( 1 + \lambda_1\right) }^4\,
541: \left( 1 + 7\,\lambda_1 \right) \\
542: \lambda_5&=&\ \
543: \s0{768}{7}\pi^8\,\lambda_1^2\,{\left( 1 + \lambda_1 \right) }^5\,
544: ( 2 + 121\,\lambda_1 + 623\,\lambda_1^2)\\
545: &\vdots&\nonumber \eea
546: \end{mathletters}%
547: %
548: for small fields. Similar explicit solutions are found for $d\neq 3$
549: or $N\neq 1$. Notice that all couplings are expressed as functions of
550: the dimensionless mass term at vanishing field, $\lambda_1$. The
551: domain of validity of this expansion is restricted to $0\le \rho\le
552: \rho_c<\infty$. From the explicit solution for large fields,
553: \eq{solutionExpansionLarge} below, it is clear that the polynomial
554: approximation cannot be extended to arbitrary large fields. The value
555: $\rho_c$ defines the radius of convergence for the polynomial
556: approximation. It is linked to the gap parameter $C$ introduced
557: earlier \cite{Litim:2000ci}. Not all values for $\lambda_1$ lead to a
558: scaling solution which remains finite and analytical for all
559: $\rho<\rho_c$. It is at this point where the quantisation of
560: $\lambda_1$ becomes manifest: only two values for $\lambda_1$
561: correspond to well-defined solutions of the fixed point equation,
562: given by the Gaussian fixed point $\lambda_1=0$ and the Wilson-Fisher
563: fixed point $\lambda_1=\lambda_{1\star}< 0$. The value for
564: $\lambda_{1\star}$ is determined by fine tuning $\lambda_1$ such that
565: the solution \eq{solutionExpansion} extends to $\rho=\rho_c$. The
566: result is $\lambda_{1\star} = -0.1860642\cdots$ for $N=1$. In the
567: other limit, for large fields $\rho\gg 1$, we find
568: %
569: \beq\label{solutionExpansionLarge}
570: u(\rho)=A\,\rho^3+\s0{1}{450\pi^2}\,A^{-1}\,\rho^{-2} +\ldots\,,
571: \eeq
572: %
573: (and similarly for $N\neq 1)$, where the dots denote subleading terms
574: in $\rho$ and the constant $A>0$ has to be fixed appropriately, in a
575: way similar to $\lambda_1$ in \eq{solutionExpansion}. It would be
576: interesting to study the analyticity properties of the fixed point
577: solutions more deeply, and to contrast it with the analysis of
578: \cite{Morris:1994ki} for the sharp cut-off flow. Despite the simple
579: explicit form of the flow, and its explicit solution for large and
580: small fields, it is more efficient to identify the scaling solution
581: and the related critical exponents using numerical methods.
582:
583: %********|*********|*********|*********|*********|*********|*********|****
584: \subsection{Scaling potential}
585: %********|*********|*********|*********|*********|*********|*********|****
586:
587: We have solved the resulting coupled set of differential equations in
588: $d=3$ dimensions, for $N=0,1,\cdots, 10$ and for truncations up to
589: $n_{\rm trunc}=20$. We have retained only the solution which has one
590: unstable eigendirection, corresponding to the Wilson-Fisher fixed
591: point. Fig.~1 shows the corresponding scaling potential $u_\star$ in
592: the vicinity of the potential minimum. In Fig.~1, the unusual
593: normalisation of the potential has been chosen only for display
594: purposes. The scaling potential has, for all $N$, a minimum at
595: non-vanishing field. For this reason, the expansion II has better
596: convergence properties than expansion I.
597:
598: %********|*********|*********|*********|*********|*********|*********|****
599: %********|*********| Fig1: pSSB_Pot_N_Scaled.ps |*********|*********|****
600: %********|*********| Fig2: pSSB_Pot1_N_Scaled.ps |*********|*********|****
601: %********|*********|*********|*********|*********|*********|*********|****
602: \begin{figure}
603: \begin{center}
604: \vskip-1cm
605: \unitlength0.001\hsize
606: \begin{picture}(1000,550)
607: \put(290,155){\framebox{\large $ c_N\, u_\star(\rho)$}}
608: \put(220,80){\large $\rho$}
609: \put(740,85){\large $\rho$}
610: \put(840,160){\framebox{\large $u'_\star(\rho)$}}
611: \psfig{file=pSSB_Pot_N_Scaled.ps,width=.45\hsize}
612: \hskip.05\hsize
613: \psfig{file=pSSB_Pot1_N_Scaled.ps,width=.45\hsize}
614: \end{picture}
615: \vskip-.75cm
616: \begin{minipage}{.47\hsize}
617: { \small {\bf Figure 1:} The scaling potential, with $c_N=40-2N$.
618: From left to right: $N=0,1,\cdots,10$.}
619: \end{minipage}
620: \hskip.05\hsize
621: \begin{minipage}{.47\hsize}
622: { \small {\bf Figure 2:} The amplitude $u'_\star(\rho)$ of the scaling
623: solution. From left to right: $N=0,1,\cdots,10$.}
624: \end{minipage}
625: \end{center}
626: \end{figure}
627: %********|*********|*********|*********|*********|*********|*********|****
628:
629:
630: %********|*********|*********|*********|*********|*********|*********|****
631: %********|*********| Fig3: pSSB_Pot2_N_Scaled.ps |*********|*********|****
632: %********|*********|*********|*********|*********|*********|*********|****
633: \begin{figure}
634: \begin{center}
635: \vskip-1cm
636: \unitlength0.001\hsize
637: \begin{picture}(1000,550)
638: \put(190,150){\framebox{\large $u'_\star(\rho)+2\rho u''_\star(\rho)$}}
639: \put(220,70){\large $\rho$}
640: \hskip.05\hsize
641: \put(520,150){
642: \begin{minipage}{.47\hsize}
643: {\small {\bf Figure 3:} The amplitude $u'_\star+2\rho u''_\star$ of
644: the scaling solution. The dots indicate the points where
645: $u'_\star=0$. From left to right: $N=0,1,\cdots,10$.}
646: \end{minipage}}
647: \psfig{file=pSSB_Pot2_N_Scaled.ps,width=.45\hsize}
648: \end{picture}
649: \end{center}
650: \vskip-1.cm
651: \end{figure}
652: %********|*********|*********|*********|*********|*********|*********|****
653:
654:
655: Fig.~2 shows the amplitude $u'$ at the fixed point. It is a monotonic
656: function of the field. It approaches its most negative values at
657: vanishing field. The flow equation would have a pole at points where
658: $1+u'$ vanishes. This is, however, never the case, because $1+u'$
659: stays always positive. Fig.~3 shows the radial mode of the scaling
660: potential $u'_\star+ 2\rho u''_\star$ in the vicinity of the potential
661: minimum. It is a monotonic function of the fields. The dots in Fig.~3
662: indicate where the derivative $u'_\star$ changes sign. Again, the most
663: negative value is attained at vanishing field and decreases with
664: increasing $N$, but it stays always above the pole of the flow
665: equation, $1+u'_\star+ 2\rho u''_\star>0$.\step
666:
667: The scaling solution is non-universal. However, critical exponents or
668: the eigenvalues of small perturbations about the scaling solutions
669: {\it are} universal. For their determination, it is sufficient to
670: study the flow of small perturbations in the vicinity of the scaling
671: potential, which is done next.
672:
673:
674: %********|*********|*********|*********|*********|*********|*********|****
675: \section{Critical exponents}\label{CriticalExponents}
676: %********|*********|*********|*********|*********|*********|*********|****
677:
678: In this section, we compute the critical exponents to leading order in
679: the derivative expansion. Numerical results are given up to six
680: significant figures (Tab.~1). A higher precision can be achieved, but
681: is not required at the present state. We find a rapid convergence of
682: the polynomial approximation, for all observables considered. Our
683: results are compared with those from other regulators.
684:
685: %********|*********|*********|*********|*********|*********|*********|****
686: \subsection{Eigenvalues}
687: %********|*********|*********|*********|*********|*********|*********|****
688:
689: Given the Wilson-Fisher fixed point solution, we can seek for
690: universal critical exponents. In the vicinity of the non-trivial fixed
691: point, we have to solve the eigenvalue equation
692: %
693: \beq\label{WFFP:1}
694: \partial_t\ \delta u^{(m)}=\omega\,\delta u^{(m)}
695: \eeq
696: %
697: in order to determine the various universal eigenvalues $\omega$.
698: Using the flow equation, setting $d=3$ and choosing $m=1$, and
699: expanding the flow to leading order about the Wilson-Fisher fixed
700: point, we find
701: %
702: \bea
703: 0&=&
704: \left[\omega+2-\0{N}{6\pi^2}\0{u''_\star}{(1+u'_\star)^2}\right]
705: \,\delta u'
706: +\,\0{1}{3\pi^2}
707: \left[ \0{2N}{1+u'_\star}
708: -\0{1+u'_\star-\rho u''_\star-2\rho^2 u'''_\star}
709: {(1+u'_\star+2\rho u''_\star)^2}\right]
710: \,\delta u''
711: \nonumber \\ &&
712: +\,\0{1}{3\pi^2}\,\0{\rho}{1+u'_\star+2\rho u''_\star}\ \delta u'''\,.
713: \label{WFFP:2}
714: \eea
715: %
716: Instead of solving \Eq{WFFP:2} directly for an eigenperturbation
717: $\delta u'$ with eigenvalue $\omega$, we follow a slightly different
718: line which is numerically less demanding. Based on the
719: polynomial expansion used in the previous section, the fixed point
720: potential is given by the set of couplings $\{\lambda_{n,\star}\}$. At
721: the fixed point, the flow of the couplings $\partial_t\lambda_n\equiv
722: \beta_n$ vanishes, $\beta_n(\lambda_{i,\star})=0$. The eigenvalues
723: $\omega$ of the stability matrix at criticality
724: %
725: \beq\label{M}
726: M_{ij}=
727: \left.\partial\beta_i/\partial\lambda_j\right|_{\lambda=\lambda_{\star}}
728: \eeq
729: %
730: correspond to the eigenvalues of \Eq{WFFP:1}. Hence, the problem of
731: finding the eigenvalues of \Eq{WFFP:1} reduces to the problem of
732: finding the eigenvalues of the stability matrix $M$.\step
733:
734: %********|*********|*********|*********|*********|*********|*********|****
735: %********|*********| Tab1 |*********|*********|*********|*********|****
736: %********|*********|*********|*********|*********|*********|*********|****
737: \begin{center}
738: \begin{tabular}{c|c|c|c|c|c}
739: $\quad{} N \quad{} $&
740: $\quad\quad\quad{}\nu \quad\quad\quad{}$&
741: $\quad\quad\quad{}\omega\quad\quad\quad{}$&
742: $\quad\quad\quad{}\omega_2\quad\quad\quad{}$&
743: $\quad\quad\quad{}\omega_3\quad\quad\quad{}$&
744: $\quad\quad\quad{}\omega_4\quad\quad\quad{}$
745: \\[.5ex] \hline%\\[-1.5ex]
746: 0
747: &0.592083
748: &0.65788
749: &3.308
750: &6.16
751: &9.2
752: \\
753: 1
754: &0.649562
755: &0.655746
756: &3.180
757: &5.912
758: &8.80
759: \\
760: 2
761: &0.708211
762: &0.671221
763: &3.0714
764: &5.679
765: &8.440
766: \\
767: 3
768: &0.761123
769: &0.699837
770: &2.9914
771: &5.482
772: &8.125
773: \\
774: 4
775: &0.804348
776: &0.733753
777: &2.9399
778: &5.330
779: &7.867
780: \\
781: 5
782: &0.837741
783: &0.766735
784: &2.9108
785: &5.2195
786: &7.665
787: \\
788: 6
789: &0.863076
790: &0.795815
791: &2.8967
792: &5.1409
793: &7.512
794: \\
795: 7
796: &0.882389
797: &0.820316
798: &2.8916
799: &5.0863
800: &7.396
801: \\
802: 8
803: &0.897338
804: &0.840612
805: &2.89163
806: &5.04848
807: &7.3086
808: \\
809: 9
810: &0.909128
811: &0.857384
812: &2.89438
813: &5.02232
814: &7.2425
815: \\
816: 10
817: &0.918605
818: &0.871311
819: &2.89846
820: &5.00420
821: &7.1921
822: \\
823: $\infty$
824: &1
825: &1
826: &3
827: &5
828: &7
829: %\\[1ex]\hline\hline
830: \end{tabular}
831: \end{center}
832: \begin{center}
833: \begin{minipage}{\hsize}
834: \vskip.3cm {\small {\bf Table 1:} The first five eigenvalues
835: $\omega_0<0<\omega_1<\omega_2<\omega_3<\omega_4$, with
836: $\nu=-1/\omega_0>0$ for various $N$ and $d=3$ dimensions.}
837: \end{minipage}
838: \end{center}
839: %********|*********|*********|*********|*********|*********|*********|****
840:
841: %********|*********|*********|*********|*********|*********|*********|****
842: \subsection{Results}
843: %********|*********|*********|*********|*********|*********|*********|****
844:
845: We have computed the eigenvalues of \eq{M} as functions of the order
846: of the polynomial truncation. The numerical results for the first five
847: eigenvalues for various $N$ and $d=3$ dimensions are given in Tab.~1
848: and in Figs.~4 -- 8. Results for other regulators are compared in
849: \cite{Litim:2001dt} (see also Sects.~\ref{SectionScheme}
850: and~\ref{Bounds}), and results for the asymmetric corrections to
851: scaling are given elsewhere. Exact results, independent on the
852: regularisation, are known for $N=\infty$. The exact eigenvalues are
853: given by $\omega_n=2n-1 +{\cal O}(1/N)$, and only the subleading
854: corrections ${\cal O}(1/N)$ depend on the regulator. For $N=-2$, it is
855: known that $\nu=\s012$. Tab.~1 shows the first five eigenvalues,
856: ordered as $\omega_0<0<\omega<\omega_2<\omega_3<\omega_4$. The
857: critical exponent $\nu$ is given by $\nu=-1/\omega_0>0$. As a function
858: of $N$, the eigenvalue $\nu$ interpolates monotonically between the
859: exact values $\nu=\s012$ for $N=-2$ and $\nu=1$ for $N=\infty$. The
860: eigenvalues $\omega$ and $\omega_2$ are non-monotonic functions of
861: $N$. For small $N$, $\omega$ decreases until $N\approx 1$, and
862: increases towards the exact asymptotic value $\omega=1$ for
863: $N\to\infty$. The eigenvalue $\omega_2$ decreases until $N\approx
864: 7-8$ before settling to the asymptotic value $\omega_2=3$ at
865: $N=\infty$. The eigenvalues $\omega_3$ and $\omega_4$ are
866: monotonically decreasing functions of $N$.
867:
868:
869: %********|*********|*********|*********|*********|*********|*********|****
870: %********|*********| Fig4: pSSB_Nu_N.ps |*********|*********|****
871: %********|*********| Fig5: pSSB_Omega_N.ps |*********|*********|****
872: %********|*********|*********|*********|*********|*********|*********|****
873: \begin{figure}[t]
874: \begin{center}
875: \vskip-.8cm
876: \unitlength0.001\hsize
877: \begin{picture}(1000,550)
878: \put(330,400){\framebox{\large $\nu(N)$}}
879: \put(210,70){\large $n_{\rm trunc}$}
880: \put(830,460){\framebox{\large $\omega(N)$}}
881: \put(710,70){\large $n_{\rm trunc}$}
882: \psfig{file=pSSB_Nu_N.ps,width=.45\hsize}
883: \hskip.05\hsize
884: \psfig{file=pSSB_Omega_N.ps,width=.45\hsize}
885: \end{picture}
886: \vskip-.75cm
887: \begin{minipage}{.45\hsize}
888: {\small {\bf Figure 4:} The exponent $\nu(N)$ as a function
889: of $N$ and of the order of the truncation. From top to bottom:
890: $N=10,4,3,2,1,0$.}
891: \end{minipage}
892: \hskip.05\hsize
893: \begin{minipage}{.45\hsize}
894: {\small {\bf Figure 5:} The eigenvalue $\omega(N)$ as a function of
895: $N$ and of the order of the truncation. From top to bottom:
896: $N=10,4,3,2,0,1$.}
897: \end{minipage}
898: \end{center}
899: \vskip-.1cm
900: \end{figure}
901: %********|*********|*********|*********|*********|*********|*********|****
902:
903:
904: Fig.~4 shows the results for $\nu$ as a function of $n_{\rm trunc}$
905: and $N$. It is seen that the expansion is very stable. It convergences
906: already within low order of the truncation towards the asymptotic
907: value. Typically, $n_{\rm trunc}\approx 6$ gives the correct result
908: below the percent level. Furthermore, the convergence is better for
909: larger values of $N$. This behaviour is observed for all eigenvalues
910: (e.g.~Figs.~4 -- 8). Fig.~5 shows the same behaviour for the smallest
911: subleading eigenvalue $\omega$. For $n_{\rm trunc}\approx 8$, it has
912: settled below the percent level of the correct result.
913: \step
914:
915:
916: %********|*********|*********|*********|*********|*********|*********|****
917: %********|*********| Fig6: pSSB_Omega2_N.ps |*********|*********|****
918: %********|*********| Fig7: pSSB_Omega3_N.ps |*********|*********|****
919: %********|*********|*********|*********|*********|*********|*********|****
920: \begin{figure}[t]
921: \begin{center}
922: {}\vskip-2cm
923: \unitlength0.001\hsize
924: \begin{picture}(1000,550)
925: \put(330,470){\framebox{\large $\omega_2(N)$}}
926: \put(210,70){\large $n_{\rm trunc}$}
927: \put(830,470){\framebox{\large $\omega_3(N)$}}
928: \put(710,70){\large $n_{\rm trunc}$}
929: \psfig{file=pSSB_Omega2_N.ps,width=.45\hsize}
930: \hskip.05\hsize
931: \psfig{file=pSSB_Omega3_N.ps,width=.45\hsize}
932: \end{picture}
933: \vskip-.75cm
934: \begin{minipage}{.47\hsize}
935: {\small {\bf Figure 6:} The critical index $\omega_2(N)$ as a
936: function of $N$ and of the order of the truncation. From top to
937: bottom: $N=0,1,2,3,4,10$.}
938: \end{minipage}
939: \hskip.05\hsize
940: \begin{minipage}{.47\hsize}
941: {\small {\bf Figure 7:} The critical index $\omega_3(N)$ as a
942: function of $N$ and of the order of the truncation. From top to
943: bottom: $N=0,1,2,3,4,10$.}
944: \end{minipage}
945: \end{center}
946: \vskip-1.cm
947: \end{figure}
948: %********|*********|*********|*********|*********|*********|*********|****
949:
950: %********|*********|*********|*********|*********|*********|*********|****
951: %********|*********| Fig8: pSSB_Omega4_N.ps |*********|*********|****
952: %********|*********|*********|*********|*********|*********|*********|****
953: \begin{figure}
954: \begin{center}
955: \unitlength0.001\hsize
956: \begin{picture}(1000,550)
957: \put(340,470){\framebox{\large $\omega_4(N)$}}
958: \put(240,70){\large $n_{\rm trunc}$}
959: \put(520,220){
960: \begin{minipage}{.47\hsize}
961: {\small {\bf Figure 8:} The critical index $\omega_4(N)$ as a
962: function of $N$ and of the order of the truncation. From top to
963: bottom: $N=0,1,2,3,4,10$. The isolated point at $n=7$ corresponds to
964: $N=10$, and the two at $n=9$ to $N=2$ (upper) and $N=3$ (lower). The
965: intermediate points (at $n=8$ and $n=10$, resp.) are missing because
966: they have a small imaginary part.}
967: \end{minipage}}
968: \hskip.02\hsize
969: \psfig{file=pSSB_Omega4_N.ps,width=.45\hsize}
970: \end{picture}
971: \end{center}
972: \vskip-1.5cm
973: \end{figure}
974: %********|*********|*********|*********|*********|*********|*********|****
975:
976: For low orders of the truncation, the eigenvalues $\omega_2$,
977: $\omega_3$ and $\omega_4$ (Figs.~6, 7 and 8, respectively) depend more
978: strongly on $n_{\rm trunc}$, and do not even lead to real eigenvalues
979: in some cases (c.f.~Figs.~7 and 8). Again, the dependence is even
980: stronger for smaller $N$. For sufficiently high order in the
981: truncation, however, all eigenvalues are real, and the convergence
982: towards the asymptotic value is fast. Typically, $\omega_2$,
983: $\omega_3$ and $\omega_4$ reach their asymptotic values below the
984: percent level for $n_{\rm trunc}\approx 10, 12$ and $14$. \step
985:
986: For all eigenvalues, the basic picture is the same: for small $n_{\rm
987: trunc}$, the truncation tends to overshoot the asymptotic value, but
988: with increasing $n_{\rm trunc}$ it relaxes towards it with a remaining
989: oscillation and decreasing amplitude. From a specific order onwards,
990: the truncation sits -- for all technical purposes -- on top of the
991: asymptotic value. For a fixed level of accuracy, a lower order in the
992: truncation is required for the dominant observables like $\nu$ or
993: $\omega$, while a higher order is required for the subleading
994: eigenvalues. Roughly speaking, to obtain the eigenvalue $\omega_n,
995: n=0,\cdots$ accurate below the percent level, a truncation with
996: $n_{\rm trunc}\approx 2n+6$ independent couplings is required.
997: %\step
998:
999:
1000: %********|*********|*********|*********|*********|*********|*********|****
1001: \subsection{Convergence and stability}
1002: %********|*********|*********|*********|*********|*********|*********|****
1003:
1004: Next, we discuss the convergence and stability of the polynomial
1005: approximation for an optimised flow. From the results presented so
1006: far (cf.~Figs.~4 - 8), we conclude that the optimised flow
1007: \eq{FlowPotentialOpt} leads to a fast convergence of the polynomial
1008: approximation for the scaling potential. More importantly, we have
1009: seen that the inclusion of further vertex functions --- increasing
1010: $n_{\rm trunc}\to n_{\rm trunc}+1$ --- does not alter the fixed point
1011: structure. Rather, it leads to a small modification of the actual
1012: fixed point solution and to minor corrections for the critical
1013: exponents. This implies that the flow is very stable, and that most of
1014: the physical information is already contained in a few leading order
1015: terms of the truncation. This result is by no means trivial. A counter
1016: example is furnished by the sharp cutoff (see also the following
1017: section), where the convergence of the polynomial approximation is
1018: poor. Here, the good convergence hinges on the use of an appropriately
1019: optimised regulator
1020: \cite{Litim:2000ci,Litim:2001up,Litim:2001fd}.
1021:
1022: %********|*********|*********|*********|*********|*********|*********|****
1023: %********|*********| Fig9: pConvergenceNu.ps |*********|*********|****
1024: %********|*********|*********|*********|*********|*********|*********|****
1025: \begin{figure}
1026: \begin{center}
1027: {}\vskip-2.cm
1028: \unitlength0.001\hsize
1029: \begin{picture}(1000,550)
1030: \put(200,360){ \fbox{{ $\displaystyle
1031: \log_{10}\left|\0{\nu_{\rm opt}-\nu_{\rm trunc}}{\nu_{\rm opt}}\right|$}}}
1032: \put(230,-10){\large $n_{\rm trunc}$}
1033: \put(520,140){
1034: \begin{minipage}{.47\hsize}
1035: { \small {\bf Figure 9:} Ising universality class. Convergence of
1036: $\nu_{\rm trunc}$ (expansion II) towards $\nu_{\rm opt}$ with
1037: increasing truncation. Points where $\nu_{\rm trunc}$ is larger
1038: (smaller) than $\nu_{\rm opt}$ are denoted by o ($\bullet$).
1039: Roughly speaking, for $2<n_{\rm trunc}<20$, the accuracy of the
1040: critical exponents improves by one decimal point every $\Delta
1041: n\approx 2-2.5$. }
1042: \end{minipage}}
1043: \hskip.02\hsize
1044: \psfig{file=Converge1.ps,width=.45\hsize}
1045: \end{picture}
1046: \end{center}
1047: \end{figure}
1048: %********|*********|*********|*********|*********|*********|*********|****
1049:
1050:
1051: Let us have a closer look at the rate of convergence towards the
1052: asymptotic values of expansion II (cf.~Fig.~9). We denote with
1053: $\nu_{\rm trunc}$ the approximate critical exponent which retains
1054: $n=n_{\rm trunc}$ independent parameters in the effective action. The
1055: semi-logarithmic plot in Fig.~9 then shows the rate with which
1056: successive approximations converge towards the asymptotic value
1057: $\nu_{\rm opt}$. The series $\nu_{\rm trunc}$ oscillates about the
1058: asymptotical values with a decreasing amplitude and, roughly, a
1059: four-fold periodicity in the pattern $++--$. The curve in Fig.~9 can
1060: be approximated by a straight line with a slope $\approx -.4$ to
1061: $-.5$. Hence, for every $\Delta n\approx 2 - 2.5$, the accuracy of
1062: $\nu_{\rm trunc}$ increases by one decimal place. Here, we have
1063: analysed the convergence for $N=1$. For larger $N$, the convergence is
1064: typically faster than for $N=1$, while for $N=0$, it is about the
1065: same. Hence, the present considerations are qualitatively the same for
1066: all $N$.
1067:
1068:
1069: %********|*********|*********|*********|*********|*********|*********|****
1070: %********|*********| Fig10: pConvergenceOpt.ps |*********|*********|****
1071: %********|*********|*********|*********|*********|*********|*********|****
1072: \begin{figure}
1073: \begin{center}
1074: \vskip-.5cm
1075: \unitlength0.001\hsize
1076: \begin{picture}(850,750)
1077: \put(200,350){\large $n_{\rm trunc}$}
1078: \put(610,350){\large $n_{\rm trunc}$}
1079: \put(280,450){\fbox{\large $\nu_{\rm Ising}$}}
1080: \put(110,510){I}
1081: \put(110,650){II}
1082: \put(600,650){I}
1083: \put(520,550){II}
1084: \psfig{file=pConvergenceOpt.ps,width=.8\hsize}
1085: \end{picture}
1086: \vskip-5cm
1087: \begin{minipage}{.92\hsize}
1088: {\small {\bf Figure 10:} The critical index $\nu$ for the Ising
1089: universality class as a function of the order of the truncation
1090: $n_{\rm trunc}$, for expansion I around vanishing field and
1091: expansion II around the non-trivial minimum. Left panel: I and II
1092: converge towards the asymptotic value. Right panel: Magnification
1093: by a factor of 625. The variation of all data points is below
1094: $10^{-3}$. II converges faster than I. Expansion I (II)
1095: fluctuates about the asymptotic value with decreasing amplitude
1096: and a six-fold (four-fold) periodicity; see also Fig.~9.}
1097: \end{minipage}
1098: \end{center}
1099: \end{figure}
1100: %********|*********|*********|*********|*********|*********|*********|****
1101:
1102: Until now, we have only employed expansion II, defined in
1103: \eq{PolyAnsatz}. It is worthwhile to employ as well expansion I,
1104: defined in \eq{PolyAnsatz0}. This has been done for the critical
1105: exponent $\nu$ of the Ising universality class in Fig.~10. The left
1106: panel shows that {\it both} expansions lead to a fast convergence
1107: towards the asymptotic value. The right panel is a magnification by
1108: 625, showing that expansion II indeed converges faster, although for
1109: $n_{\rm trunc}=10$ their difference is already below $10^{-3}$.
1110: Expansion I fluctuates about the asymptotic value with decreasing
1111: amplitude and a six-fold periodicity in the pattern $+++---$.
1112:
1113: %********|*********|*********|*********|*********|*********|*********|****
1114: \subsection{Convergence and scheme dependence} \label{SectionScheme}
1115: %********|*********|*********|*********|*********|*********|*********|****
1116:
1117:
1118: In this section, we discuss the convergence of the ERG flows and the
1119: polynomial expansion for various regulators. In Fig.~11, we have
1120: computed the critical exponent $\nu$ $(N=1)$ for the sharp cutoff
1121: $r_{\rm sharp}(y)=1/\,\theta(1-y)-1$, the quartic regulator $r_{\rm
1122: quart}(y)=y^{-2}$ and the optimised regulator \eq{ropt}. The left
1123: (right) panel uses the expansion I (II). Both $r_{\rm quart}$ and
1124: $r_{\rm opt}$ are optimised regulators in the sense coined in section
1125: \ref{Optimisation}. From Fig.~11, three results are
1126: noteworthy. First, for the expansion I, we confirm that the
1127: convergence is very poor for the sharp cutoff. For both the quartic
1128: and the optimised regulator we find a good convergence. Second, the
1129: convergence is additionally improved by switching to the expansion II.
1130: Third, the critical exponents obey $\nu_{{\rm sharp}}> \nu_{{\rm
1131: quart}}> \nu_{{\rm opt}}$. Hence, the better the convergence and
1132: the stability of the flow, the smaller the resulting critical exponent
1133: $\nu$. This observation is also linked to the convergence of the
1134: derivative expansion \cite{Litim:2001dt}.
1135: %\step
1136:
1137: %********|*********|*********|*********|*********|*********|*********|****
1138: %********|*********| Fig11: pConvergenceERG.ps |*********|*********|****
1139: %********|*********|*********|*********|*********|*********|*********|****
1140: \begin{figure}
1141: \begin{center}
1142: \unitlength0.001\hsize
1143: \begin{picture}(850,750)
1144: \put(200,370){\large $n_{\rm trunc}$}
1145: \put(610,370){\large $n_{\rm trunc}$}
1146: \put(300,650){I}
1147: \put(700,650){II}
1148: \put(280,450){\fbox{\large $\nu_{\rm Ising}$}}
1149: \psfig{file=pConvergenceERG.ps,width=.8\hsize}
1150: \end{picture}
1151: \vskip-5cm
1152: \begin{minipage}{.92\hsize}
1153: { \small {\bf Figure 11:} The critical index $\nu$ for the Ising
1154: universality class. Results are given for Expansion I (left panel)
1155: and II (right panel), and for the sharp cutoff (upper curves) the
1156: quartic regulator (middle curves) and the optimised regulator
1157: (lower curves).}
1158: \end{minipage} \end{center}
1159: \end{figure}
1160: %********|*********|*********|*********|*********|*********|*********|****
1161:
1162: In \cite{Litim:2000ci}, it has been shown that the gap $C$ (the
1163: radius of convergence for amplitude expansions) is linked to the
1164: radius of convergence $C'$ for the expansions \eq{PolyAnsatz0} and
1165: \eq{PolyAnsatz}. An alternative way for identifying $C'$ consists in
1166: studying the complex structure of the scaling solution. For real
1167: $\rho$, the scaling solution is finite and real for all $\rho<\infty$.
1168: In the complex plane, the scaling solution has (various) poles. Those
1169: closest to the chosen expansion point constrain the radius of
1170: convergence $C'$. For the sharp cutoff and the expansion I, this has
1171: been analysed in \cite{Morris:1994ki} for $N=1$. It was found that
1172: $\nu_{\rm sharp}$ cannot be determined to an accuracy better than
1173: $8\cdot 10^{-3}$. Hence, the polynomial expansion I for a sharp
1174: cutoff does {not} converge beyond a certain level. \step
1175:
1176: However, the findings of \cite{Morris:1994ki} do not imply that
1177: polynomial truncations are not trustworthy {\it per se}. To the
1178: contrary, the decisive difference between ``good'' or ``bad''
1179: convergence properties stems essentially from an appropriate choice of
1180: the regularisation. When optimised, the regularisation implies a
1181: significant improvement for either expansion. In this light, the
1182: non-convergence of the sharp cutoff flow within expansion I is
1183: understood as a deficiency of the sharp cutoff regularisation. This
1184: picture is consistent with the results of \cite{Aoki:1998um}, who
1185: showed that expansion II leads to an improved convergence over
1186: expansion I --- {even} for the sharp cutoff.
1187:
1188:
1189:
1190: %********|*********|*********|*********|*********|*********|*********|****
1191: \section{Bounds on critical exponents} \label{Bounds}
1192: %********|*********|*********|*********|*********|*********|*********|****
1193:
1194: In this section, we discuss how the critical exponents computed in the
1195: previous sections, depend on the regularisation. This discussion is
1196: mandatory because observables computed from a truncated flow, are
1197: known to depend spuriously on the regularisation. It is decisive to
1198: understand the range over which $\nu_{\rm ERG}$ may vary as a function
1199: of the IR regulator. The origin of the spurious RS dependence is
1200: easily understood. The regulator, while regulating the flow, also
1201: modifies the coupling amongst all vertex functions of the theory.
1202: These regulator induced contributions are of no relevance for the
1203: integrated full flow, but they do matter for approximated flows, like
1204: in the present case. It is argued that the smallest value for the
1205: exponent $\nu_{{}_{\rm ERG}}$ is obtained for the optimised regulator
1206: $R_{\rm opt}$. Prior to this, we recall the results obtained
1207: previously in the literature, where, by a number of groups
1208: \cite{Hasenfratz:1986dm,Morris:1994ki,Comellas:1997tf,Liao:2000sh,Litim:2001hk,Litim:2001dt},
1209: critical exponents have been computed based on \eq{ERG} for different
1210: regulators to leading order in the derivative expansion. For all $N$,
1211: all previously published results obey
1212: %
1213: \beq\label{range-lit}
1214: \nu_{{\rm sharp}} \ge
1215: \nu_{{}_{\rm ERG}} \ge
1216: \nu_{{\rm min}} \,.
1217: \eeq
1218: %
1219: The regulators studied in the literature cover the sharp cutoff and a
1220: variety of smooth cutoff (exponential, power-law), and classes of
1221: regulators interpolating between the sharp cutoff and specific smooth
1222: cutoffs. Most results have been published for the Ising universality
1223: class $N=1$. For any $N\ge 0$, the smallest value $\nu_{{\rm min}}$
1224: obtained in the literature is larger than the value $\nu_{\rm opt}$:
1225: $\nu_{{\rm min}}>\nu_{\rm opt}$. For a detailed comparison of critical
1226: exponents to leading and subleading order in the derivative expansion,
1227: and a comparison to results from other methods and experiment, we
1228: refer to Ref.~\cite{Litim:2001dt}.
1229:
1230: %********|*********|*********|*********|*********|*********|*********|****
1231: \subsection{Upper boundary}
1232: %********|*********|*********|*********|*********|*********|*********|****
1233:
1234: Now, we turn to a general discussion on the scheme dependence of
1235: $\nu$. At first sight, \eq{range-lit} suggests that the possible range
1236: for $\nu$ is bounded from above and from below. Let us assess the two
1237: boundaries. We begin by showing that the inequality $\nu_{{\rm
1238: sharp}}\ge \nu_{{}_{\rm ERG}}$ does {not} hold for generic
1239: regulator. Indeed, the upper boundary in \eq{range-lit} can be
1240: overcome by choosing regulators which lead to a worse convergence than
1241: the sharp cutoff. To see this more explicitly, consider a class of
1242: regulators discussed in Appendix~\ref{VariantsOpt}. It is given by
1243: %
1244: \beq\label{Ra}
1245: R_a(q^2)=a\,(k^2-q^2)\,\theta(k^2-q^2)\,,
1246: \eeq
1247: %
1248: and is a variant of the optimised regulator \eq{Ropt}, to which it
1249: reduces for $a=1$. For $a\to\infty$, it corresponds to the sharp cut
1250: off. The regulator leads to an effective radius of convergence $C_a=a$
1251: for $a<1$, and $C_a=\s0{a}{2a-1}$ for $a\ge 1$. Since $C_a\ll 1$ for
1252: $a\ll 1$, we expect that the corresponding critical exponents will
1253: become large. Our results for $R_a$ in \eq{Ra} are given in Fig.~12
1254: (for more details, see Appendix~\ref{VariantsOpt} and Tab.~3). We find
1255: that $\nu_a$ is a monotonously increasing function for decreasing
1256: $a\le 1$. In particular, for small $a$ we find indeed that
1257: $\nu_a>\nu_{\rm sharp}$. Hence, this result confirms the above
1258: picture: Flows with a poor radius of convergence lead to large
1259: numerical values for $\nu$. \step
1260:
1261: %********|*********|*********|*********|*********|*********|*********|****
1262: %********|*********| Tab2 |*********|*********|*********|*********|****
1263: %********|*********|*********|*********|*********|*********|*********|****
1264: \begin{center}
1265: \begin{tabular}{c||c|c|c|c|c|c|c}
1266: $\gamma$&
1267: $\s012$&
1268: $\s034$&
1269: $\s045$&
1270: $\s056$&
1271: 1&
1272: $\s032$&
1273: 2\\[.5ex] \hline%\\[-3ex]
1274: $\quad\nu_\gamma\quad$&
1275: $\ \ \ 1\ \ \ {}$&
1276: $\ .7354\ {}$&
1277: $\ .7216\ {}$&
1278: $\ .7142\ {}$&
1279: $\ .6895\ {}$&
1280: $\ .6604\ {}$&
1281: $\ .6496\ {}$\\[-1ex]
1282: \end{tabular}
1283: \end{center}
1284: \begin{center}
1285: \begin{minipage}{.9\hsize}
1286: \vskip.3cm {\small {\bf Table 2:} Critical exponents $\nu$ (Ising
1287: universality class) for the flows $\ell_\gamma$ and various
1288: $\gamma$.\\}
1289: \end{minipage}
1290: \end{center}
1291: %********|*********|*********|*********|*********|*********|*********|****
1292:
1293: %********|*********|*********|*********|*********|*********|*********|****
1294: %********|*********| Fig12: pBoundary.ps |*********|*********|****
1295: %********|*********|*********|*********|*********|*********|*********|****
1296: \begin{figure}
1297: \begin{center}
1298: {}\vskip-2.cm
1299: \unitlength0.001\hsize
1300: \begin{picture}(1000,550)
1301: \put(290,230){ \begin{tabular}{ccl}
1302: $\ \ R_a:\ {}$ & $\ \bullet$\\[-.3ex]
1303: $\ \ R_\gamma:\ {}$ & \ o \\[-.3ex]
1304: $\ \ R_c:\ {}$ & $\stackrel{\rm \ opt}{\put(0,0){\line(50,0){50}}}$
1305: \end{tabular}}
1306: \put(330,330){\fbox{{\Large $\displaystyle \nu_{{}_{\rm ERG}}$}}}
1307: \put(260,-10){\large $x$}
1308: \put(250,56){\small phys}
1309: \put(230,85){\small opt}
1310: \put(230,120){\small sharp}
1311: \put(230,395){\small large-$N$}
1312: \put(520,180){
1313: \begin{minipage}{.47\hsize}
1314: { \small {\bf Figure 12:} Ising universality class. The critical
1315: exponent $\nu$ for various classes of regulators. For display
1316: purposes, we use $x=\s023(\gamma-\s012)$ for $R_\gamma$,
1317: $x=\0{a}{a+1}$ for $R_a$ and $x\equiv c$ for $R_c$. Boundaries:
1318: The full line (opt) corresponds to $R_c$ and denotes the lower
1319: boundary, the upper full line (large-$N$) denotes the upper
1320: boundary. For comparison: The dashed line (sharp) indicates the
1321: sharp cut off value, and the thick full line (phys) the physical
1322: value.}
1323: \end{minipage}}
1324: \hskip.02\hsize
1325: \psfig{file=pBoundary.ps,width=.45\hsize}
1326: \end{picture}
1327: \end{center}
1328: \end{figure}
1329: %********|*********|*********|*********|*********|*********|*********|****
1330:
1331: Next, we turn to a class of regulators aimed at probing the maximal
1332: spread of flows. We recall that, to leading order in the derivative
1333: expansion, all regulator-dependence is contained in the functions
1334: $\ell(\omega)$. In Ref.~\cite{Litim:2001hk}, it has been shown that
1335: the function $\ell(\omega)$ decays at most as $\omega^{-1}$, if the
1336: regulator, obeying the basic constraints \eq{I}-\eq{III}, is not a
1337: strongly oscillating function of momenta. Hence, regulators leading
1338: to a function
1339: %
1340: \beq\label{ell-gamma}
1341: \ell_\gamma(\omega)\sim(1+\omega)^{1-\gamma}
1342: \eeq
1343: %
1344: define for $\gamma=2$ a boundary in the space of regulators. This is
1345: the case for $R_{\rm opt}$. No regulator can be found such that
1346: $\gamma>2$ \cite{Litim:2001hk}. The proportionality constant in
1347: \eq{ell-gamma} is irrelevant because it can be scaled away into the
1348: fields in \eq{FlowPotential}. The second boundary is set by a mass
1349: term regulator $R_{\rm mass}=k^2$: For a mass term, the corresponding
1350: flow is a Callan-Symanzik flow which is not a Wilsonian flow in the
1351: strict sense. This comes about because the condition \eq{II} does no
1352: longer hold true for $R_{\rm mass}$ in the limit $q^2\to\infty$. In
1353: higher dimensions, this may lead to an insufficiency in the
1354: integrating-out of large momentum modes. Inserting $R_{\rm mass}$
1355: into \eq{Id} in $d=3$ dimensions leads to \eq{ell-gamma} with
1356: $\gamma=\s012$. This sets the second boundary. The regulator
1357: $R_\gamma$ is explicitly known for the cases $\gamma=2,\s032,1$ and
1358: $\s012$, and corresponds, respectively, to the optimised regulator,
1359: the quartic regulator $R_{\rm quart}=k^4/q^2$, the sharp cutoff and a
1360: mass term regulator $R_{\rm mass}=k^2$. For all other values of
1361: $\gamma\in [\s012,2]$, $R_\gamma$ can be recovered explicitly from
1362: $\ell_\gamma(\omega)$ \cite{Litim:2001hk}. In the present case, we
1363: only need to know that such regulators exist.\step
1364:
1365:
1366: We have computed the critical index $\nu_\gamma$ for
1367: $\ell_\gamma(\omega)$ with $\gamma\in[\s012,2]$, and our results are
1368: given in Tab.~2. In Fig.~12, the results are denoted by open circles
1369: for $x=\s023(\gamma-\s012)$. As a result, the function $\nu_\gamma$
1370: increases for decreasing $\gamma$, $\nu_\gamma\ge\nu_{\rm opt}$. In
1371: particular, once $\gamma<1$, the results obey $\nu_\gamma>\nu_{\rm
1372: sharp}$. When $\gamma\to\s012$, the eigenvalues at criticality
1373: approach their large-$N$ values $\nu\to 1$ and $\omega_n\to 2n-1$ for
1374: any $N$. Note that the large-$N$ limit is exact in that it is
1375: independent of the regularisation \cite{Litim:2001dt}. The large
1376: numerical value for $\nu_{\gamma=1/2}$ is due to the deficiencies of
1377: the Callan-Symanzik flow. We conjecture that the large-$N$ limit
1378: $\nu_{{}_{\, {\rm large}\,N}}$ corresponds, for any $N$, to an upper
1379: boundary for {any} regulator
1380: %
1381: \beq
1382: \nu_{{}_{\rm ERG}}\le \nu_{{}_{\, {\rm large}\,N}} = 1\,.
1383: \eeq
1384: %
1385:
1386:
1387: %********|*********|*********|*********|*********|*********|*********|****
1388: \subsection{Lower boundary}
1389: %********|*********|*********|*********|*********|*********|*********|****
1390:
1391: Next, we assess the lower boundary. It would be important to know whether
1392: the optimised regulator leads to the smallest attainable value for
1393: $\nu$ in the present approximation. We are not aware of a general
1394: proof for this statement. However, strong evidence is provided by
1395: studying alterations of the optimised regulator. We have done so for
1396: various classes of regulators, three of which are discussed here more
1397: explicitly. For more details, we defer to the
1398: Appendices~\ref{VariantsOpt} and \ref{SlidingOpt}. In
1399: Appendix~\ref{VariantsOpt}, we employ variants of the optimised
1400: regulator, given by the class $R_a$ of \eq{Ra}, and by the class $R_b$
1401: defined as
1402: %
1403: \beq\label{Rb}
1404: R_b(q^2)=
1405: (k^2-q^2)\,\theta(k^2-q^2)\,\theta(q^2-\s012 k^2)
1406: +(bk^2+(1-2b)q^2)\,\theta(\s012 k^2-q^2)\,.
1407: \eeq
1408: %
1409: for $0\leq b\leq\infty$. This class contains the optimised regulator
1410: \eq{Ropt} for $b=1$. The limit $b\to\infty$ corresponds to a variant
1411: of the standard sharp cutoff. Another new class of regulators $R_c$
1412: is studied in Appendix~\ref{SlidingOpt}, where we allow for an
1413: additional $k$-dependence on a mass scale within the regulator. We use
1414: the class
1415: %
1416: \bea\label{Rc}
1417: R_c(q^2) &=&
1418: (k^2-q^2-c\,m^2_k) \,\theta (k^2-q^2-c\,m^2_k)\,.
1419: \eea
1420: %
1421: Here, $c$ is a free parameter and $m^2_k=U_k''(\phi=0)$ is the mass
1422: term at vanishing field. For $c=0$, it turns into the optimised
1423: regulator \eq{Ropt}. Due to the implicit scale dependence of $m_k$ on
1424: $k$, the corresponding flow equations are substantially different from
1425: the usual one. Most notably, they contain terms proportional to the
1426: flow of $m_k$. \step
1427:
1428: The classes $R_\gamma$, $R_a$, $R_b$ and $R_c$ probe ``orthogonal''
1429: directions in the space of regulators. $R_\gamma$ is sensitive to the
1430: analyticity structure of the flow, $R_a$ and $R_b$ are sensitive to
1431: alterations of the function $r(y)$ in the low momentum regime, and
1432: $R_c$, while keeping the shape of the regulator $r(y)$ fixed, alters
1433: the implicit modifications due to an additional running mass term. As
1434: such, $R_a$, $R_b$ and $R_c$ can be seen as variants of the optimised
1435: regulator. The class $R_\gamma$ covers the largest domain of
1436: qualitatively different flows \cite{Litim:2001ky}.
1437:
1438: %********|*********|*********|*********|*********|*********|*********|****
1439: %********|*********| Fig13: pRelativ.ps |*********|*********|****
1440: %********|*********|*********|*********|*********|*********|*********|****
1441: \begin{figure}
1442: \begin{center}
1443: \unitlength0.001\hsize
1444: \begin{picture}(1000,550)
1445: \put(300,450){\fbox{{\Large $\displaystyle \0{\nu_{{}_{\rm ERG}}}{\nu_{\rm opt}}-1$}}}
1446: \put(238,72){\large $N$}
1447: \put(445,99){\large $\infty$}
1448: \put(90,200){\small Wegner-Houghton}
1449: \put(155,155){\small opt}
1450: \put(210,300){\small Callan-Symanzik}
1451: \put(520,200){
1452: \begin{minipage}{.47\hsize}
1453: { \small {\bf Figure 13:} The spread of the critical exponent $\nu$ for
1454: various $N$ to leading order in the derivative expansion. The
1455: upper bound is set by the large-$N$ limit (Callan-Symanzik flow).
1456: The sharp cut-off results (Wegner-Houghton flow) are given for
1457: comparison.}
1458: \end{minipage}}
1459: \hskip.02\hsize
1460: \psfig{file=pRelativ.ps,width=.45\hsize}
1461: \end{picture}
1462: \end{center}
1463: \vskip-1cm
1464: \end{figure}
1465: %********|*********|*********|*********|*********|*********|*********|****
1466:
1467: We have computed the critical exponent $\nu$ for the Ising
1468: universality class for all these classes of regulators. A part of our
1469: results for $N=1$ is displayed in Fig.~12. Similar results are found
1470: for all $N$. The classical (mean field) value for $\nu$ is $\nu_{\rm
1471: mf}=\s012$. The results for $R_a$ in \eq{Ra} are displayed in
1472: Fig.~12 by full circles, with $x=\s0a{a+1}$. Some numerical values are
1473: collected in Tab.~3 (cf.~Appendix~\ref{VariantsOpt}). It is found that
1474: all critical indices $\nu_a$ obey $\nu_{{}_{\, {\rm large}\,N}}\ge
1475: \nu_a\ge \nu_{\rm opt}>\nu_{\rm mf}$. This proves that alterations of
1476: the regulator in the low momentum region do not lead to values for
1477: $\nu$ smaller than $\nu_{\rm opt}$. An analogous result is found for
1478: all regulators $R_b$ in \eq{Rb}, where $\nu_{{}_{\, {\rm
1479: large}\,N}}\ge\nu_{b}\ge \nu_{\rm opt}$. Some numerical values
1480: are given in Tab.~4 (Appendix~\ref{VariantsOpt}). In Fig.~12, our
1481: results for $R_b$ are represented by open circles, and those for $R_c$
1482: by a dashed line, with $x\equiv c$. All regulators $R_c$ from \Eq{Rc}
1483: lead to the {\it same} critical exponent as $R_{\rm opt}$ in
1484: \Eq{Ropt}, $\nu_c\equiv \nu_{\rm opt}$ (see
1485: Appendix~\ref{SlidingOpt}). \step
1486:
1487:
1488: In Fig.~13, we discuss the spread of $\nu_{{}_{\rm ERG}}$ to leading
1489: order in the derivative expansion for all $N\ge 0$. The spread
1490: $\nu_{{}_{\,{\rm large}\,N}}/\nu_{\rm opt}-1$ is a $N$-dependent
1491: quantity. For $N=1$, the spread is about $0.54$, and hence quite
1492: large. For comparison, the relative width with respect to the sharp
1493: cut off $\nu_{\rm sharp}/\nu_{\rm opt}-1$ is roughly $0.06$ and
1494: significantly smaller. With increasing $N$, the spread vanishes as
1495: $\sim 1/N$. This follows trivially from the fact that the ERG flow, to
1496: leading order in the derivative expansion, becomes exact in the
1497: large-$N$ limit \cite{Litim:2001dt}.
1498: \step
1499:
1500: In summary, the critical exponent $\nu$, as a function of the infrared
1501: regularisation, is bounded. The upper boundary is realised for flows
1502: with a mass term regulator, e.g.~Callan-Symanzik flows. The lower
1503: boundary is given by
1504: %
1505: \beq \label{range}
1506: \nu_{{}_{\rm ERG}} \ge \nu_{{\rm opt}}
1507: \eeq
1508: %
1509: to leading order in the derivative expansion. Hence, $\nu_{{\rm opt}}$
1510: appears indeed to be the smallest value attainable within the present
1511: approximation. The inequality \eq{range} provides a quantitative basis
1512: for the optimisation procedure which has lead to $R_{\rm opt}$ in the
1513: first place. For the observable $\nu$, we have equally shown that the
1514: optimised regulator corresponds, at least, to a ``local minimum'' in
1515: the space of all regularisations. Furthermore, we have established
1516: flat directions in the space of regulators. Based on the conceptual
1517: reasons which have lead to $R_{\rm opt}$
1518: \cite{Litim:2000ci,Litim:2001up}, we expect that $\nu_{\rm opt}$ even
1519: corresponds to the global minimum. At present, however, we have no
1520: regulator-independent proof for this conjecture.
1521:
1522:
1523: %********|*********|*********|*********|*********|*********|*********|****
1524: \section{Discussion and conclusions}\label{Discussion}
1525: %********|*********|*********|*********|*********|*********|*********|****
1526:
1527: We studied in detail $O(N)$ symmetric scalar theories at criticality,
1528: using the ERG method to leading order in the derivative expansion.
1529: This included a complete investigation of the Gaussian fixed point in
1530: $d>2$ for arbitrary regulator, and the computation of universal
1531: critical exponents and subleading corrections at the Wilson-Fisher
1532: fixed point for $d=3$. Furthermore, we studied the spurious scheme
1533: dependence for the critical exponent $\nu$ at the Wilson-Fisher fixed
1534: point in three dimensions. One of the main new results is that the
1535: leading critical exponent, as a function of the IR regulator, is
1536: bounded from above and from below as
1537: %
1538: \beq\label{FinalBound}
1539: \nu_{{\, {\rm opt}}} \le
1540: \nu_{{}_{\rm ERG}}\le
1541: \nu_{{}_{\, {\rm large}\,N}}\,.
1542: \eeq
1543: %
1544: This result has been achieved by studying the maximal domain of ERG
1545: flows in the present approximation, ranging from Callan-Symanzik flows
1546: to optimised flows and variants thereof. The {qualitative} result --
1547: the existence of a non-trivial Wilson-Fisher fixed point -- is very
1548: stable, although the spread of values for $\nu$ is fairly large (see
1549: Fig.~13). Supposedly, this is a consequence of a small anomalous
1550: dimension $\eta$, which constrains higher order corrections. The
1551: spread would shrink to zero only to sufficiently high order in the
1552: derivative expansion or in the large-$N$ limit.\step
1553:
1554: The important {quantitative} question is: Which value for $\nu$ could
1555: be considered as a good approximation to the physical theory? In view
1556: of the regulator dependence, a prediction solely based on
1557: \eq{FinalBound} is of little use. Our answer to this problem is
1558: entirely based upon the structure of the ERG flow. We proposed to use
1559: specific regulators which lead to more stable ERG flows in the space
1560: of all action functionals. The numerical determination of critical
1561: exponents and subleading corrections to scaling as given in Tab.~1, is
1562: based on an optimised flow. We expect that the results should be in
1563: the vicinity of the physical theory. In the present approximation,
1564: the results for $\nu_{\rm opt}$ are indeed closest to the physical
1565: ones \cite{Zinn-Justin:1989mi},
1566: %
1567: \beq\label{nu_Phys}
1568: \nu_{{\rm phys}}<\nu_{{\, {\rm opt}}}\,.
1569: \eeq
1570: %
1571: The understanding of the spurious scheme dependence reduced the
1572: ambiguity in $\nu$ to a small range about $\nu_{\rm opt}$. Typically,
1573: the results from optimised flows other than $R_{\rm opt}$ are close to
1574: the values achieved by $R_{\rm opt}$, and hence close to the lower
1575: boundary of \eq{FinalBound}. In this light, optimised flows are
1576: solutions to a ``minimum sensitivity condition'' in the space of all
1577: IR regularisations \cite{Litim:2001fd}. \step
1578:
1579:
1580: Next we turn to the Callan-Symanzik flow, which is the flow with a
1581: mass term regulator $R_{\rm mass}=k^2$. We argued that it defines the
1582: upper boundary of values for $\nu$. It is quite remarkable that this
1583: flow, for {any} $N$, leads to the same eigenvalues at criticality
1584: given by the large-$N$ result. Here, this result has been achieved
1585: numerically. It would be helpful if it could also be understood
1586: analytically. The large numerical value for $\nu$ reflects the poor
1587: convergence properties of a Callan-Symanzik flow, essentially due to
1588: deficiencies in the integrating-out of large momentum modes. \step
1589:
1590: Now we discuss our results concerning polynomial approximations. The
1591: reliability of this additional truncation is guaranteed if the
1592: approximation convergences reasonably fast. Here, we have established
1593: that optimised flows converge very rapidly within the local potential
1594: approximation. The efficiency is remarkable: a simple approximation
1595: with only six independent operators --- say, the running v.e.v.~and
1596: five running vertex functions up to $(\phi^a\phi_a)^5$ --- reproduces
1597: the physical result for the exponent $\nu$ at the percent level. A
1598: better agreement cannot be expected, given that anomalous dimensions
1599: of the order of a few percent have been neglected. These findings are
1600: in contrast to earlier computations based on the sharp cut-off, where
1601: the polynomial approximation has lead to spurious fixed point
1602: solutions, even to high order in the approximation
1603: \cite{Margaritis:1988hv}. Hence, the efficiency of the formalism
1604: not only depends on the choice for the degrees of freedom and the
1605: truncation, but additionally, and strongly, on the IR regulator. We
1606: conclude that polynomial approximations are reliable for all technical
1607: purposes, and even to low orders, if they are backed-up by appropriate
1608: regulators. These considerations should be useful in more complex
1609: theories whose algebraic complexity requires polynomial
1610: approximations, e.g.~quantum gravity.\step
1611:
1612: It would be interesting to apply the present ideas to theories like
1613: QCD, where the propagating modes are strongly modified in the low
1614: momentum regime due to confinement \cite{QCD1,Ellwanger:1996qf}. Then,
1615: an optimised regulator is found by requiring that the regularised
1616: inverse propagator is again flat, i.~e.~momentum-independent for small
1617: momenta. Interesting choices are $R = (k^2-X)\,\theta(k^2-X)$ and $X =
1618: \Gamma_k^{(2)}[\phi=\phi_0]$ and variants thereof. Here $\phi_0$
1619: denotes a non-propagating background field. This conjecture is
1620: supported by first results. \\[1ex]
1621:
1622: %********|*********|*********|*********|*********|*********|*********|****
1623: {\bf Acknowledgments:}
1624: %********|*********|*********|*********|*********|*********|*********|****
1625: I thank J.M.~Pawlowski for useful discussions and comments on the
1626: manuscript. This work has been supported by the European Community
1627: through the Marie-Curie fellowship HPMF-CT-1999-00404.\step
1628:
1629: %********|*********|*********|*********|*********|*********|*********|****
1630: \setcounter{section}{0}
1631: \renewcommand{\thesection}{\Alph{section}}
1632: \renewcommand{\thesubsection}{\arabic{subsection}}
1633: \renewcommand{\theequation}{\Alph{section}.\arabic{equation}}
1634: %********|*********|*********|*********|*********|*********|*********|****
1635:
1636: %********|*********|*********|*********|*********|*********|*********|****
1637: \section{Gaussian fixed point}\label{Gauss}
1638: %********|*********|*********|*********|*********|*********|*********|****
1639:
1640: In this appendix, we discuss the Gaussian fixed point of the flow
1641: equation \eq{FlowPotentialOpt} in $d>2$ dimensions and for arbitrary
1642: regulator. The Gaussian fixed point corresponds to the specific
1643: solution $u_\star(\rho)={\rm const.}$ All higher derivatives of the
1644: potential vanish, $u^{(n)}_\star(\rho)=0$. From the flow equation, we
1645: deduce that
1646: %
1647: \beq\label{GFP:1}
1648: u_\star=\0{2v_d}{d} N \int^\infty_0 dy\, y^{\s0d2 -1}\0{-r'(y)}{1+r(y)}\,.
1649: \eeq
1650: %
1651: For the optimised regulator, we find $u_\star=4Nv_d/d^2$. More
1652: information can be extracted by studying small perturbations $\delta
1653: u^{(m)}$ around the $m$-th derivative of the scaling solution
1654: $u_\star(\rho)^{(m)}$. The eigenperturbations obey the differential
1655: equation
1656: %
1657: \beq\label{GFP:EW}
1658: \partial_t\, \delta u^{(m)}=\omega\, \delta u^{(m)}
1659: \eeq
1660: %
1661: with eigenvalues $\omega$. Expanding the flow equation to leading
1662: order in $\delta u$, the eigenvalues obey
1663: %
1664: \beq\label{GFP:2}
1665: 0=
1666: \left[\omega+d-(d-2)m\right]\, \delta u^{(m)}
1667: +\left[2A(\0N2+m)-(d-2)\rho\right]\,\delta u^{(m+1)}
1668: +2\rho A \,\delta u^{(m+2)}\,.
1669: \eeq
1670: %
1671: Here, the scheme-dependent coefficient $A$ is given by
1672: %
1673: \beq
1674: A= 2v_d\int^\infty_0 dy \,y^{\s0d2 -3}\0{-r'(y)}{[1+r(y)]^2}
1675: \eeq
1676: %
1677: and $0<A<\infty$. For the optimised regulator, $A_{\rm opt}=4v_d/d$.
1678: Introducing new variables $x=(d-2)\rho/(2A)$ and $f(x)=\delta
1679: u(\rho)$, the differential equation \eq{GFP:2} transforms into the
1680: (generalised) Laguerre differential equation
1681: %
1682: \beq\label{GFP:Laguerre}
1683: 0=
1684: \left(\0{d+\omega}{d-2}-m\right) f^{(m)}(x)
1685: +\left(\0N2 +m-x\right)f^{(m+1)}(x)
1686: + x f^{(m+2)}(x)\,.
1687: \eeq
1688: %
1689: We consider only polynomial solutions to \Eq{GFP:Laguerre}. The
1690: requirement that solutions to \Eq{GFP:Laguerre} are bounded by
1691: polynomials fixes the possible eigenvalues as
1692: %
1693: \beq\label{GFP:omega}
1694: \omega=(d-2)(n+m)-d
1695: \eeq
1696: %
1697: for non-negative integers $n$ and $m$.
1698: %Within the ERG approach,
1699: %non-polynomial solutions to \Eq{GFP:Laguerre} have been studied in
1700: %\cite{Gies:2001xr}. Here, they are not discussed any further.
1701: Apart
1702: from an irrelevant normalisation constant, the $n$-th eigensolution to
1703: \Eq{GFP:Laguerre} are given by the (generalised) Laguerre polynomials
1704: %
1705: \beq\label{GFP:LPoly}
1706: \delta u^{(m)}(\rho)=
1707: L^{m-1+N/2}_{n}\left(\0{2A\,\rho}{d-2}\right)\,.
1708: \eeq
1709: %
1710: \Eq{GFP:LPoly} is the most general eigensolution at the Gaussian fixed
1711: point in $d>2$ dimensions with eigenvalues given by \Eq{GFP:omega}.
1712: The result holds for arbitrary regulator function. The scheme
1713: dependence enters only the argument of the Laguerre polynomials in
1714: \Eq{GFP:LPoly}. It is interesting to note that the eigenvalues are
1715: independent of the regularisation scheme. Furthermore, the rescaled
1716: differential equation \eq{GFP:Laguerre} is also independent of the
1717: regulator.\step
1718:
1719: In $d=3$ dimensions, the relevant and marginal operators are
1720: $L^{N/2-1}_0$, $L^{N/2-1}_1$, $L^{N/2-1}_2$ and $L^{N/2-1}_3$ with
1721: eigenvalues $\omega=-3,-2,-1$ and $0$ for $m=0$, or $L^{N/2}_0$,
1722: $L^{N/2}_1$ and $L^{N/2}_2$ with eigenvalues $\omega=-2,-1$ and $0$
1723: for $m=1$. In $d=4$ dimensions, the relevant and marginal operators
1724: are $L^{N/2-1}_0$, $L^{N/2-1}_1$ and $L^{N/2-1}_2$ with eigenvalues
1725: $\omega=-4,-2$ and $0$ for $m=0$, or $L^{N/2}_0$ and $L^{N/2}_1$ with
1726: eigenvalues $\omega=-2$ and $0$ for $m=1$. For $m\ge \s0d{d-2}$, all
1727: eigenoperators are marginal or irrelevant.\step
1728:
1729:
1730: For $2m+N=1$ $(2m+N=3)$, the solution \Eq{GFP:LPoly} can be rewritten
1731: in terms of Hermite polynomials of even (odd) degree,
1732: %
1733: \bea \label{GFP:m=0}
1734: m=\s0{1-N}{2}:\quad \delta u^{(m)}(\rho)&=&
1735: \0{(-1)^n}{n!}\,2^{-2n}\,
1736: H_{2n}\left(\sqrt{\0{2A\rho}{d-2}}\right)\\
1737: \label{GFP:m=1}
1738: m=\s0{3-N}{2}:\quad \delta u^{(m)}(\rho)&=&
1739: \0{(-1)^n}{n!}\,2^{-2n-1}\,
1740: \sqrt{\0{d-2}{2A\rho}}\,
1741: H_{2n+1}\left(\sqrt{\0{2A\rho}{d-2}}\right)\,.
1742: \eea
1743: %
1744: Let us finally mention that some of these solutions have been given
1745: earlier in the literature for the case of a sharp cutoff regulator
1746: (with $A_{\rm sharp}=1$): for $N=1$ and $m=1$, \Eq{GFP:m=1} has been
1747: given in Ref.~\cite{Hasenfratz:1986dm}, and \Eq{GFP:LPoly} for $m=1$
1748: has been given in Ref.~\cite{Comellas:1997tf}.
1749:
1750: %********|*********|*********|*********|*********|*********|*********|****
1751: \section{Variants of the optimised cutoff}\label{VariantsOpt}
1752: %********|*********|*********|*********|*********|*********|*********|****
1753:
1754: In this appendix, we discuss variants of the optimised regulator
1755: \eq{Ropt}. The aim is to probe whether certain alterations of the
1756: regulator may lead to lower values for the critical exponent $\nu$.
1757: Here, the properties of the regularisation are changed in the low
1758: momentum region by modifying the function $r(y)$. In the following
1759: appendix, we discuss modifications of $r(y)$ through the introduction
1760: of additional (theory-dependent) $k$ dependent parameters.
1761: %\step
1762:
1763: %********|*********|*********|*********|*********|*********|*********|****
1764: %********|*********| Fig14: pRopta.ps |*********|*********|****
1765: %********|*********|*********|*********|*********|*********|*********|****
1766: \begin{figure}
1767: \begin{center}
1768: \vskip-1cm
1769: \unitlength0.001\hsize
1770: \begin{picture}(1000,550)
1771: \put(250,250){\framebox{\large $y(1+r_a)$}}
1772: \put(220,75){\large $y$}
1773: \hskip.05\hsize
1774: \put(520,200){
1775: \begin{minipage}{.47\hsize}
1776: {\small {\bf Figure 14:} The function $y(1+r_a)$ for the class of
1777: regulators \eq{Radef}. All lines for different $a$ coincide for large
1778: momenta $y\in [1,\infty]$ (full line), but differ for small
1779: momenta $y\in [0,1[$ (dashed lines). The dashed lines clock-wise
1780: from the bottom: $a=\s012, 1, 2$ and $\infty$. The short dashed
1781: line corresponds to $a=0$ (no regulator). }
1782: \end{minipage}}
1783: \psfig{file=pRa.ps,width=.45\hsize}
1784: \end{picture}
1785: \end{center}
1786: \vskip-2cm
1787: \end{figure}
1788: %********|*********|*********|*********|*********|*********|*********|****
1789:
1790: %********|*********|*********|*********|*********|*********|*********|****
1791: \subsection{Definition}
1792: %********|*********|*********|*********|*********|*********|*********|****
1793:
1794: In this appendix, we discuss two variants of the optimised regulator
1795: \eq{Ropt}. First, we consider the class of regulators given by
1796: %
1797: \beq\label{Radef}
1798: R_a(q^2)=a\,(k^2-q^2)\,\theta(k^2-q^2)\,.
1799: \eeq
1800: %
1801: These regulators have a compact support. They vanish for all
1802: $q^2>k^2$. In the infrared, they have a mass-like limit $R_a(q^2\to
1803: 0)= ak^2$ for all $0<a<\infty$. The limit $a\to\infty$ corresponds to
1804: the sharp cut-off case. For $a=0$, the regulator is removed
1805: completely. For $a=1$, the regulator \eq{Radef} reduces to the optimised
1806: regulator \eq{Ropt}. As a function of $a$, the regulators differ only
1807: in the momentum regime $y\in [0,1[\ $, where $y\equiv q^2/k^2$. The
1808: dimensionless functions $r_a$ corresponding to \eq{Radef} are given by
1809: %
1810: \beq
1811: r_a(y)=a\, (\s01y-1)\,\theta(1-y)\,.
1812: \eeq
1813: %
1814: In Fig.~1, we have displayed the function $y(1+r_a)$ for various
1815: cases. The full line corresponds to the range $y\in [1,\infty]$, for
1816: all regulators \eq{Radef}. The dashed lines, clock-wise from the bottom,
1817: correspond to $a=\s012, 1, 2$ and $\infty$. The gaps associated to
1818: \eq{Radef} are given by $C_a=\s0{a}{2a-1}$ for $a\ge 1$, and $C_a=a$ for
1819: $\s012<a<1$. They are obtained from the normalised analogue of \eq{Radef},
1820: chosen such that $r_a(\s012)=1$.
1821: \step
1822:
1823: Second, we consider another variant of the optimised regulator, where
1824: the properties of the regularisation are changed only in the low
1825: momentum region by modifying the function $r(y)$. Consider the class
1826: of regulators given by
1827: %
1828: \beq\label{Rbdef} R_b(q^2)=
1829: (k^2-q^2)\,\theta(k^2-q^2)\,\theta(q^2-\s012 k^2)
1830: +(bk^2+(1-2b)q^2)\,\theta(\s012 k^2-q^2)\,. \eeq
1831: %
1832: These regulators have a compact support. They vanish for all
1833: $q^2>k^2$. In the infrared, they have a mass-like limit $R_b(q^2\to
1834: 0)= b\,k^2$ for all $0<b<\infty$. At first sight, it may seem that
1835: \eq{Rbdef} is not a viable regulator for $b=0$, because $R_{b=0}$
1836: vanishes in the infrared limit (no gap). However, for $b=0$ the
1837: function $\partial_t R_{b=0}$ vanishes identically for all $q^2<\s012
1838: k^2$. Hence, \eq{Rbdef} provides a gap because
1839: $R_{b=0}(\s012k^2)=\s012k^2>0$. For $b=1$, the regulator \eq{Rbdef}
1840: reduces to the optimised regulator \eq{Ropt}. As a function of $b$,
1841: the regulators differ only in the momentum regime $y\in [0,\s012[\ $,
1842: where $y\equiv q^2/k^2$. The dimensionless functions $r_b$
1843: corresponding to \eq{Rbdef} are given by
1844: %
1845: \beq
1846: r_b(y)=
1847: (\s01y-1)\,\theta(1-y)\,\theta(y-\s012)
1848: +(\s0by+1-2b)\,\theta(\s012 -y)\,.
1849: \eeq
1850: %
1851: In Fig.~15, we have displayed the function $y(1+r_b)$ for various
1852: cases. The full line corresponds to the range $y\in [\s012,\infty]$,
1853: for all regulators \eq{Rbdef}. The dashed lines, clock-wise from the
1854: bottom, correspond to $b=0, \s012, 1, 2$ and $\infty$. By
1855: construction, the regulator is normalised as $r_b(\s012)=1$. The
1856: associated gaps are given by $C_b=1$ for $b\ge 1$ and $b=0$ (see
1857: below), and $C_b=b$ for $0<b<1$. For comparison, we have also given
1858: the curve for the standard sharp cutoff (dotted line). The
1859: corresponding gap is $C_{\rm sharp}=\s012$.
1860:
1861: %********|*********|*********|*********|*********|*********|*********|****
1862: %********|*********| Fig15: pRopta.ps |*********|*********|****
1863: %********|*********|*********|*********|*********|*********|*********|****
1864: \begin{figure}
1865: \begin{center}
1866: \vskip-.5cm
1867: \unitlength0.001\hsize
1868: \begin{picture}(1000,550)
1869: \put(250,250){\framebox{\large $y(1+r_b)$}}
1870: \put(220,75){\large $y$}
1871: \hskip.05\hsize
1872: \put(520,210){
1873: \begin{minipage}{.47\hsize}
1874: {\small {\bf Figure 15:} The function $y(1+r_b)$ for the class of
1875: regulators \eq{Rbdef}. All lines for different $b$ coincide for
1876: large momenta $y\in [\s012,\infty]$ (full line), but differ for
1877: small momenta $y\in [0,\s012]$ (dashed lines). The dashed lines
1878: clock-wise from the bottom: $b=0, \s012, 1, 2$ and $\infty$. The
1879: standard sharp cutoff regulator (dotted line) is given for
1880: comparison. }
1881: \end{minipage}}
1882: \psfig{file=pRopta.ps,width=.45\hsize}
1883: \end{picture}
1884: \end{center}
1885: {}\vskip-1.5cm
1886: \end{figure}
1887: %********|*********|*********|*********|*********|*********|*********|****
1888:
1889: The regulators $R_a$ and $R_b$ have a similar low momentum limit for
1890: $q^2\to 0$, e.g.~$R_a(q^2\to 0)=R_{b=a}(q^2\to 0)=a\,k^2$. The crucial
1891: difference between them concerns the intermediate momentum regime
1892: $q^2\approx k^2$. Here, the regulator $R_b$ leads by construction to a
1893: plateau for $y(1+r_b)$, which is absent for $R_a$.
1894:
1895: %********|*********|*********|*********|*********|*********|*********|****
1896: \subsection{Flows}
1897: %********|*********|*********|*********|*********|*********|*********|****
1898:
1899: In order to employ \eq{Radef} for the computation of critical exponents
1900: in $d=3$, the associated flows $\ell(\omega)\to \ell_a(\omega)$ have
1901: to be computed. For the function $\ell_a(\omega)$ and for $a\le 1$, we
1902: find in $d=3$ (similar expressions are found for any $d$)
1903: %
1904: \beq
1905: \label{la-}
1906: \ell_a(\omega)=
1907: \0{2a}{1-a}\left[1-\sqrt{\0{a+\omega}{{1-a}}}
1908: \arctan\sqrt{\0{1-a}{a+\omega}}\right]\,.
1909: \eeq
1910: %
1911: The region $a>1$ is obtained by analytical continuation:
1912: %
1913: \beq
1914: \label{la+}
1915: \ell_a(\omega)=
1916: \0{2a}{1-a}\left[1-\sqrt{\0{{a+\omega}}{{a-1}}}
1917: {\rm arctanh}\sqrt{\0{a-1}{a+\omega}}\right]\,.
1918: \eeq
1919: %
1920: These flows have the following structure. They have poles on the
1921: negative $\omega$-axis at $\omega=-1$. For $a=\infty$, the function
1922: decays only logarithmically for asymptotically large $\omega$. In the
1923: limiting cases $a=1$ and $\infty$, we find
1924: %
1925: \bea
1926: \label{la1}
1927: \ell_{a=1}(\omega)&=& \s023 (1+\omega)^{-1}\\
1928: \label{laInf}
1929: \ell_{a=\infty}(\omega)&=&-\ln(1+\omega)+{\rm const.}
1930: \eea
1931: %
1932: For $a=1$, in \eq{la1}, both expressions \eq{la-} and \eq{la+} have
1933: the same limit discussed earlier in \cite{Litim:2001up}. In the limit
1934: $a\to\infty$, the resulting regulator is equivalent to the standard
1935: sharp cut-off.
1936: \step
1937:
1938: In order to employ the regulator $R_b$ from \eq{Rbdef} for the
1939: computation of critical exponents in $d=3$, the associated flows
1940: $\ell(\omega)\to \ell_b(\omega)$ have to be computed. In full analogy
1941: to the preceding computation, we find in $d=3$ (similar expressions
1942: are found for any $d$) for the function $\ell_b(\omega)$ and for $b\le
1943: 1$
1944: %
1945: \beq
1946: \label{lopta-}
1947: \ell_b(\omega)=
1948: \s023 (1-2^{-3/2})(1+\omega)^{-1}
1949: +\0{b}{\sqrt{2}(1-b)}\left[1-\sqrt{\0{b+\omega}{{1-b}}}
1950: \arctan\sqrt{\0{1-b}{b+\omega}}\right]\,.
1951: \eeq
1952: %
1953: The region $b>1$ is obtained by analytical continuation:
1954: %
1955: \beq
1956: \label{lopta+}
1957: \ell_b(\omega)=
1958: \s023 (1-2^{-3/2})(1+\omega)^{-1}
1959: +\0{b}{\sqrt{2}(1-b)}\left[1-\sqrt{\0{{b+\omega}}{{b-1}}}
1960: {\rm arctanh}\sqrt{\0{b-1}{b+\omega}}\right]\,.
1961: \eeq
1962: %
1963: These flows have the following structure. They have poles on the
1964: negative $\omega$-axis at $\omega=-1$ due to the first term on the
1965: r.h.s.~in \eq{lopta-} and \eq{lopta+}. For large $\omega$ and $b<\infty$,
1966: both expressions decay as $\omega^{-1}$. For $b=\infty$, the decay is
1967: only logarithmic. Let us consider the three limiting cases $a=0, 1$
1968: and $\infty$. We find
1969: %
1970: \bea
1971: \label{lopta0}
1972: \ell_{b=0}(\omega)&=&\s023 (1-2^{-3/2})(1+\omega)^{-1}\\
1973: \label{lopta1}
1974: \ell_{b=1}(\omega)&=& \s023 (1+\omega)^{-1}\\
1975: \label{loptaInf}
1976: \ell_{b=\infty}(\omega)&=&\s023 (1-2^{-3/2})(1+\omega)^{-1}
1977: -\s01{2\sqrt{2}}\ln(1+\omega)+{\rm const.}
1978: \eea
1979: %
1980: For $b=0$, the momentum regime $q^2<\s012 k^2$ does not contribute to
1981: the flow and the effective gap for $b=0$ is $C_0=1$. This is seen
1982: directly from \eq{lopta-} and \eq{lopta0}: the first term of
1983: \eq{lopta-} stems from the momentum interval $\s012 k^2\leq q^2\leq
1984: k^2$, which is the only term surviving in \eq{lopta0}. For $b=1$, in
1985: \eq{lopta1}, both terms of \eq{lopta-} combine to the known result
1986: discussed earlier in \cite{Litim:2001up}. Finally, we turn to the
1987: limit $b\to\infty$. The resulting regulator is similar to the
1988: standard sharp cut-off, with, however, an important difference. For
1989: the sharp cutoff, the function $y(1+r_b)$ has no plateau in the
1990: momentum regime $\s012 k^2\leq q^2\leq k^2$ (see Fig.~15), which leads
1991: to $\ell_{\rm sharp}(\omega)=-\ln(1+\omega)$. In \eq{laInf}, the
1992: sharp-cutoff-like logarithmic term is clearly seen, and is due to the
1993: momentum integration with $y\in [0,\s012]$. However, a decisive
1994: difference is the additional term in \eq{laInf}. Notice also that the
1995: constant in \eq{loptaInf} is actually infinite, but field independent.
1996: Hence, it is irrelevant for a computation of critical exponents (only
1997: the functions $\partial_\omega \ell_b(\omega)$ are needed). \step
1998:
1999: %********|*********|*********|*********|*********|*********|*********|****
2000: \subsection{Results}
2001: %********|*********|*********|*********|*********|*********|*********|****
2002:
2003:
2004: For $R_a$, we have computed the critical exponent $\nu$ for the Ising
2005: universality class using the flow equation \eq{FlowPotential} with
2006: \eq{la-}, \eq{la+} (see Tab.~3). We confirm that
2007: $\nu_{a=\infty}=\nu_{\rm sharp}$ and that $\nu_{a=1}=\nu_{\rm opt}$.
2008: For all $a>1$ $(a<1)$, $\nu_a$ is a monotonically increasing
2009: (decreasing) function with increasing $a$, hence $\nu_a\ge\nu_{\rm
2010: opt}$. Notice that the smallest value for $\nu$ is obtained for the
2011: largest value for the gap parameter. \step
2012:
2013: %********|*********|*********|*********|*********|*********|*********|****
2014: %********|*********| Tab3 |*********|*********|*********|*********|****
2015: %********|*********|*********|*********|*********|*********|*********|****
2016: \begin{center}
2017: %\begin{tabular}{c||lllll}
2018: \begin{tabular}{c||c|c|c|c|c|c|c|c|c|c}
2019: $a$&
2020: $10^{-2}$&
2021: $10^{-1}$&
2022: $\s012$&
2023: 1&
2024: 2&
2025: $10$&
2026: $10^{2}$&
2027: $10^{3}$&
2028: $10^{4}$&
2029: $\infty$\\[.5ex] \hline%\\[-3ex]
2030: $C_a$&
2031: $10^{-2}$&
2032: $10^{-1}$&
2033: $\s012$&
2034: 1&
2035: $\s023$&
2036: $\s0{10}{19}$&
2037: $\s0{100}{199}$&
2038: $\s0{1000}{1999}$&
2039: $\s0{10000}{19999}$&
2040: $\s012$\\[.5ex] \hline%\\[-1.5ex]
2041: $\quad\nu_a\quad$&
2042: $\ .776\ {}$&
2043: $\ .677\ {}$&
2044: $\ .652\ {}$&
2045: $\ .650\ {}$&
2046: $\ .651\ {}$&
2047: $\ .665\ {}$&
2048: $\ .683\ {}$&
2049: $\ .688\ {}$&
2050: $\ .689\ {}$&
2051: $\ .690\ {}$\\
2052: \end{tabular}
2053: \end{center}
2054: \begin{center}
2055: \begin{minipage}{.95\hsize}
2056: \vskip.3cm {\small {\bf Table 3:} Critical exponents $\nu$ (Ising
2057: universality class) for the regulator $R_a$ and various $a$.}
2058: \end{minipage}
2059: \end{center}
2060: %********|*********|*********|*********|*********|*********|*********|****
2061:
2062: %********|*********|*********|*********|*********|*********|*********|****
2063: %********|*********| Tab4 |*********|*********|*********|*********|****
2064: %********|*********|*********|*********|*********|*********|*********|****
2065: \begin{center}
2066: \begin{tabular}{c||c|c|c|c|c|c}
2067: $b$&
2068: $0$&
2069: $\s012$&
2070: $1$&
2071: $10$&
2072: $100$&
2073: $\infty$
2074: \\[.5ex] \hline%\\[-3ex]
2075: $\quad\nu_b\quad$&
2076: $\ .6495\ {}$&
2077: $\ .6518\ {}$&
2078: $\ .6495\ {}$&
2079: $\ .6594\ {}$&
2080: $\ .6675\ {}$&
2081: $\ .6699\ {}$\\
2082: \end{tabular}
2083: \end{center}
2084: \begin{center}
2085: \begin{minipage}{.95\hsize}
2086: \vskip.3cm {\small {\bf Table 4:} Critical exponents $\nu$ (Ising
2087: universality class) for the flows with $R_b$ and various
2088: $b$.}
2089: \end{minipage}
2090: \end{center}
2091: %********|*********|*********|*********|*********|*********|*********|****
2092:
2093:
2094: For $R_b$, we have also computed the critical exponent $\nu$ for the
2095: Ising universality class using the flow equation \eq{FlowPotential}
2096: with \eq{lopta-}, \eq{lopta+}. We find that
2097: $\nu_{b=0}=\nu_{b=1}=\nu_{\rm opt}$. For $b<1$, we have
2098: $\nu_b>\nu_{\rm opt}$. However, for too small values of $b$, $1\gg
2099: b>0$, the regulator leads to a very small gap and the polynomial
2100: approximation does no longer converge to a definite result, which is
2101: an artifact of the regulator. For $b> 1$, $\nu_b$ is a monotonic
2102: function of $b$ with $\nu_{b=\infty}\ge\nu_b>\nu_{\rm opt}$. It is
2103: interesting to discuss the case $b=\infty$ in more detail. Here,
2104: $\nu_{b=\infty}=.6699\cdots$ which should be compared to $\nu_{\rm
2105: sharp}=.6895\cdots$ and to $\nu_{\rm opt}=.6495\cdots$. From the
2106: structure of the flow, it is clear that the difference $\nu_{\rm
2107: sharp}-\nu_{b=\infty}$ has to be attributed to momenta with $y\in
2108: [\s012,1]$. Hence, the ``flattening'' of the standard sharp cutoff
2109: reduces the critical exponent by a few percent. On the technical
2110: level, this reduction is attributed to the non-logarithmic term in
2111: \eq{loptaInf}. However, the smallest value for $\nu$ is obtained only
2112: in case the logarithmic term is absent, as it happens in both
2113: \eq{lopta0} and \eq{lopta1}.
2114:
2115:
2116: %********|*********|*********|*********|*********|*********|*********|****
2117: \section{Cutoffs with intrinsic scaling}\label{SlidingOpt}
2118: %********|*********|*********|*********|*********|*********|*********|****
2119:
2120: In this appendix, we study regulators with additional intrinsic
2121: scale-dependent parameters.
2122:
2123: %********|*********|*********|*********|*********|*********|*********|****
2124: \subsection{Definition}
2125: %********|*********|*********|*********|*********|*********|*********|****
2126:
2127: Up to now, we have considered IR regulators $R(q^2)$ which depend on
2128: momenta only through the combination $q^2/k^2$, cf.~\Eq{r}. In
2129: \Eq{r}, the essential IR cutoff is provided by the function
2130: $r(q^2/k^2)$, which cuts off the momentum scale $q^2$ in a way which
2131: is independent of the particular theory studied. The $k$ dependence of
2132: $R(q^2)$ is the trivial $k$ dependence linked to the dimensionality of
2133: $R$. The situation is different once further $k$-dependent functions
2134: are introduced into the regulator. Typically, this is done by
2135: replacing
2136: %
2137: \beq\label{Rm} R(q^2) \to R(q^2,\{Z_k, m^2_k,\ldots\})\,. \eeq
2138: %
2139: Here, the set $\{Z_k, m^2_k,\ldots\}$ denotes scale-dependent
2140: parameters of the specific theory studied, like the wave-function
2141: renormalisation $Z_k$ or mass parameters $m_k$. It is expected that
2142: the substitution \eq{Rm} leads to an improved convergence and
2143: stability of the flow. A well-known example for \eq{Rm} is given by
2144: $R(q^2) \to Z_k R(q^2)\,,$ which is often used beyond the leading
2145: order in a derivative expansion. Here, the introduction of $Z_k$ in
2146: the regulator simplifies the study of scaling solutions. In the
2147: context of non-Abelian gauge theories, regulators like \Eq{Rm} have
2148: been used in \cite{Ellwanger:1996qf} based on the replacement $q^2\to
2149: \bar\Gamma^{(2)}_k[\phi_0]$ in the regulator, and hence
2150: %
2151: \beq\label{RGamma} R(q^2) \to R[\bar\Gamma^{(2)}_k]\,. \eeq
2152: %
2153: Again, this type of regulator has been motivated to stabilise the flow
2154: and to encompass possible poles in the flow due to mass terms for the
2155: gluonic fields \cite{Ellwanger:1996qf}. Notice that
2156: $\bar\Gamma^{(2)}_k$ in \Eq{RGamma} cannot be the full field-dependent
2157: functional, because elsewise the flow equation would no longer be the
2158: correct one. Instead, one has to evaluate it for some fixed
2159: background field $\phi=\phi_0$ (hence the bar on
2160: $\bar\Gamma^{(2)}_k$). In the present case, and to leading order in
2161: the derivative expansion, we have
2162: %
2163: \beq\label{Gamma0}
2164: \bar\Gamma_k^{(2)}[\phi_0]=q^2+U_k''(\phi_0)\,.
2165: \eeq
2166: %
2167: Here, $U''(\phi_0)$ corresponds to a scale-dependent effective mass
2168: term. Within the non-convex part of the potential, or in a phase with
2169: spontaneous symmetry breaking, we have $m^2_k\equiv U''(\phi_0)<0$.
2170: Hence the regulator \Eq{RGamma} is a special case of \Eq{Rm}. In the
2171: remainder, the optimised regulator \eq{ropt} is used to define a class
2172: of regulators of the form \eq{Rm}, namely
2173: %
2174: \bea\label{RMod}
2175: R_c(q^2) &=&
2176: (k^2-q^2-c\,m^2_k) \,\theta (k^2-q^2-c\,m^2_k)\,.
2177: \eea
2178: %
2179: Hence, $R_c(q^2)=R_{\rm opt}(q^2+c\,m_k^2)$. Effectively, this
2180: corresponds to the replacement $k^2\to k^2_{\rm eff}(k)=k^2+c\,m^2_k$.
2181: In terms of $r(y)$ defined in \Eq{r}, the regulator is given as
2182: %
2183: \bea
2184: r_c(y) &=& (\s0{1-c\,\bar\omega}{y}-1)
2185: \,\theta (1-y-c\,\bar\omega)\,,
2186: \eea
2187: %
2188: and the dimensionless mass parameter is
2189: %
2190: \beq \bar \omega \equiv m^2_k/k^2\,.\eeq
2191: %
2192: Below, we use $m^2_k\equiv U''(\phi_0=0)$. The regulator \eq{RMod} can
2193: be seen as a `sliding' cutoff, because at any scale $k$, only the
2194: $k$-dependent momentum interval $y\in [0,1-c\,\bar\omega_k]$
2195: contributes to the flow. For $c=0$, the regulator reduces to
2196: \eq{Ropt}, while for $c=1$ it turns into a regulator of the form
2197: \eq{RGamma} with \eq{Gamma0}. Due to the additional dependence on
2198: $m_k^2/k^2$, the structure of the flow equation is now different.
2199:
2200: %********|*********|*********|*********|*********|*********|*********|****
2201: \subsection{Flows}
2202: %********|*********|*********|*********|*********|*********|*********|****
2203:
2204: Let us compute the flow for the regulator \Eq{RMod}. Inserting
2205: \Eq{RMod} into \Eq{Id}, and after some straightforward algebra, we
2206: find
2207: %
2208: \beq\label{ell-mod} \ell_c(\omega)= \s02d (1-c\,\bar
2209: \omega)^{d/2} \,
2210: \0{1-c\,\bar\omega-\s0{c}{2}\partial_t\bar\omega}{1-c\,\bar
2211: \omega+\omega} \eeq
2212: %
2213: Notice that the function \Eq{ell-mod} depends now on $\bar \omega$ and
2214: on the {\it flow} of $\bar \omega$. This is generic to
2215: regulators of the form \eq{Rm} or \eq{RGamma}, because the implicit
2216: scale dependence of $m^2_k$ leads to an additional term $\sim
2217: \partial_t m^2_k$ in the flow equation. The flow equation
2218: \eq{FlowPotential} with \Eq{ell-mod} becomes
2219: %
2220: \bea\label{FlowPotentialMod}
2221: \partial_t u+du-(d-2) \rho u'
2222: &=&
2223: %\0{4v_d}{d}
2224: (N-1)(1-c\,\bar \omega)^{d/2} \, \0{1-c\,\bar
2225: \omega-\s0{c}{2} \partial_t \bar \omega}{1-c\,\bar
2226: \omega+u'}
2227: \nonumber\\
2228: &+&
2229: (1-c\,\bar \omega)^{d/2} \, \0{1-c\,\bar
2230: \omega-\s0{c}{2} \partial_t \bar \omega}{1-c\,\bar
2231: \omega+u'+2\rho u''} \, \eea
2232: %
2233: Here, in order to simplify the expressions, we have rescaled the
2234: irrelevant numerical factor $\s04d v_d$ into the fields and the
2235: potential. In order to find an explicit form of the flow, the running
2236: mass term needs to be specified. We chose $\bar \omega=u'(\rho=0)$.
2237: This choice is motivated by the fact that the function $u'$, on the
2238: fixed point, reaches its most negative value at vanishing field. For
2239: $c=0$, the original flow may run into a pole at $u'=-1$. The
2240: present choice shifts the pole to $u'(\rho)=-1+c\, u'(0)$. Since
2241: $u'(\rho)-u'(0)\ge 0$ for a scaling solution, the right-hand side of
2242: \eq{FlowPotentialMod} has no longer an explicit pole for $c=1$.
2243: This has been the motivation for the structure of the regulator used
2244: in \cite{Ellwanger:1996qf}. However, as we shall see, the full flow
2245: still has an implicit pole due to the flow of $\bar \omega$ in
2246: \eq{FlowPotentialMod}. In terms of
2247: %
2248: \beq\label{MassScales}
2249: \lambda_0\equiv u'(\rho=0)\,,\quad \lambda_1\equiv u''(\rho=0)\,,
2250: \eeq
2251: %
2252: and after inserting \Eq{ell-mod} into \Eq{FlowPotentialMod} and
2253: collecting terms proportional to the flow of $\lambda_0$, we find
2254: %
2255: \beq\label{flowmass} \partial_t \lambda_0
2256: =-\0{2\lambda_0+\lambda_1{(N+2){(1-c\,\lambda_0)^{1+d/2}}}{
2257: [1+\lambda_0(1-c)]^{-2}}}{1-\s012 c\lambda_1
2258: {(N+2){(1-c\,\lambda_0)^{d/2}}}{
2259: [1+\lambda_0(1-c)]^{-2}}}\,. \eeq
2260: %
2261: For the quartic coupling, and suppressing terms proportional to
2262: $u'''(0)$, we find
2263: %
2264: \bea \partial_t\lambda_1 &=& -\lambda_1 \left(4-d- {\lambda_1}\,
2265: \frac{2(N+8)\,{\left( 1 - c\,\lambda_0 \right) }^{d/2}}{
2266: {\left[1 + (1-c)\lambda_0 \right]}^3} \right)\times \nonumber \\
2267: \label{flowlambda}
2268: &&\left( 1 - c\,\lambda_0 +\0{c}{2} \frac{\lambda_0
2269: +\0{\lambda_1}{2} {\left( N+2 \right){\left( 1 -c\,\lambda_0
2270: \right) }^{1+d/2}}{ {\left[ 1 + \left( 1 -c \right)
2271: \,\lambda_0 \right]^{-2} }}}{ 1 - c\0{\lambda_1}{2}
2272: {(N+2)\left( 1 - c\,\lambda_0 \right)^{d/2}}{\left[ 1 +
2273: \left( 1 - c \right) \,\lambda_0 \right]^{-2}}} \right)
2274: \eea
2275: %
2276: The structure of the flows \eq{flowmass} and \eq{flowlambda} is easily
2277: understood. The denominator in \eq{flowmass} stems from a resummation
2278: of the back coupling of $\partial_t \lambda_0$. The denominator
2279: becomes trivial for $c=0$. The numerator contains the usual scaling
2280: term and a modified threshold behaviour, which depends now on $c$. A
2281: similar structure appears for the flow of the quartic coupling.
2282: Simpler forms of the flow are obtained for $c=0$ (no back-coupling of
2283: a running mass term) or $c=1$. For $c=0$, we have
2284: %
2285: \bea\label{flowmass0} \partial_t \lambda_0 &=&-2\lambda_0
2286: +\lambda_1 \0{N+2}{(1+\lambda_0)^{2}}\,,\\
2287: \label{flowlambda0}
2288: \partial_t\lambda_1 &=& -(4-d)\lambda_1 + \lambda_1^2\,
2289: \frac{2(N+8)}{(1 + \lambda_0 )^3}\,,
2290: \eea
2291: %
2292: while for $c=1$, the result is
2293: %
2294: \bea\label{flowmass1} \partial_t \lambda_0 &=& -\0{2\lambda_0
2295: +\lambda_1{(N+2){(1-\lambda_0)^{1+d/2}}}}{1-\s012 \lambda_1
2296: {(N+2){(1-\lambda_0)^{d/2}}}} \,, \\
2297: \label{flowlambda1}
2298: \partial_t\lambda_1 &=& -\lambda_1 (1+ \lambda_0)\frac{4-d-
2299: {\lambda_1}\, {2(N+8)\,{\left( 1 -\lambda_0 \right)
2300: }^{d/2}}}{1-\s012\lambda_1 (N+2)(1-\lambda_0)^{d/2}}\,. \eea
2301: %
2302: Hence, \eq{flowmass0} and \eq{flowlambda0} have a putative pole at
2303: $\lambda_0=-1$. In turn, \eq{flowmass1} and \eq{flowlambda1} have
2304: a putative pole at $(N+2)\lambda_1={2}(1-\lambda_0)^{-d/2}$. The
2305: putative pole is absent for $N=-2$.
2306:
2307:
2308: %********|*********|*********|*********|*********|*********|*********|****
2309: \subsection{Results}
2310: %********|*********|*********|*********|*********|*********|*********|****
2311:
2312: From the explicit form of the function \eq{ell-mod}, and without
2313: having made yet the explicit choice \eq{MassScales} for the mass term
2314: $m^2_k$, we conclude that the non-universal fixed point solution
2315: $\partial_t u=0$ of \eq{FlowPotential} for either $R_{\rm opt}$ from
2316: \eq{Ropt}, or the class of regulators $R_c$ from \eq{RMod} are related
2317: by a simple rescaling of the fields. The reason is the following: on
2318: the fixed point $\partial_t u=0$, we also have $\partial_t
2319: (m_k^2/k^2)=0$. Hence, the functions \Eq{ell-mod} do no longer depend
2320: on the flow of the mass term. Hence, \Eq{ell-mod} becomes
2321: $\ell(\omega)=\s02d C^{d/2+1}/(C+\omega)$, with $C\equiv 1-c\, \bar
2322: \omega={\rm const}$ on a fixed point. This function is equivalent to a
2323: flow derived from $R_{\rm opt}$. Hence, it suffices to rescale $u\to
2324: u/C^{d/2}$ and $\rho\to\rho/C^{d/2+1}$ in order to transform the fixed
2325: point solution for arbitrary $c$ onto the fixed point equation for
2326: $c=0$.\step
2327:
2328: Next, we check the $c$ dependence of critical exponents. A priori, the
2329: critical exponents are sensitive to the flow in the vicinity of the
2330: fixed point, and hence to the additional terms $\partial_t \lambda_0$
2331: in the flow equation. We have solved the flow \eq{FlowPotentialMod}
2332: with \eq{flowmass} numerically, within the polynomial approximation
2333: defined in \Eq{PolyAnsatz0} up to $n_{\rm trunc} =20$. The eigenvalues
2334: at criticality are found to be {\it independent} on the parameter $c$.
2335: The results correspond to the lower full line in Fig.~12 (and $x\equiv
2336: c$). Furthermore, we found that the critical exponents for different
2337: $c$ agree for every single order in the truncation. Hence, for $N=1$,
2338: the exponent $\nu_{\rm trunc}$ as obtained from \eq{FlowPotentialMod}
2339: with \eq{flowmass} are given by the line I in Fig.~10. \step
2340:
2341: In summary, the introduction of scale-dependent parameters into the
2342: regulator has lead to a significant change of the flow equations and
2343: their pole structure. Hence, the approach to a fixed point solution,
2344: and stability properties of the flow are quite different. Here, we
2345: studied the replacement $R_{\rm opt}(q^2)\to R_{\rm
2346: opt}(q^2+c\,m^2_k)$. Universal quantities are independent of $c$,
2347: because the modified regulator is linked to the optimised one through
2348: the replacement $k^2\to k^2_{\rm eff}(k)=k^2+c\,m^2_k$ in the flow
2349: equation. This should be irrelevant for universal observables at a
2350: fixed point, as has been confirmed explicitly. For the same reason,
2351: the entire class of regulators $R_a(q^2+c\,m^2_k)\equiv a R_{\rm
2352: opt}(q^2+c\,m^2_k)$ leads to $c$-independent critical exponents.
2353: Still, the flows are completely different for different $a$.
2354: Therefore, $c$ can be used as a free parameter to stabilise the flow,
2355: without affecting the physical result. If the substitution $R(q^2)\to
2356: R(q^2+c\,m^2_k)$ cannot be rephrased as a redefinition of the infrared
2357: scale, it is expected that also universal observables no longer remain
2358: insensitive to free parameters like $c$. This case should be studied
2359: separately.
2360:
2361:
2362: %********|*********|*********|*********|*********|*********|*********|****
2363: \begin{thebibliography}{99}
2364: %********|*********|*********|*********|*********|*********|*********|****
2365:
2366:
2367: %\cite{Zinn-Justin:1989mi}
2368: \bibitem{Zinn-Justin:1989mi}
2369: J.~Zinn-Justin,
2370: {\it Quantum Field Theory And Critical Phenomena},
2371: {Oxford, Clarendon (1989)}.
2372:
2373:
2374: \bibitem{Polchinski}
2375: J.\,Polchinski, Nucl.~Phys.~{\bf B231} (1984) 269.
2376:
2377: \bibitem{continuum}
2378: C.\,Wetterich, Phys.~Lett.~{\bf B301} (1993) 90.
2379:
2380: %\cite{Ellwanger:1994mw}
2381: \bibitem{Ellwanger:1994mw}
2382: U.~Ellwanger,
2383: %``Flow equations for N point functions and bound states,''
2384: Z.\ Phys.\ C {\bf 62} (1994) 503
2385: [hep-ph/9308260].
2386: %;\\
2387: %%CITATION = HEP-PH 9308260;%%
2388:
2389: %\cite{Morris:1994qb}
2390: \bibitem{Morris:1994qb}
2391: T.~R.~Morris,
2392: %``The Exact renormalization group and approximate solutions,''
2393: Int.\ J.\ Mod.\ Phys.\ {\bf A9} (1994) 2411
2394: [hep-ph/9308265].
2395: %%CITATION = HEP-PH 9308265;%%
2396:
2397: %\cite{Bagnuls:2000ae}
2398: \bibitem{Bagnuls:2000ae}
2399: For recent reviews, see:\\
2400: C.~Bagnuls and C.~Bervillier,
2401: %``Exact renormalization group equations: An introductory review,''
2402: Phys.\ Rept.\ {\bf 348} (2001) 91
2403: [hep-th/0002034],\\
2404: J.~Berges, N.~Tetradis and C.~Wetterich,
2405: %``Non-perturbative renormalisation flow in quantum field theory
2406: %and statistical physics,''
2407: hep-ph/0005122.
2408: %%CITATION = HEP-PH 0005122;%%
2409:
2410:
2411: %\cite{Litim:2000ci}
2412: \bibitem{Litim:2000ci}
2413: D.~F.~Litim,
2414: %``Optimisation of the exact renormalisation group,''
2415: Phys.\ Lett.\ {\bf B486} (2000) 92
2416: [hep-th/0005245].
2417: %%CITATION = HEP-TH 0005245;%%
2418:
2419:
2420: %\cite{Litim:2001up}
2421: \bibitem{Litim:2001up}
2422: D.~F.~Litim,
2423: %``Optimised renormalisation group flows,''
2424: Phys.\ Rev.\ D {\bf 64} (2001) 105007
2425: [hep-th/0103195].
2426: %%CITATION = HEP-TH 0103195;%%
2427:
2428: %\cite{Litim:2001fd}
2429: \bibitem{Litim:2001fd}
2430: D.~F.~Litim,
2431: %``Mind the gap,''
2432: Int.\ J.\ Mod.\ Phys.\ A {\bf 16} (2001) 2081
2433: [hep-th/0104221].
2434: %%CITATION = HEP-TH 0104221;%%
2435:
2436: %\cite{Litim:2001dt}
2437: \bibitem{Litim:2001dt}
2438: D.~F.~Litim,
2439: %``Derivative expansion and renormalisation group flows,''
2440: JHEP {\bf 0111} (2001) 059
2441: [hep-th/0111159].
2442: %%CITATION = HEP-TH 0111159;%%
2443:
2444: %\cite{Litim:1998nf}
2445: \bibitem{Litim:1998nf}
2446: For a review, see: D.~F.~Litim and J.~M.~Pawlowski,
2447: %``On gauge invariant Wilsonian flows,''
2448: hep-th/9901063.
2449: %%CITATION = HEP-TH 9901063;%%
2450:
2451:
2452: %\cite{QuantumGravity1}
2453: \bibitem{QuantumGravity1}
2454: M.~Reuter,
2455: %``Nonperturbative Evolution Equation for Quantum Gravity,''
2456: Phys.\ Rev.\ D {\bf 57} (1998) 971
2457: hep-th/9605030].
2458: %%CITATION = HEP-TH 9605030;%%
2459:
2460: %\cite{Hasenfratz:1986dm}
2461: \bibitem{Hasenfratz:1986dm}
2462: A.~Hasenfratz and P.~Hasenfratz,
2463: %``Renormalization Group Study Of Scalar Field Theories,''
2464: Nucl.\ Phys.\ B {\bf 270} (1986) 687.
2465: %[Helv.\ Phys.\ Acta {\bf 59} (1986) 833].
2466: %%CITATION = NUPHA,B270,687;%%
2467:
2468:
2469: %\cite{Margaritis:1988hv}
2470: \bibitem{Margaritis:1988hv}
2471: A.~Margaritis, G.~Odor and A.~Patkos,
2472: %``Series Expansion Solution Of The Wegner-Houghton Renormalization Group Equation,''
2473: Z.\ Phys.\ C {\bf 39} (1988) 109.
2474: %%CITATION = ZEPYA,C39,109;%%
2475:
2476: %\cite{Tetradis:1994ts}
2477: \bibitem{Tetradis:1994ts}
2478: N.~Tetradis and C.~Wetterich,
2479: %``Critical exponents from effective average action,''
2480: Nucl.\ Phys.\ B {\bf 422} (1994) 541
2481: [hep-ph/9308214].
2482: %%CITATION = HEP-PH 9308214;%%
2483:
2484: %\cite{Morris:1994ie}
2485: \bibitem{Morris:1994ie}
2486: T.~R.~Morris,
2487: %``Derivative expansion of the exact renormalization group,''
2488: Phys.\ Lett.\ {\bf B329} (1994) 241
2489: [hep-ph/9403340].
2490: %%CITATION = HEP-PH 9403340;%%
2491:
2492: %\cite{Ball:1995ji}
2493: \bibitem{Ball:1995ji}
2494: R.~D.~Ball, P.~E.~Haagensen, J.~I.~Latorre and E.~Moreno,
2495: %``Scheme independence and the exact renormalisation group,''
2496: Phys.\ Lett.\ {\bf B347} (1995) 80.
2497: %%CITATION = HEP-TH 9411122;%%
2498:
2499: %\cite{Litim:1995ex}
2500: \bibitem{Litim:1995ex}
2501: D.~F.~Litim and N.~Tetradis,
2502: %``Approximate solutions of exact renormalization group equations,''
2503: hep-th/9501042.
2504: %%CITATION = HEP-TH 9501042;%%
2505:
2506: %\cite{Tetradis:1996br}
2507: \bibitem{Tetradis:1996br}
2508: N.~Tetradis and D.~F.~Litim,
2509: %``Analytical Solutions of Exact Renormalization Group Equations,''
2510: Nucl.\ Phys.\ {\bf B464} (1996) 492
2511: [hep-th/9512073].
2512: %%CITATION = HEP-TH 9512073;%%
2513:
2514: %\cite{Comellas:1997tf}
2515: \bibitem{Comellas:1997tf}
2516: J.~Comellas and A.~Travesset,
2517: %``O(N) models within the local potential approximation,''
2518: Nucl.\ Phys.\ B {\bf 498} (1997) 539
2519: [hep-th/9701028].
2520: %%CITATION = HEP-TH 9701028;%%
2521:
2522: %\cite{Morris:1998xj}
2523: \bibitem{Morris:1998xj}
2524: T.~R.~Morris and M.~D.~Turner,
2525: %``Derivative expansion of the renormalization group in O(N) scalar field theory,''
2526: Nucl.\ Phys.\ B {\bf 509} (1998) 637
2527: [hep-th/9704202].
2528: %%CITATION = HEP-TH 9704202;%%
2529:
2530:
2531: %\cite{Liao:2000sh}
2532: %\bibitem{LPS}
2533: \bibitem{Liao:2000sh}
2534: S.~B.~Liao, J.~Polonyi and M.~Strickland,
2535: %``Optimisation of renormalisation group flow,''
2536: Nucl.\ Phys.\ {\bf B567} (2000) 493 [hep-th/9905206].
2537: %%CITATION = HEP-TH 9905206;%%
2538: %\href{\wwwspires?eprint=HEP-TH/9905206}{SPIRES}
2539:
2540: %\cite{VonGersdorff:2000kp}
2541: \bibitem{VonGersdorff:2000kp}
2542: G.~Von Gersdorff and C.~Wetterich,
2543: %``Nonperturbative renormalization flow
2544: %and essential scaling for the Kosterlitz-Thouless transition,''
2545: Phys.\ Rev.\ B {\bf 64} (2001) 054513 [hep-th/0008114].
2546: %%CITATION = HEP-TH 0008114;%%
2547:
2548: %\cite{Litim:2001hk}
2549: \bibitem{Litim:2001hk}
2550: D.~F.~Litim and J.~M.~Pawlowski,
2551: %``Predictive power of renormalisation group flows: A comparison,''
2552: Phys.\ Lett.\ B {\bf 516} (2001) 197
2553: [hep-th/0107020].
2554: %D.~F.~Litim and J.~M.~Pawlowski,
2555: %in preparation.
2556: %%CITATION = HEP-TH 0107020;%%
2557:
2558: %\cite{Bohr:2001gp}
2559: \bibitem{Bohr:2001gp}
2560: O.~Bohr, B.J.~Sch\"afer and J.~Wambach,
2561: %``Renormalization group flow equations and the phase transition in O(N)
2562: %models,''
2563: Int.\ J.\ Mod.\ Phys.~{\bf A16} (2001) 3823 [hep-ph/0007098],\\
2564: %[hep-ph/0007098];
2565: %%CITATION = HEP-PH 0007098;%%
2566: %\cite{Bonanno:2000yp}
2567: %\bibitem{Bonanno:2000yp}
2568: A.~Bonanno and D.~Zappal\`a,
2569: %``Towards an accurate determination of the critical exponents
2570: %with the renormalization group flow equations,''
2571: Phys.\ Lett.\ B {\bf 504} (2001) 181 [hep-th/0010095],\\
2572: %%CITATION = HEP-TH 0010095;%%
2573: %\cite{Mazza:2001bp}
2574: %\bibitem{Mazza:2001bp}
2575: M.~Mazza and D.~Zappal\`a,
2576: %``Proper time regulator and renormalization group flow,''
2577: Phys.\ Rev.\ D {\bf 64} (2001) 105013 [hep-th/0106230].
2578: %%
2579:
2580: %\cite{Liao:1996fp}
2581: \bibitem{Liao:1996fp}
2582: S.~B.~Liao,
2583: %``On connection between momentum cutoff
2584: %and the proper time regularisations,''
2585: Phys.\ Rev.\ {\bf D53} (1996) 2020
2586: [hep-th/9501124].
2587: %%CITATION = HEP-TH 9501124;%%
2588:
2589: %\cite{Litim:2001ky}
2590: \bibitem{Litim:2001ky}
2591: D.~F.~Litim and J.~M.~Pawlowski,
2592: %``Perturbation theory and renormalisation group equations,''
2593: %``Completeness and consistency of renormalisation group flows,''
2594: hep-th/0111191; hep-th/0202188.
2595: %%CITATION = HEP-TH 0111191;%%
2596: %%CITATION = HEP-TH 0202188;%%
2597:
2598: %\cite{Litim:1997nw}
2599: \bibitem{Litim:1997nw}
2600: D.~F.~Litim,
2601: %``Scheme independence at first order phase
2602: %transitions and the renormalisation group,''
2603: Phys.\ Lett.\ {\bf B393} (1997) 103 [hep-th/9609040].
2604: %%CITATION = HEP-TH 9609040;%%
2605:
2606: %\cite{Freire:2001sx}
2607: \bibitem{Freire:2001sx}
2608: F.~Freire and D.~F.~Litim,
2609: Phys.\ Rev.\ D {\bf 64} (2001) 045014 [hep-ph/0002153].
2610: %%CITATION = HEP-PH 0002153;%%
2611:
2612: %\cite{Latorre:2000qc}
2613: %\bibitem{Latorre:2000qc}
2614: %J.~I.~Latorre and T.~R.~Morris,
2615: %JHEP {\bf 0011} (2000) 004 [hep-th/0008123].
2616: %%CITATION = HEP-TH 0008123;%%
2617:
2618:
2619: \bibitem{ScalarQED}
2620: %\cite{Litim:1997jd}
2621: %\bibitem{Litim:1997jd}
2622: D.~F.~Litim, C.~Wetterich and N.~Tetradis,
2623: %``Nonperturbative analysis of the Coleman-Weinberg phase transition,''
2624: Mod.\ Phys.\ Lett.\ A {\bf 12} (1997) 2287,\\
2625: %[hep-ph/9407267],\\
2626: %%CITATION = HEP-PH 9407267;%%
2627: %\cite{ScalarQED-N}
2628: %\bibitem{ScalarQED-N}
2629: B.~Bergerhoff, D.~F.~Litim, S.~Lola and C.~Wetterich,
2630: Int.\ J.\ Mod.\ Phys.\ {\bf A11} (1996) 4273,\\
2631: %\bibitem{ScalarQED}
2632: %%CITATION = COND-MAT 9502039;%%
2633: B.~Bergerhoff, F.~Freire, D.~F.~Litim, S.~Lola and C.~Wetterich,
2634: Phys.\ Rev.\ {\bf B53} (1996) 5734.
2635: %%CITATION = HEP-PH 9503334;%%
2636:
2637: %\cite{QuantumGravity2}
2638: \bibitem{QuantumGravity2}
2639: O.~Lauscher and M.~Reuter, hep-th/0108040, hep-th/0110021;\\
2640: %%CITATION = HEP-TH 0108040;%%
2641: %%CITATION = HEP-TH 0110021;%%
2642: M.~Reuter and F.~Saueressig, hep-th/0110054.
2643: %%CITATION = HEP-TH 0110054;%%
2644:
2645: %\cite{Morris:1994ki}
2646: \bibitem{Morris:1994ki}
2647: T.~R.~Morris,
2648: %``On truncations of the exact renormalization group,''
2649: Phys.\ Lett.\ B {\bf 334} (1994) 355
2650: [hep-th/9405190].
2651: %%CITATION = HEP-TH 9405190;%%
2652:
2653:
2654: %\cite{Aoki:1998um}
2655: \bibitem{Aoki:1998um}
2656: K.~Aoki, K.~Morikawa, W.~Souma, J.~Sumi and H.~Terao,
2657: %``Rapidly converging truncation scheme of the exact renormalization group,''
2658: Prog.\ Theor.\ Phys.\ {\bf 99} (1998) 451.
2659: %[hep-th/9803056].
2660: %%CITATION = HEP-TH 9803056;%%
2661:
2662: %\cite{Golner:1986}
2663: \bibitem{Golner:1986}
2664: G.~R.~Golner, Phys.~Rev.~{\bf B33} (1986) 7863.
2665:
2666: %\cite{Wetterich:1993be}
2667: \bibitem{Wetterich:1993be}
2668: C.~Wetterich,
2669: %``The Average action for scalar fields near phase transitions,''
2670: Z.\ Phys.\ C {\bf 57} (1993) 451.
2671: %%CITATION = ZEPYA,C57,451;%%
2672:
2673: %\cite{Kubyshin:2001gz}
2674: %\bibitem{Kubyshin:2001gz}
2675: %Y.~Kubyshin, R.~Neves and R.~Potting,
2676: %``Polchinski ERG equation in O(N) scalar field theory,''
2677: %Int.\ J.\ Mod.\ Phys.\ A {\bf 16} (2001) 2065
2678: %[hep-th/0101191].
2679: %%CITATION = HEP-TH 0101191;%%
2680:
2681:
2682: \bibitem{QCD1}
2683: %\cite{Litim:1998qi}
2684: %\bibitem{Litim:1998qi}
2685: D.~F.~Litim and J.~M.~Pawlowski,
2686: %``Flow equations for Yang-Mills theories in general axial gauges,''
2687: Phys.\ Lett.\ B {\bf 435} (1998) 181
2688: [hep-th/9802064]; hep-th/0203005.
2689: %%CITATION = HEP-TH 9802064;%%
2690: %%CITATION = HEP-TH 0203005;%%
2691:
2692:
2693: %\cite{Ellwanger:1996qf}
2694: \bibitem{Ellwanger:1996qf}
2695: U.~Ellwanger, M.~Hirsch and A.~Weber,
2696: %``Flow equations for the relevant part of the pure Yang-Mills action,''
2697: Z.\ Phys.\ C {\bf 69} (1996) 687
2698: [hep-th/9506019],\\
2699: %%CITATION = HEP-TH 9506019;%%
2700: %\cite{Bergerhoff:1998cv}
2701: %\bibitem{Bergerhoff:1998cv}
2702: B.~Bergerhoff and C.~Wetterich,
2703: %``Effective quark interactions and QCD propagators,''
2704: Phys.\ Rev.\ D {\bf 57} (1998) 1591
2705: [hep-ph/9708425].
2706: %%CITATION = HEP-PH 9708425;%%
2707:
2708:
2709:
2710: \end{thebibliography}
2711: \end{document}
2712:
2713:
2714: