1: \documentclass[12pt,a4paper]{article}
2: \usepackage{amsmath,epsfig}
3: \setlength{\textwidth}{15cm}
4: \begin{document}
5:
6: \begin{titlepage}
7: %
8: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
9: %
10: \begin{flushright}
11: %{\bf DRAFT 4}\\
12: MPI-PhT/2000-44\\
13: %hep-th/9612039
14: \end{flushright}
15:
16: \vspace{1cm}
17:
18: \begin{center}
19: {\LARGE Kosterlitz-Thouless theory and lattice artifacts}
20: \vspace{1.6cm}
21:
22: {{\large J. Balog\footnote{on leave of absence from the Research
23: Institute for Particle and Nuclear Physics, Budapest, Hungary}\\[8mm]
24: {\small\sl Max-Planck-Institut f\"{u}r Physik}\\
25: {(\small\sl Werner Heisenberg Institut)}\\[3mm]
26: {\small\sl F\"{o}hringer Ring 6, 80805 Munich, Germany}}}
27:
28: \vspace{2.4cm}
29: {\bf Abstract}
30: \end{center}
31:
32: \begin{quote}
33: The massive continuum limit of the $1+1$ dimensional $O(2)$
34: nonlinear $\sigma$-model (XY model) is studied using its
35: equivalence to the Sine-Gordon model at its asymptotically free
36: point. It is shown that leading lattice artifacts are universal but
37: they vanish only as inverse powers of the logarithm of the correlation
38: length. Such leading artifacts are calculated for the case of the scattering
39: phase shifts and the correlation function of the Noether current using
40: the bootstrap S-matrix and perturbation theory respectively.
41: \end{quote}
42: \vfill
43: \setcounter{footnote}{0}
44: %
45: \end{titlepage}
46:
47:
48: %
49: %
50: \newcommand{\be}{\begin{equation}}
51: %\newcommand{\ee}{\end{equation}}
52: \newcommand{\ba}{\begin{eqnarray}}
53: \newcommand{\ea}{\end{eqnarray}}
54: %
55: %
56:
57:
58: \section{Introduction}
59: In this paper we study the properties of the two-dimensional
60: $O(2)$ nonlinear $\sigma$-model, better known as the XY model.
61: This model has been the subject of extensive theoretical and numerical
62: analysis, starting with the seminal papers of Kosterlitz and Thouless (KT)
63: \cite{KT}. For a review of KT theory, see \cite{ZinnJ}.
64:
65: Analitical work is usually based on a series of mappings that starts at
66: the original (lattice) XY model and arrives at the Sine-Gordon (SG) model
67: or its fermionic equivalent, a (deformed) version of the chiral Gross-Neveu
68: (CGN) model. This latter formulation is most useful if one wants to study
69: questions related to the dynamically generated $SU(2)$ symmetry of the model.
70:
71: Most papers on the XY model study its properties interesting for
72: Statistical Physics, in particular the pecularities of the KT phase
73: transition, which is of infinite order. In this paper we look at the XY
74: model as an example of $1+1$ dimensional relativistic Quantum Field Theory.
75: More precisely, we study the massive continuum limit of the lattice theory,
76: which, in the language of Statistical Physics, means that we approach the
77: KT phase transition point from the high temperature phase.
78:
79: Treating the XY model as the $n=2$ member of the family of $O(n)$
80: nonlinear $\sigma$-models gives additional insights, since a lot is
81: known about the $n\geq3$ models~\cite{ZZ}. More importantly, using the
82: SG language, we show that the approach to the continuum limit in this
83: model is much slower than in most other lattice models. Lattice artifacts
84: vanish in this model, instead of the
85: usual Symanzik type behaviour~\cite{Sym} (i.e. integer powers of the
86: lattice spacing), as inverse powers of the logarithm of the lattice
87: spacing only. On the other hand, we can show that the leading artifacts
88: are universal and calculable. Our main result is Eqs. (\ref{gammaexp})
89: and (\ref{xiexp}) in Section 4, which allow us to calculate leading
90: lattice artifacts in terms of SG data.
91:
92: We can make use of the fact that the SG model is exactly solvable
93: and its bootstrap S-matrix is exactly known. We calculate the leading
94: artifacts for the scattering phase shifts using the bootstrap results.
95: An alternative method is perturbation theory (PT). Since the SG model
96: is asymptotically free if we use suitable expansion parameters
97: \cite{Amit}, the methods of renormalization group (RG) improved PT
98: are thus available.
99:
100: In Section 2 we review the relation of the XY model to the $O(n)$ models
101: with $n\geq3$ and describe the chain of mappings leading from the XY model
102: to the SG model and its equivalent fermionic formulation.
103:
104: In Section 3 we recall the analysis of the phase diagram of the model
105: in the vicinity of the KT phase transition point. This is described in the
106: SG language.
107:
108: In Section 4 we explain how to calculate the lattice artifacts and apply
109: this to the case of the scattering phase shifts.
110:
111: Finally in Section 5 we calculate the lattice artifacts for the
112: two-point correlation function of the Noether current corresponding
113: to the $O(2)$ symmetry. Here we use the method of RG improved PT.
114: To calculate the value of a non-perturbative constant needed here,
115: we also consider the system in the presence of an external field coupled
116: to the Noether charge. (This calculation is analogous to the one used
117: previously to determine the $M/\Lambda$ ratio for the $O(n)$ models
118: \cite{HMN}.) We give here a parameter-free two-loop formula for the
119: lattice artifacts.
120:
121: A precision MC study of the massive continuum limit of the $O(2)$ model
122: will be described in a forthcoming paper \cite{paper3}.
123:
124:
125:
126: \section{From the XY model to the Sine-Gordon model and beyond}
127:
128: In this section we describe in some detail the chain of mappings starting
129: with the XY model and ending at the SG model and its fermionic equivalent.
130:
131: We can treat the XY model as the $n=2$ member of the family of
132: $O(n)$ nonlinear $\sigma$-models with Lagrangean
133: \be
134: {\cal L}^{O(n)}=\frac{1}{2g^2}\,\partial_\mu S^a\partial_\mu S^a
135: \quad ;\quad S^aS^a=1,\qquad a=1,2,\dots n.
136: \label{LagOn}
137: \end{equation}
138: The $n\geq3$ models are known to be integrable. Polyakov \cite{Poly}
139: and L\"uscher \cite{Lusch} have shown the existence of
140: respectively local and nonlocal higher spin conserved charges, whose
141: existence implies quantum integrability. Assuming the spectrum of the
142: model consisted of an $O(n)$ vector multiplet of massive particles the
143: exact S-matrix of the $n\geq3$ models was found by bootstrap methods
144: \cite{ZZ}:
145: \be
146: S_{ab;cd}(\theta)=
147: \sigma_1(\theta)\delta_{ab}\delta_{cd}+
148: \sigma_2(\theta)\delta_{ac}\delta_{bd}+
149: \sigma_3(\theta)\delta_{ad}\delta_{bc},
150: \label{OnSmatrix}
151: \end{equation}
152: where
153: \begin{alignat}{1}
154: \sigma_1(\theta)&=\frac{-2\pi i\theta}{i\pi-\theta}
155: \,\cdot\,\frac{s^{(2)}(\theta)}
156: {(n-2)\theta-2\pi i}\,,\nonumber\\
157: \sigma_2(\theta)&=(n-2)\theta
158: \,\cdot\,\frac{s^{(2)}(\theta)}
159: {(n-2)\theta-2\pi i}\,,\label{sig2}\\
160: \sigma_3(\theta)&=-2\pi i
161: \,\cdot\,\frac{s^{(2)}(\theta)}
162: {(n-2)\theta-2\pi i}\nonumber
163: \end{alignat}
164: and the \lq isospin 2' phase shift $s^{(2)}$ is given by
165: \be
166: s^{(2)}(\theta)=-\exp\left\{2i\int_0^\infty\frac{d\omega}{\omega}
167: \sin\omega\theta \tilde K_n(\omega)\right\}
168: \label{iso2}
169: \end{equation}
170: with
171: \be
172: \tilde K_n(\omega)=
173: \left[\frac{e^{-\pi\omega}+e^{-2\pi\frac{\omega}{n-2}}}
174: {1+e^{-\pi\omega}}\right].
175: \label{Kn}
176: \end{equation}
177:
178: Much less is known about the $O(2)$ model. A simple observation
179: is that (\ref{sig2}) and also (\ref{Kn}) have a smooth
180: $n\to2$ limit. It is natural to assume that the $O(2)$ model is
181: also integrable, its spectrum consists of a single $O(2)$ doublet
182: of massive particles whose scattering is indeed described by the
183: $n\to2$ limit of the S-matrix (\ref{OnSmatrix}).
184:
185: Although taking the formal $n\to2$ limit of the bootstrap results
186: valid for $n\geq3$ is not convincing in itself, the conclusion
187: turns out to be correct because as we will see it also follows from
188: the Kosterlitz-Thouless theory \cite{KT} of the XY model.
189:
190: Before turning to the KT theory we make a small digression to
191: discuss the two-dimensional Sine-Gordon (SG) model.
192: Its Lagrangean can be written as
193: \begin{equation}
194: {\cal L}^{\rm SG}=\frac12\partial_\mu\phi\partial_\mu\phi
195: +\frac{\alpha}{\beta^2a^2}\left[1-\cos(\beta\phi)\right],
196: \label{LagSG}
197: \end{equation}
198: where $\alpha$ is the dimensionless mass parameter, $a$ is
199: a constant of dimension mass$^{-1}$ and $\beta$ is
200: the SG coupling. It is also integrable and its spectrum and S-matrix
201: was also found in \cite{ZZ}. The spectrum depends on $\beta$
202: in a complicated way but it becomes simple for the range
203: $8\pi>\beta^2>4\pi$ when it is free of any bound states
204: and consists of a single $O(2)$ vector of massive particles whose S-matrix
205: can again be written as (\ref{OnSmatrix}) but now
206: \begin{alignat}{1}
207: \sigma_1(\theta)&=\cosh\frac{i\pi\nu}{2}
208: \sinh\frac{\theta\nu}{2}\,\,\frac{s^{(2)}_{\rm SG}(\theta)}
209: {\sinh\frac{(i\pi-\theta)\nu}{2}}\,,\nonumber\\
210: \sigma_2(\theta)&=-\sinh\frac{i\pi\nu}{2}
211: \sinh\frac{\theta\nu}{2}\,\,\frac{s^{(2)}_{\rm SG}(\theta)}
212: {\cosh\frac{(i\pi-\theta)\nu}{2}}\,,\label{sig2SG}\\
213: \sigma_3(\theta)&=\cosh\frac{i\pi\nu}{2}
214: \cosh\frac{\theta\nu}{2}\,\,\frac{s^{(2)}_{\rm SG}(\theta)}
215: {\cosh\frac{(i\pi-\theta)\nu}{2}}\,,\nonumber
216: \end{alignat}
217: where we have introduced the parametrization
218: \be
219: \nu=\frac{8\pi}{\beta^2}-1.
220: \label{gammapar}
221: \end{equation}
222: The \lq isospin 2' phase shift for the SG model is
223: \be
224: s^{(2)}_{\rm SG}(\theta)=-\exp\left\{2i\int_0^\infty\frac{d\omega}{\omega}
225: \sin\omega\theta\,\tilde k(\omega)\right\}
226: \label{iso2SG}
227: \end{equation}
228: with
229: \be
230: \tilde k(\omega)=\frac{\sinh\frac{\pi\omega(1-\nu)}{2\nu}}
231: {2\cosh\frac{\pi\omega}{2}\sinh\frac{\pi\omega}{2\nu}}.
232: \label{tildek}
233: \end{equation}
234: Note that in the $\beta^2\to8\pi$ ($\nu\to0$) limit the SG S-matrix
235: coincides with the $n\to2$ limit of the $O(n)$ S-matrix, in particular
236: $\lim_{\nu\to0}\tilde k(\omega)=\lim_{n\to2}\tilde K_n(\omega)$.
237:
238: The identification of the XY model with the $\nu\to0$ limit of the
239: SG model is surprising since in this limit the bootstrap
240: S-matrix (\ref{sig2SG}) becomes SU(2) symmetric, coinciding with
241: the S-matrix of the $SU(2)$ chiral Gross-Neveu (CGN) model \cite{pew}.
242: It is not obvious where this enlarged symmetry comes from.
243: The existence of a nontrivial XY model is even more surprising in the
244: light of the fact that the beta-function of the coupling $g^2$
245: in (\ref{LagOn}) vanishes for $n=2$ and by making the substitution
246: $S^1=\cos\varphi\,,\,S^2=\sin\varphi$ the Lagrangean (\ref{LagOn})
247: naively becomes free.
248:
249: Kosterlitz and Thouless \cite{KT} argued that the fact that $\varphi$
250: is a periodic (angular) variable plays an important role and therefore
251: the model has nontrivial dynamics. They have shown that topologically
252: nontrivial objects, vortices, are present in typical spin configurations
253: and their interaction makes the theory nontrivial.
254:
255: The standard lattice action of the XY model is
256: \be
257: S_{\rm XY}=K\sum_{x,\mu}\big[1-
258: \cos\big(\varphi(x)-\varphi(x+\hat\mu)\big)\big].
259: \label{standardS}
260: \end{equation}
261: We denoted by $K$ the inverse of the XY model coupling to avoid confusion
262: with the SG coupling $\beta$. Assuming universality, not only the cosine
263: function but any other $2\pi$-periodic function $W(\varphi)$ which has a local
264: minimum at $\varphi=0$ defines a possible XY model lattice action.
265: The Villain model action \cite{Villain} is chartacterized by
266: \be
267: W(\varphi)=-\frac{1}{K_V}\ln\Big[\sum_m\exp\big\{-\frac{K_V}{2}
268: (\varphi-2\pi m)^2\big\}\Big].
269: \label{VillainS}
270: \end{equation}
271:
272: Kosterlitz and Thouless showed that typical spin configurations
273: can be represented as a mixture of smooth, topologically trivial
274: configurations (spin waves) and a gas of vortices (of integer topological
275: charge). The KT vortices are not interacting with the spin waves, but
276: there is a logarithmic interaction potential between the vortices which
277: are therefore identical to a two-dimensional Coulomb gas.
278: This spin wave $+$ Coulomb gas (SWCG) picture is only approximate
279: if we start from the standard action (\ref{standardS}) but it is an exact
280: duality transformation \cite{Jose} for the Villain action corresponding to
281: (\ref{VillainS}).
282: That the XY model with standard action is in the same universality class
283: as the Villain model was demonstrated using Monte Carlo renormalization
284: group techniques \cite{Hasenbusch}. On the other hand, it has been
285: shown rigorously \cite{FrohSpen} that the Coulomb gas has a phase transition
286: point at some finite critical coupling $K_c$. KT interpreted this phase
287: transition as one of vortex condensation and by a (heuristic) energy-entropy
288: consideration showed that in the vicinity of $K_c$ vortices of topological
289: charge $\pm1$ only are important, higher vortices can be neglected.
290: It is easy to see that this system (SWCG with unit charge vortices only)
291: is exactly equivalent to the SG model. In ref. \cite{Amit} it was shown that
292: the extremal SG fixed point $\beta^*=\sqrt{8\pi}\,,\alpha^*=0$ is
293: appropriate to discribe the KT phase transition. The renormalizabilty
294: of the SG model around this point was explicitly demonstrated up to
295: two-loop order in a simultaneous perturbative expansion in $\alpha$
296: and $\delta=\frac{\beta^2-8\pi}{8\pi}$.
297:
298: Finally, there is a further transformation that explains the dynamical
299: $SU(2)$ symmetry of the XY model. The SG model can be exactly mapped
300: \cite{Banksetal} to a fermionic model formulated in terms of a two-component
301: Dirac fermion $\psi$. The transformation is similar to the well-known one
302: that relates the SG model to the massive Thirring model \cite{Coleman}.
303: Here the fermionic model is a deformation of the chiral Gross-Neveu model
304: with four-fermion interaction:
305: \be
306: {\cal L}_F=i(\overline{\psi}\gamma_\mu\partial_\mu\psi)-
307: g_0(J_\mu^1)^2-g_0(J_\mu^2)^2-(g_0+f_0)(J_\mu^3)^2\,,
308: \label{LagF}
309: \end{equation}
310: where
311: \be
312: J_\mu^a=\frac{1}{2}\,\overline{\psi}\gamma_\mu\sigma^a\psi
313: \label{SU2curr}
314: \end{equation}
315: is the fermionic $SU(2)$ current. The relation between the SG couplings
316: $\delta\,,\alpha$ and the fermion couplings $g_0\,,f_0$ is
317: \be
318: \alpha=\frac{8}{\pi}g_0+\cdots,\qquad\qquad
319: \delta=-\frac{1}{\pi}(g_0+f_0)+\cdots,
320: \label{g0f0}
321: \end{equation}
322: where the dots indicate that the relations (\ref{g0f0}) receive higher
323: order corrections in perturbation theory. In the fermionic formulation
324: the KT fixed point is the Gaussian one and for vanishing deformation
325: parameter, $f_0=0$, the model is manifestly $SU(2)$ symmetric. The
326: corresponding relation in the SG language is $\alpha+8\delta=0$ at
327: lowest order.
328:
329: To summarize, the XY model in the vicinity of the Kosterlitz-Thouless
330: transition point is believed to be described by the SG model with
331: extremal coupling $\beta=\sqrt{8\pi}$. This is further equivalent to
332: the two-component chiral Gross-Neveu model around its Gaussian point.
333: We will use the SG language throughout this paper.
334:
335:
336:
337:
338:
339:
340:
341:
342: \section{The SG description of the $O(2)$ model}
343:
344: In this section we review the SG description of the KT theory
345: closely following the approach of Amit et al. \cite{Amit}.
346: Without loss of generality we can adopt the somewhat
347: unusual regularization scheme of the authors,
348: since, as we will see, all important results are universal,
349: i.e. independent of the regularization scheme. Nevertheless, it would
350: be interesting to repeat all the calculations below using some of the
351: more customary regularizations like the lattice or dimensional regularization.
352:
353: Our starting point is the Euklidean Lagrangian \cite{Amit}
354: \begin{equation}
355: {\cal L}=\frac12\partial_\mu\phi\partial_\mu\phi
356: +\frac{m_0^2}{2}\phi^2+\frac{\alpha_0}{\beta_0^2a^2}\left[
357: 1-\cos(\beta_0\phi)\right],
358: \label{Lag}
359: \end{equation}
360: where $m_0$ is an IR regulator mass and $a$ is the UV
361: cutoff (of dimension length). We have denoted the dimensionless
362: SG couplings by $\beta_0$ and $\alpha_0$ to emphasize that they are
363: bare (unrenormalized). UV regularized correlation functions are
364: calculated by using
365: \begin{equation}
366: G_0(x)=\frac{1}{2\pi}\,K_0\left(m_0\sqrt{x^2+a^2}\right)
367: \label{G0}
368: \end{equation}
369: where $K_0$ is the modified Bessel function, as the
370: $\phi$ propagator. Our strategy is slightly different from
371: \cite{Amit}, who really considered the renormalization of
372: the massive SG model (\ref{Lag}) of mass $m_0$. We treat
373: $m_0$ as an IR regulator mass and consider IR stable
374: physical quantities for which we can take the limit
375: ${m_0\to0}$ already at the UV regularized level
376: (before UV renormalization). All renormalization constants are,
377: for example, IR stable and independent of $m_0$.
378:
379: The SG coupling $\beta_0$ is close to its special value $\sqrt{8\pi}$
380: and a simultaneous perturbative expansion in the mass
381: parameter $\alpha_0$ and the deviation $\delta_0$ is defined, where
382: \be
383: \beta_0^2=8\pi(1+\delta_0)
384: \end{equation}
385: and the parameters are renormalized according to
386: \ba
387: \alpha_0&=&Z_\alpha\alpha\,,\\
388: \beta_0^2&=&Z_\phi^{-1}\beta^2\,.
389: \ea
390: Here the Z-factors are functions of the renormalized couplings
391: $\alpha$ and $\delta$ and the combination
392: \be
393: l=\ln\mu a\,,
394: \end{equation}
395: where $\mu$ is an arbitrary mass parameter (basically the normalization
396: point). Similarly, a renormalization constant $Z$ is necessary to make
397: ${\cal G}$, the spin-spin 2-point function finite:
398: \be
399: {\cal G}_R=Z{\cal G}\,.
400: \end{equation}
401:
402:
403:
404:
405:
406:
407:
408: For vanishing mass parameter $\alpha$ the Lagrangian (\ref{Lag}) is trivial
409: and $Z_\phi=1$ since there is no need to renormalize the SG coupling. The
410: spin-spin correlation function (which is an exponential of
411: the basic field $\phi$) gets renormalized even in this point, but in this
412: case its renormalization constant is simply
413: \be
414: Z=(\mu a)^{-{1\over4(1+\delta)}}\,.
415: \end{equation}
416: In addition, there is a symmetry $\alpha\leftrightarrow-\alpha$
417: (which corresponds to a $\pi\over\beta_0$ shift in the basic field).
418:
419: Taking into account the above constraints, the perturbative expansion
420: of the Z-factors must be of the form
421: \ba
422: Z_\phi&=&1+f_1\alpha^2l+\alpha^2\delta(\bar f_2l^2+f_2l)+\cdots\,,\\
423: Z_\alpha&=&1+g_1\delta l+\alpha^2(\bar g_2l^2+g_2l)+\delta^2(\bar g_3l^2+
424: g_3l)+\cdots\,,\\
425: Z&=&e^{-{l\over4(1+\delta)}}\big\{1+\alpha^2(\bar h_1l^2+h_1l)+
426: \cdots\big\}\,.
427: \ea
428: Amit et al. \cite{Amit} found the following results.
429: \ba
430: f_1={1\over32}\,,\qquad\qquad\bar f_2&=&-{1\over16}\,,\qquad\qquad\ \
431: f_2=-{3\over32}\,,\nonumber\\
432: g_1=-2\,,\qquad\qquad\bar g_2&=&\phantom{-}{1\over32}\,,\qquad\qquad\ \
433: g_2=-{5\over64}\,,\label{Am2}\\
434: \bar g_3&=&\phantom{+}\ 2\,,\qquad\qquad\ \ \ g_3=
435: \phantom{-}\ 0\,.\nonumber
436: \ea
437: Furthermore $\bar h_1=-1/256$, but the number $h_1$ is not known at present.
438: The above two-loop beta-function coefficients were calculated also by
439: other methods. The original calculation has recently been reconsidered and
440: the results (\ref{Am2}) have been confirmed \cite{BH}.
441:
442:
443:
444: The spin-spin correlation function satisfies the equation
445: \be
446: {\cal D}{\cal G}=\gamma{\cal G}\,,
447: \label{RGE}
448: \end{equation}
449: where ${\cal D}$ is the renormalization group (RG) operator
450: \be
451: {\cal D}=-a{\partial\over\partial a}+D=
452: -a{\partial\over\partial a}
453: +\beta_\alpha(\alpha_0,\delta_0){\partial\over\partial\alpha_0}
454: +\beta_\delta(\alpha_0,\delta_0){\partial\over\partial\delta_0}
455: \end{equation}
456: and the $\beta$ and $\gamma$-functions are given by
457: \ba
458: \beta_\alpha(\alpha_0,\delta_0)&=&-g_1\alpha_0\delta_0-g_2\alpha_0^3-
459: g_3\alpha_0\delta_0^2+\cdots\,,\label{beta1}\\
460: \beta_\delta(\alpha_0,\delta_0)&=&f_1\alpha_0^2+(f_1+f_2)\alpha_0^2\delta_0
461: +\cdots\,,\label{beta2}\\
462: \gamma(\alpha_0,\delta_0)&=&-{1\over4}+{1\over4}\delta_0-{1\over4}\delta_0^2+
463: h_1\alpha_0^2+\cdots\,.\label{gamma}
464: \ea
465:
466: Now, as is well known, not all $\beta$-function coefficients are universal.
467: For example, under a redefinition
468: \ba
469: \tilde\alpha_0&=&\alpha_0+L\alpha_0\delta_0+\cdots\,,\label{Redef1}\\
470: \tilde\delta_0&=&\delta_0+K\alpha_0^2+\cdots\,,\label{Redef2}
471: \ea
472: the coefficients change as
473: \ba
474: \tilde g_1&=&g_1\,,\phantom{-Kg_1-Lf_1,}\qquad\qquad\tilde f_1=f_1\,,
475: \nonumber\\
476: \tilde g_2&=&g_2-Kg_1-Lf_1\,,\qquad\qquad\tilde f_2=f_2
477: -2Kg_1-2Lf_1\,,\label{redef2}\\
478: \tilde g_3&=&g_3\,.\nonumber
479: \ea
480: ((\ref{Redef1}-\ref{Redef2}) is the most
481: general perturbative redefinition respecting
482: the $\alpha_0\leftrightarrow-\alpha_0$ symmetry together with the requirement
483: that for $\alpha_0=0$ $\delta_0$ is not redefined.)
484: From (\ref{redef2}) we see that
485: in addition to the one-loop coefficients $g_1$ and $f_1$ there exist also
486: two-loop invariants. They are $g_3$ and the combination
487: \be
488: J=2g_2-f_2\,.
489: \label{J}
490: \end{equation}
491:
492: Two important physical quantities are the correlation length
493: \be
494: \xi={1\over Ma}=e^{\Psi(\alpha_0,\delta_0)}\,,
495: \label{xi}
496: \end{equation}
497: where $M$ is the mass of the physical particle
498: and the dimensionless susceptibility
499: \be
500: \chi={1\over a^2}\int d^2 x\,{\cal G}(x)=e^{\Phi(\alpha_0,\delta_0)}\,.
501: \label{chi}
502: \end{equation}
503: From (\ref{RGE}) it follows that the exponents satisfy
504: \ba
505: D\Psi&=&1\,,\label{Psi}\\
506: D\Phi&=&2+\gamma\,.\label{Phi}
507: \ea
508:
509: It is useful to introduce the RG invariant quantity $Q(\alpha_0,\delta_0)$
510: which satisfies $DQ=0$. Introducing the inverse function $k(\delta_0,Q)$
511: that satisfies
512: \be
513: \alpha_0=k(\delta_0,Q(\alpha_0,\delta_0))\,,
514: \end{equation}
515: we can define new $\beta$- and $\gamma$-functions:
516: \ba
517: b(\delta_0,Q)&=&\beta_\delta(k(\delta_0,Q),\delta_0)\,,\label{b}\\
518: \Gamma(\delta_0,Q)&=&\gamma_\delta(k(\delta_0,Q),\delta_0)\,.\label{Gamma}
519: \ea
520: The advantage of using the variables $\delta_0$ and $Q$ is that the RG
521: invariant $Q$ can be treated in many respects as if it were a numerical
522: constant and $\delta_0$ were a single coupling constant. For example,
523: if we write
524: \be
525: \Psi(\alpha_0,\delta_0)=\Psi_1(\delta_0,Q(\alpha_0,\delta_0))
526: \end{equation}
527: and
528: \be
529: \Phi(\alpha_0,\delta_0)=\Phi_1(\delta_0,Q(\alpha_0,\delta_0))
530: \end{equation}
531: then the functions $\Psi_1$ and $\Phi_1$ can be determined from
532: \be
533: {\partial\over\partial\delta_0}\Psi_1(\delta_0,Q)={1\over b(\delta_0,Q)}
534: \label{Psi1diff}
535: \end{equation}
536: and
537: \be
538: {\partial\over\partial\delta_0}\Phi_1(\delta_0,Q)=
539: {2+\Gamma(\delta_0,Q)\over b(\delta_0,Q)}
540: \label{Phi1diff}
541: \end{equation}
542: respectively.
543:
544: Using (\ref{beta1}) and (\ref{beta2}) $Q$ can be determined.
545: \be
546: Q(\alpha_0,\delta_0)=\frac{1}{32}\alpha_0^2-2\delta_0^2+
547: 2g_2\alpha_0^2\delta_0+F_2\delta_0^3+\cdots
548: \label{Q}
549: \end{equation}
550: and using this in (\ref{b}) we find
551: \be
552: b(\delta_0,Q)=Q+2\delta_0^2+AQ\delta_0+B\delta_0^3+\cdots\,,
553: \label{bb}
554: \end{equation}
555: while the $\Gamma$ function to this order is
556: \be
557: \Gamma(\delta_0,Q)=-{1\over4}+{1\over4}\delta_0+\cdots\,.
558: \end{equation}
559: Here
560: \be
561: A=1-\frac{J}{f_1}\,,\qquad\qquad
562: B=-\frac{2}{3}g_3+\frac{2}{3}A\,,\qquad\qquad
563: F_2=\frac{2}{3}g_3+\frac{4}{3}A
564: \label{ABF2}
565: \end{equation}
566: or numerically
567: \be
568: A=3\,,\qquad\qquad B=2\,,\qquad\qquad F_2=4\,,
569: \label{ABF2num}
570: \end{equation}
571: but we will keep these constants in the following to explicitly
572: demonstrate that all our results are universal.
573:
574: We know that a one-parameter renormalizable subspace
575: in the $\delta_0$--$\alpha_0$ plane is equivalent to the SU(2)
576: chiral Gross-Neveu model. (This is most evident in the fermionic formulation.)
577: This subspace must correspond to the $Q=0$ RG trajectory
578: because we know that it goes through the point
579: $\alpha_0=\delta_0=0$. Moreover, it must be the $\delta_0<0$ branch
580: of the $Q=0$ trajectory, since it is the one that is asymptotically
581: free in perturbation theory.
582:
583: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
584: \begin{figure}[htb]
585: \leavevmode
586: \epsfxsize=170mm
587: %\epsfysize=19.5cm
588: \hspace{-1.0cm}
589: \epsfbox{PD.eps}
590:
591: \vspace{0.1cm}
592:
593: \caption{\footnotesize
594: Phase diagram of the KT theory. The entire Region I is
595: critical. The model is massive in Region III ($Q>0$) and
596: Region II ($Q<0$). The dotted lines are RG trajectories.
597: The full lines $S_1$ and $S_2$ correspond to the critical
598: surface of the XY model and the (bare) CGN model respectively.
599: The dashed curve is the XY model hitting the critical surface
600: at $c$.
601: }
602: \label{extra1}
603: \end{figure}
604:
605:
606: Following \cite{Amit} the phase diagram of our model is represented
607: in Figure 1.
608: The CGN model corresponds to the separatrix $S_2$ on this plot,
609: which will be referred to as PD for short. Region III
610: corresponds to $Q>0$, whereas Regions I and II correspond to
611: $Q<0,\delta_0>0$ and $Q<0,\delta_0<0$, respectively.
612:
613: In the neighbourhood of $S_2$, i.e. close to the CGN case
614: we have
615: \be
616: b(\delta_0,Q)=b_0(\delta_0)+Qb_1(\delta_0)+\cdots\,,
617: \end{equation}
618: where
619: \ba
620: b_0(\delta_0)&=&2\delta_0^2+B\delta_0^3+\cdots\,,\label{b0}\\
621: b_1(\delta_0)&=&1+A\delta_0+\cdots\,.\label{b1}
622: \ea
623: It is a crucial observation that near the CGN line the correlation
624: length exponent $\Psi_1$ is a smooth analytic function of $Q$:
625: \be
626: \Psi_1(\delta_0,Q)=H_0(\delta_0)+QH_1(\delta_0)+\cdots\,,
627: \label{Psi1CGN}
628: \end{equation}
629: where
630: \ba
631: H_0(\delta_0)&=&-{1\over2\delta_0}-\frac{B}{4}\ln\vert2\delta_0\vert+
632: \Psi_0+\cdots\,,\\
633: H_1(\delta_0)&=&{1\over12\delta_0^3}+\frac{A-B}{8\delta_0^2}+\cdots\,.
634: \ea
635: We will calculate the value of the nonperturbative constant $\Psi_0$
636: in Section 5.
637:
638: Now we can integrate (\ref{Psi1diff}) using the perturbative expansion
639: (\ref{bb}) and for $Q>0$ we get
640: \ba
641: \Psi_1(\delta_0,Q)&=&{1\over\sqrt{2Q}}\Bigg({\pi\over2}+{\rm arctg}
642: {\sqrt{2}\delta_0\over\sqrt{Q}}\Bigg)-\frac{B}{8}\ln2(Q+2\delta_0^2)
643: \nonumber\\
644: &+&\frac{2A-B}{8}{Q\over Q+2\delta_0^2}+\Psi_0+\cdots\,.\label{pertI}
645: \ea
646: Here the dots stand for higher order terms in the perturbative expansion
647: (these are, in principle, calculable) and also for an unknown
648: (nonperturbative) function of $Q$ only. For small $Q$ (\ref{pertI})
649: becomes
650: \be
651: \Psi_1(\delta_0,Q)={\pi\over\sqrt{2Q}}\,\Theta(\delta_0)+
652: H_0(\delta_0)+QH_1(\delta_0)+\cdots\,,
653: \label{smallQ}
654: \end{equation}
655: where the dots stand for terms higher order in Q. They come from the
656: higher perturbative terms of (\ref{pertI}) and also from the
657: nonperturbative function mentioned above. The point is that there are
658: no terms, singular in $Q$, coming from any of these two sources. This
659: is obvious for the perturbative terms, but must also be true for the
660: nonperturbative contributions, since otherwise $\Psi_1$ would be
661: singular on the $S_2$ line. Requiring it to be nonsingular on $S_2$,
662: we force $\Psi_1$ to diverge on $S_1$, which is therefore part of the
663: critical surface of the phase diagram PD. An other line, on which we
664: know the correlation length must diverge is the $\delta_0$
665: axis $\alpha_0=0$, because this corresponds to a free, massless model.
666:
667: For $Q<0$, it is convenient to parametrize $Q$ in terms of the
668: $\delta_0$ value at which the RG trajectory intersects this axis.
669: In other words, we have to express $Q$ in terms of $d$ that solves
670: \be
671: b(d,Q)=0\,.
672: \label{d}
673: \end{equation}
674: (Note that in this parametrization $\vert\delta_0\vert\geq\vert d\vert$,
675: because of $\alpha_0^2\geq0$.)
676: The perturbative solution of (\ref{d}) is
677: \be
678: Q=-2d^2+(2A-B)d^3+\cdots\,.
679: \label{Qsol}
680: \end{equation}
681: Using this parametrization the perturbative solution for $Q<0$ is
682: \ba
683: \Psi_1(\delta_0,Q)&=&-{1\over4d}\ln\Big({\delta_0+d\over\delta_0-d}
684: \Big)\Big[1+\frac{2A-3B}{4}d\Big]
685: +\frac{2A-B}{8}{d\over\delta_0+d}\nonumber\\
686: &-&\frac{B}{4}\ln2\vert\delta_0+d\vert+y(d)+\cdots\,,\label{pertII}
687: \ea
688: where
689: \ba
690: {\rm for\ }d<0\qquad y(d)&=&\Psi_0+\cdots\,,\label{dneg}\\
691: {\rm for\ }d>0\qquad y(d)&=&\infty\,.\label{dpos}
692: \ea
693: (\ref{dneg}) is obtained by matching (\ref{pertII}) to
694: (\ref{Psi1CGN}) for small $Q$, while (\ref{dpos}) is formally true
695: since this is the only way to achieve that the correlation length
696: diverges on the positive half of the $\delta_0$ axis.
697: (Without the infinite constant $\xi$ actually {\it vanishes}
698: there.)
699:
700: Now we can discuss the phase diagram of our model. The entire Region I
701: is critical. The massive phase is Regions II and III and the critical
702: surface bordering them is $S_1$ plus the negative part of the $\delta_0$
703: axis. They are smoothly connected across $S_2$, which is the (bare)
704: CGN model. The $O(2)$ NLS model corresponds to the dashed curve of PD.
705: In a MC experiment we are approaching the critical point $c$ from the
706: massive phase (Region III). We will denote the $\delta_0$ coordinate of
707: $c$ by $d_0$. Because the RG trajectories are running
708: basically parallel to $S_1$, it is physically irrelevant at which point
709: the critical surface is reached and therefore the parameter $d_0$ is
710: irrelevant. The continuum model will be the same for all points on
711: $S_1$, including the origin. But the origin is the point, where
712: (coming along $S_2$) the continuum CGN model is defined! So our
713: continuum theory is inevitably identical to the (massive part of the)
714: $SU(2)$-invariant CGN model.
715:
716: If we start from somewhere in the middle of Region II, we can define
717: a massive continuum limit by approaching the negative half of the
718: $\delta_0$ axis. The intercept $d$ is then relevant. The continuum
719: theory is the SG model with
720: \be
721: \beta^2=8\pi(1+d).
722: \label{betad}
723: \end{equation}
724:
725: Returning to the $O(2)$ model, the dashed trajectory can be parametrized
726: as
727: \ba
728: \delta_0&=&d_0+d_1\tau+\cdots\,,\\
729: \alpha_0&=&k(d_0,0)+\alpha_1\tau+\cdots\,,
730: \ea
731: where
732: \be
733: \tau=K_c-K
734: \label{tau}
735: \end{equation}
736: is the reduced coupling and we have assumed that physical
737: quantities are analytic in $K$. ($K_c=1.1197(5)$
738: \cite{Hasenbusch}.) Then also $Q$ is analytic in $\tau$:
739: \be
740: Q\sim\tau\,.
741: \end{equation}
742: From (\ref{smallQ}) we see that along the $O(2)$ curve
743: \be
744: \ln\xi={\pi\over\sqrt{2Q}}+H_0(d_0)+\cdots\,,
745: \label{logxi}
746: \end{equation}
747: where the dots stand for terms analytic in $Q$ (and vanishing for $Q=0$).
748: In other words, for the $O(2)$ model \cite{KT}
749: \be
750: \xi=C\exp\Big\{\frac{b}{\sqrt{\tau}}\Big\}\big(1+
751: {\cal O}(\sqrt{\tau})\big),
752: \label{xiKT}
753: \end{equation}
754: where the constants $C$ and $b$ are not universal. This is the famous KT
755: formula showing that the phase transition is of infinite order in the
756: reduced temperature.
757:
758: It is more important for us that (\ref{logxi}) can be rewritten as
759: \be
760: Q={\pi^2\over2(\ln\xi +u)^2}+{\cal O}\big((\ln\xi)^{-5}\big),
761: \label{Qdef}
762: \end{equation}
763: where
764: \be
765: u=-H_0(d_0)\,,
766: \label{u}
767: \end{equation}
768: which is given (if $d_0$ is sufficiently small) perturbatively by
769: \be
770: u\approx{1\over2 d_0}+\frac{B}{4}\ln(2d_0)-\Psi_0\,.
771: \end{equation}
772: Note that the leading $1/(\log\xi)^2$
773: term in (\ref{Qdef}) is universal and only the subleading terms
774: (depending on the value of the parameter $u$) are model-dependent.
775:
776: The susceptibility exponent $\Phi_1$ can be studied similarly. It can
777: be written as
778: \be
779: \Phi_1={7\over4}\,\Psi_1+\tilde\Phi_1={7\over4}\,\Psi_1+
780: {1\over16}\ln(2\delta_0^2+Q)+\cdots\,,
781: \end{equation}
782: where $\tilde\Phi_1$ satisfies
783: \be
784: {\partial\tilde\Phi_1\over\partial\delta_0}={\Gamma(\delta_0,Q)+{1\over4}
785: \over b(\delta_0,Q)}\,.
786: \end{equation}
787:
788: Now the crucial observation is again that, for small $Q$,
789: \be
790: \Phi_1={7\over4}\,\Psi_1+c_1+c_2Q+\cdots\,,
791: \label{Phi1smallQ}
792: \end{equation}
793: because the (calculable) perturbative terms are analytic in $Q$, while
794: the (not calculable) purely $Q$-dependent terms in $\tilde\Phi_1$ must
795: also be analytic in $Q$ otherwise these latter singularities would
796: also turn up for the CGN line $S_2$, where they must not.
797:
798: From (\ref{Phi1smallQ}) we have
799: \be
800: \chi=\xi^{{7\over4}}\,e^{c_1}\Big(1+c_2Q+\cdots\Big)\,,
801: \label{chixi}
802: \end{equation}
803: i.e. there are no (multiplicative) log corrections in the scaling
804: relation for the susceptibility. The possibility of such
805: multiplicative logarithmic corrections is discussed in \cite{JankeKI}.
806:
807: \section{Determination of the lattice artifacts}
808:
809:
810: Recall that the RG invariant $Q$ has
811: a completely different meaning for Region III (which contains the
812: massive phase of the XY model) and for Region II (where the usual
813: massive SG model with $\beta^2<8\pi$ can be defined). Indeed, in
814: the positive $\delta_0$ part of Region III,
815: close to $S_1$, the (positive) parameter $Q$ merely measures the
816: distance from the XY
817: critical surface on which it vanishes, whereas in Region II $Q$ is
818: a relevant (negative) parameter related to the SG coupling $\beta$
819: by (\ref{d}) and (\ref{betad}).
820: Our main assumption is that in spite of this difference
821: physical quantities are smoothly depending on $Q$ in the vicinity of the
822: separatrix $S_2$ connecting the two regions. More precisely, we will
823: assume that close to $S_2$ any physical quantity $U$ has the form
824: \be
825: U(Q)=U_0+U_1Q+{\cal O}(Q^2).
826: \label{Qexp}
827: \end{equation}
828: Here $U_0=U(0)$ is its value for the CGN model (and thus also in the
829: continuum limit of the XY model). The first correction coefficient $U_1$
830: can be calculated from the SG model as follows. Using the identification
831: (\ref{d}) and its perturbative solution (\ref{Qsol}) together with
832: (\ref{betad}) and (\ref{gammapar}) we have
833: \be
834: Q=-2\nu^2 +{\cal O}(\nu^4).
835: \label{Qgamma}
836: \end{equation}
837: This means that if in the SG model, close to the CGN point $\nu=0$,
838: we have
839: \be
840: U(\nu)=u_0+u_1\nu^2+{\cal O}(\nu^4),
841: \label{gammaexp}
842: \end{equation}
843: then
844: \be
845: U_0=u_0\qquad\qquad{\rm and}\qquad\qquad U_1=-u_1/2.
846: \label{u0u1}
847: \end{equation}
848: Translated to the language of lattice artifacts by (\ref{Qdef}) we thus
849: have
850: \be
851: U(\xi)=u_0-\frac{u_1\pi^2}{4(\ln\xi+u)^2}+{\cal O}\big((\ln\xi)^{-4}\big).
852: \label{xiexp}
853: \end{equation}
854: This means that lattice artifacts typically go away very slowly, only as
855: $1/(\log\xi)^2$. On the other hand the leading artifacts are universal and
856: calculable.
857:
858: We apply this method first to the scattering phase shifts. Recall
859: the SG model S-matrix (\ref{OnSmatrix}) with (\ref{sig2SG}).
860: The three distinct S-matrix eigenvalues are
861: \ba
862: s^{(0)}_{\rm SG}(\theta)=2\sigma_1(\theta)+\sigma_2(\theta)+\sigma_3(\theta)
863: &=&\phantom{-}\frac
864: {\sinh\frac{\nu}{2}(i\pi+\theta)}
865: {\sinh\frac{\nu}{2}(i\pi-\theta)}s^{(2)}_{\rm SG}(\theta),\label{s0}\\
866: s^{(1)}_{\rm SG}(\theta)=\sigma_2(\theta)-\sigma_3(\theta)
867: &=&-\frac
868: {\cosh\frac{\nu}{2}(i\pi+\theta)}
869: {\cosh\frac{\nu}{2}(i\pi-\theta)}s^{(2)}_{\rm SG}(\theta)\label{s1}
870: \ea
871: and
872: \be
873: s^{(2)}_{\rm SG}(\theta)=\sigma_2(\theta)+\sigma_3(\theta)
874: =-\exp\Big\{2i\int_0^\infty\frac{d\omega}{\omega}
875: \sin(\theta\omega)\,\tilde k(\omega)\Big\}
876: \label{s2}
877: \end{equation}
878: where $\tilde k(\omega)$ is given by (\ref{tildek}).
879:
880: We now write
881: \be
882: s^{(i)}_{\rm SG}(\theta)=\exp\big\{2i\delta_0^{(i)}(\theta)+2i
883: \nu^2\delta_1^{(i)}(\theta)+{\cal O}(\nu^4)\big\}
884: \label{phaseshifts}
885: \end{equation}
886: for $i=0,1$ and $2$. Here $\delta^{(i)}_0$ are the CGN phase shifts
887: which, as remarked before, coincide with the $n\to2$ limit of the
888: $O(n)$ phase shifts.
889: The first correction coefficients can be obtained by a simple
890: calculation. The result is
891: \be
892: \delta^{(0)}_1(\theta)=0,\qquad
893: \delta^{(1)}_1(\theta)=\frac{\pi\theta}{6},\qquad
894: \delta^{(2)}_1(\theta)=-\frac{\pi\theta}{12}.
895: \label{1correction}
896: \end{equation}
897: This can be used to obtain the leading lattice artifacts in the
898: XY model by the relation (\ref{xiexp}).
899:
900:
901: \section{Current-current 2-point function and free energy}
902:
903: Consider the 2-point function of the Noether current
904: \be
905: J_\mu=i{\beta_0\over2\pi}\epsilon_{\mu\nu}\partial_\nu\phi\,.
906: \label{defJ}
907: \end{equation}
908: Its Fourier transform $I(p)$ is defined by
909: \be
910: \langle J_\mu(x)J_\nu(y)\rangle=\int {d^2p\over(2\pi)^2}\,
911: e^{-ip(x-y)}{p_\mu p_\nu-p^2\delta_{\mu\nu}\over p^2}\,I(p)\,.
912: \end{equation}
913: It is easy to calculate $I(p)$ to second order:
914: \be
915: I(p)={2\over\pi}\Bigg\{1+\delta_0+\alpha_0^2\Big[\Omega(p,a)+\cdots
916: \Big]\Bigg\}\,,
917: \label{curr}
918: \end{equation}
919: where
920: \be
921: \Omega(p,a)=f_1(\ln pa +\Omega_0)
922: \label{Omega}
923: \end{equation}
924: and
925: \be
926: \Omega_0=C-{1\over2}-\ln2\,,
927: \label{Omega0}
928: \end{equation}
929: $C$ being Euler's constant.
930:
931: Standard RG considerations give
932: \be
933: I(p,\delta_0,Q,a)=I(p_0,\bar\lambda,Q,a)=\tilde I(p/M,Q)\,,
934: \label{RGI}
935: \end{equation}
936: where the running coupling is the solution of
937: \be
938: \Psi_1(\bar\lambda,Q)=\ln{p\over M}-\ln p_0a\,.
939: \label{lambdabar}
940: \end{equation}
941:
942: Let us consider the $Q<0$, $\delta_0<0$ case (Region II) first.
943: For $p\rightarrow\infty$ also $\Psi_1\rightarrow\infty$ and therefore
944: $\bar\lambda\rightarrow d$, where $d$ is defined by
945: \be
946: b(d,Q)=0\qquad\qquad{\rm or\ equivalently}
947: \qquad\qquad\alpha_0=k(d,Q)=0\,.
948: \end{equation}
949: This gives
950: \be
951: \tilde I(\infty,Q)={2\over\pi}(1+d)\,,
952: \label{Iinfty}
953: \end{equation}
954: where $d$ parametrizes $Q$ according to (\ref{Qsol}).
955: (\ref{Iinfty}) is consistent with the identification (\ref{betad}).
956:
957: Next we study the case of small $Q$, because this is relevant if
958: we are interested in the approach to the continuum limit along
959: the $O(2)$ curve. We assume that $\tilde I$ is analytic in $Q$:
960: \be
961: \tilde I(z,Q)=\tilde I_0(z)+Q\tilde I_1(z)+\cdots\,.
962: \label{Iexpand}
963: \end{equation}
964: It is a standard exercise to obtain the asymptotic
965: expansion of the coefficients $\tilde I_0$ and $\tilde I_1$ in perturbation
966: theory. One gets
967: \ba
968: \tilde I_0(p/M)&=&{2\over\pi}\Bigg\{1-\lambda+2\kappa\lambda^2+\cdots\Bigg\}
969: \,,\label{I0}\\
970: \tilde I_1(p/M)&=&{2\over3\pi}\Bigg\{{1\over2\lambda}+\kappa
971: +\frac{2B-3A}{4}+\cdots\Bigg\}\,,\label{I1}
972: \ea
973: where
974: \be
975: \kappa=\Omega_0-\Psi_0
976: \label{kappa}
977: \end{equation}
978: and the coupling $\lambda$ is the solution of
979: \be
980: {1\over2\lambda}-\frac{B}{4}\ln(2\lambda)=\ln{p\over M}\,.
981: \end{equation}
982: The asymptotic expansions (\ref{I0}) and (\ref{I1}) are valid for
983: $\lambda\ll1$, i.e. for $p\rightarrow\infty$ but also $Q$ must
984: satisfy
985: \be
986: Q\ll6\lambda^2
987: \label{Qsmaller}
988: \end{equation}
989: so that the expansion (\ref{Iexpand}) makes sense.
990:
991: It is by now standard how the nonperturbative constant $\kappa$
992: can be calculated. For this it is necessary to consider the free
993: energy in an external field and we now turn to this calculation.
994: We follow here \cite{Zamolodchikov} and start from the modified
995: Lagrangian
996: \be
997: {\cal L}_h={1\over2}(\partial_\mu\phi)^2-{\alpha_0\over\beta_0^2 a^2}
998: \cos(\beta_0\phi)+{\beta_0 h\over2\pi}\partial_1\phi\,,
999: \label{Lagh}
1000: \end{equation}
1001: which corresponds to adding a term $ihJ_2$ to the Lagrangian density.
1002: The modified ground state energy must be of the form
1003: \be
1004: {\cal E}(h)=-{h^2\over\pi}\,{\cal F}(h)\,,
1005: \end{equation}
1006: where ${\cal F}(h)$ is dimensionless.
1007:
1008: In perturbation theory we get
1009: \be
1010: {\cal F}(h)=1+\delta_0+\alpha_0^2\Big[\tilde\Omega(h,a)+\cdots
1011: \Big]\,,
1012: \label{calF}
1013: \end{equation}
1014: where
1015: \be
1016: \tilde\Omega(h,a)=f_1(\ln ha +\tilde\Omega_0)
1017: \label{TOmega}
1018: \end{equation}
1019: and
1020: \be
1021: \tilde\Omega_0=C-{1\over2}+\ln2\,.
1022: \label{TOmega0}
1023: \end{equation}
1024: For $Q=0$ therefore ${\cal F}(h)$ has the following asymptotic expansion
1025: \be
1026: {\cal F}(h)=1-{1\over2\ln{h\over M}}-\frac{B}{8}
1027: {\ln\ln {h\over M}\over\ln^2{h\over M}}+
1028: {\kappa+\ln4\over2\ln^2{h\over M}}+
1029: {\cal O}\Bigg({\Big(\ln\ln {h\over M}\Big)^2\over\ln^3{h\over M}}
1030: \Bigg)\,.
1031: \label{pertF}
1032: \end{equation}
1033: This is to be compared to the exact result \cite{Forgacs}
1034: \be
1035: {\cal F}(h)=1-{1\over2\ln{h\over M}}-
1036: {\ln\ln {h\over M}\over4\ln^2{h\over M}}+
1037: {{3\over2}\ln2-{1\over4}-{1\over2}\ln\pi
1038: \over2\ln^2{h\over M}}+
1039: {\cal O}\Bigg({\Big(\ln\ln {h\over M}\Big)^2\over\ln^3{h\over M}}
1040: \Bigg)\,.
1041: \label{exF}
1042: \end{equation}
1043: It is a nontrivial check on the overall consistency of our considerations
1044: that using (\ref{ABF2num}) for the numerical value of our
1045: universal constant $B$ the third term in (\ref{pertF}) matches the
1046: corresponding one in (\ref{exF}).
1047: Comparing (\ref{pertF}) to (\ref{exF}) we also get
1048: \be
1049: \kappa=-{1\over4}-\ln\sqrt{2\pi}\,.
1050: \label{kappaEX}
1051: \end{equation}
1052: Using the exact value (\ref{kappaEX}) we obtain the asymptotic
1053: expansion of the first correction in $Q$:
1054: \be
1055: \tilde I_1(p/M)={2\over3\pi}\Bigg\{{1\over2\lambda}-{3\over2}-
1056: \ln\sqrt{2\pi}\Bigg\}={2\over3\pi}\Bigg\{{1\over2\lambda}-2.419
1057: \Bigg\}\,.
1058: \label{deltaI}
1059: \end{equation}
1060: Note that the leading correction term (\ref{deltaI}) does not contain any
1061: free parameter.
1062:
1063: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1064:
1065:
1066:
1067: \vspace{1cm}
1068: {\tt Acknowledgements}
1069:
1070: \noindent
1071: This investigation was supported in part by the Hungarian National
1072: Science Fund OTKA (under T030099 and T029802). I thank the members of the
1073: \lq$\sigma$ collaboration',
1074: M. Niedermaier, F. Niedermayer, A. Patrascioiu, E. Seiler and P. Weisz,
1075: for many thorough and interesting discussions. I also thank the
1076: Max-Planck-Institut f\"ur Physik for hospitality.
1077:
1078:
1079:
1080: \clearpage
1081:
1082:
1083: \begin{thebibliography}{99}
1084:
1085: \bibitem{KT}
1086: J. M. Kosterlitz and D. J. Thouless,
1087: J. Phys. {\bf C6} (1973) 1181;\\
1088: J. M. Kosterlitz,
1089: J. Phys. {\bf C7} (1974) 1046.
1090:
1091: \bibitem{ZinnJ}
1092: J. Zinn-Justin, {\it Quantum Field Theory and Critical
1093: Phenomena}, Oxford, 1989.
1094:
1095: \bibitem{ZZ}
1096: A. B. and Al. B. Zamolodchikov,
1097: Ann. Phys. {\bf 120} (1979) 253; \\
1098: Nucl. Phys. {\bf B133} (1978) 525.
1099:
1100: \bibitem{Sym}
1101: K. Symanzik, Nucl. Phys. {\bf B226} (1983) 187.
1102:
1103: \bibitem{Amit}
1104: D. J. Amit, Y. Y. Goldschmidt and G. Grinstein, J. Phys. {\bf A13}
1105: (1980) 585.
1106:
1107: \bibitem{HMN}
1108: P.~Hasenfratz, M.~Maggiore and F.~Niedermayer,
1109: Phys. Lett. {\bf B245} (1990) 522;\\
1110: P.~Hasenfratz and F.~Niedermayer,
1111: Phys. Lett. {\bf B245} (1990) 529.
1112:
1113: \bibitem{paper3}
1114: J.~Balog, M.~Niedermaier, F.~Niedermayer, A.~Patrascioiu, E.~Seiler
1115: and P.~Weisz, in preparation.
1116:
1117: \bibitem{Poly}
1118: A. Polyakov,
1119: Phys. Lett. {\bf B72} (1977) 224.
1120:
1121: \bibitem{Lusch}
1122: M.~L\"uscher,
1123: Nucl. Phys. {\bf B135} (1978) 1.
1124:
1125: \bibitem{pew}
1126: B. Berg and P. Weisz, Nucl. Phys. {\bf B146} (1979) 205;\\
1127: E. Abdalla, B. Berg and P. Weisz, Nucl. Phys. {\bf B157} (1979) 387.
1128:
1129: \bibitem{Villain}
1130: J. Villain, J. Physique {\bf 36} (1975) 581.
1131:
1132: \bibitem{Jose}
1133: J. V. Jos\'e et al., Phys. Rev. {\bf B16} (1977) 1217.
1134:
1135: \bibitem{Hasenbusch}
1136: M. Hasenbusch, M. Marcu and K. Pinn, Physica {\bf A208} (1994) 124.
1137:
1138: \bibitem{FrohSpen}
1139: J. Fr\"ohlich and T. Spencer, Comm. Math. Phys. {\bf 81} (1981) 527.
1140:
1141: \bibitem{Banksetal}
1142: T. Banks, D. Horn and H. Neuberger, Nucl. Phys. {\bf B108} (1976) 119.
1143:
1144: \bibitem{Coleman}
1145: S. Coleman, Phys. Rev. {\bf D11} (1975) 2088.
1146:
1147: \bibitem{BH}
1148: J. Balog and \'A. Heged\H us, J. Phys. {\bf A33} (2000) 6543.
1149:
1150: \bibitem{JankeKI}
1151: R. Kenna and A. C. Irving, Nucl. Phys. {\bf B485} (1997) 583;\\
1152: Phys. Lett. {\bf B351} (1995) 273.\\
1153: W. Janke, {\tt hep-lat/9609045}.
1154:
1155: \bibitem{Zamolodchikov}
1156: Al. B. Zamolodchikov, Int. J. of Mod. Phys. {\bf A10} (1995) 1125.
1157:
1158: \bibitem{Forgacs}
1159: P. Forg\'acs, S. Naik and F. Niedermayer, Phys. Lett.
1160: {\bf B283} (1992) 282.
1161:
1162:
1163:
1164:
1165:
1166: \end{thebibliography}
1167:
1168: \end{document}
1169:
1170: