1: \documentclass[12pt]{article}
2: %\usepackage{showkeys}
3: \usepackage{epsf,amssymb,psfrag,amsmath}
4: %\usepackage{hyperref}
5: %\usepackage[active]{srcltx}
6: \catcode`\@=11
7: %-------------------------------------------------------------
8: \textwidth 173mm \textheight 235mm \topmargin -50pt \oddsidemargin -0.45cm \evensidemargin -0.45cm
9: %-------------------------------------------------------------
10: %\def \thesection {\arabic{section}.}
11: %\def \thesubsection {\thesection\arabic{subsection}.}
12: %\def \thesubsubsection {\thesubsection\arabic{subsubsection}.}
13: %-------------------------------------------------------------
14: \def \be {\begin{equation}}
15: \def \ee {\end{equation}}
16: \def \ba {\begin{eqnarray}}
17: \def \ea {\end{eqnarray}}
18: \def \baa {\begin{eqnarray*}}
19: \def \eaa {\end{eqnarray*}}
20: \def \bb {\begin {thebibliography} }
21: \def \eb {\end{thebibliography}}
22: %\def \lab #1 {\label{#1} \mbox{\# ${#1}$}}
23: \def \lab #1 {\label{#1}}
24: \newcommand \ci [1] {\cite{#1}}
25: \newcommand \bi [1] {\bibitem{#1}}
26: \newcommand\re[1]{({\ref{#1}})}
27: \def \qqquad {\qquad\quad}
28: \def \qqqquad {\qquad\qquad}
29: %-------------------------------------------------------------
30: \def \matrix #1 {\left(\begin{array}{cc} #1 \end{array}\right)}
31: \def \Tr {\mathop{\rm Tr}\nolimits}
32: \def \tr {\mathop{\rm tr}\nolimits}
33: \def \Im {\mathop{\rm Im}\nolimits}
34: \def \Re {\mathop{\rm Re}\nolimits}
35: \def \res{\mathop{\rm res}\nolimits}
36: \def \e {\mathop{\rm e}\nolimits}
37: \newcommand\lr[1]{{\left({#1}\right)}}
38: \newcommand \widebar [1] {\overline{#1}}
39: \newcommand\bin[2]{\left({#1}\atop{#2}\right)}
40: \newcommand \vev [1] {\langle{#1}\rangle}
41: \newcommand \VEV [1] {\left\langle{#1}\right\rangle}
42: \newcommand \ket [1] {|{#1}\rangle}
43: \newcommand \bra [1] {\langle {#1}|}
44: \newcommand \partder [1] {{\partial \over\partial #1}}
45: \newcommand{\as}{\ifmmode\alpha_{\rm s}\else{$\alpha_{\rm s}$}\fi}
46: \newcommand{\asbar}{\ifmmode\bar{\alpha}_{\rm s}\else{$\bar{\alpha}_{\rm s}$}\fi}
47: \newcommand{\ft}[2]{{\textstyle\frac{#1}{#2}}}
48: \newcommand{\inclfig}[2]{\mbox{\epsfxsize=#1cm \epsfbox{#2.ps}}}
49: \newcommand{\insertfig}[2]{\mbox{\epsfxsize=#1cm \epsfbox{#2.eps}}}
50:
51: \font\cmss=cmss12 \font\cmsss=cmss10 at 11pt
52: \def\inbar{\,\vrule height1.5ex width.4pt depth0pt}
53: \def\IC{\relax\hbox{$\inbar\kern-.3em{\rm C}$}}
54: \def\IZ{\relax{\hbox{\cmss Z\kern-.4em Z}}}
55: \def\IR{{\hbox{{\rm I}\kern-.2em\hbox{\rm R}}}}
56: \def\R{{\tiny \IR}}
57: \def\IP{{\hbox{{\rm I}\kern-.2em\hbox{\rm P}}}}
58: \def\II{\hbox{{1}\kern-.25em\hbox{l}}}
59:
60: \def\numberbysection{\@addtoreset{equation}{section}
61: \def\theequation{\thesection.\arabic{equation}}}
62: \numberbysection
63:
64: \newcommand \mybf[1] {\mbox{\boldmath$\scriptstyle {#1} $}}
65: \newcommand \Mybf[1] {\mbox{\boldmath$ {#1} $}}
66: \newbox\lett\newdimen\lheight\newdimen\lwidth
67: \def\ontop#1#2{\setbox\lett=\hbox{#2}\lheight\ht\lett
68: \multiply\lheight by 12 \divide\lheight by 10\relax%
69: \lwidth\wd\lett \multiply\lwidth by 8 \divide\lwidth by 10\relax #2\kern-\lwidth%
70: \raise\lheight\hbox{{$\scriptstyle #1$}}\kern.1ex}
71:
72:
73:
74: \begin{document}
75:
76:
77: \begin{titlepage}
78: \begin{flushright}
79: \begin{tabular}{l}
80: LPT--Orsay--06--04 \\ %[-1mm]
81: ITEP-TH-02/06 \\
82: hep-th/0601112
83: \end{tabular}
84: \end{flushright}
85:
86: \vskip2cm
87:
88: \centerline{\large \bf Logarithmic scaling in gauge/string correspondence}
89:
90: \vspace{1cm}
91:
92: \centerline{\sc A.V. Belitsky$^a$, A.S. Gorsky$^b$, G.P. Korchemsky$^c$}
93:
94: \vspace{10mm}
95:
96: \centerline{\it $^a$Department of Physics and Astronomy, Arizona State
97: University} \centerline{\it Tempe, AZ 85287-1504, USA}
98:
99: \vspace{3mm}
100:
101: \centerline{\it $^b$Institute of Theoretical and Experimental Physics }
102: \centerline{\it B. Cheremushkinskaya ul. 25, 117259 Moscow, Russia}
103:
104: \vspace{3mm}
105:
106: \centerline{\it $^c$Laboratoire de Physique Th\'eorique\footnote{Unit\'e
107: Mixte de Recherche du CNRS (UMR 8627).},
108: Universit\'e de Paris XI}
109: \centerline{\it 91405 Orsay C\'edex, France}
110:
111:
112: \def\thefootnote{\fnsymbol{footnote}}%
113: \vspace{1cm}
114:
115: \centerline{\bf Abstract}
116:
117: \vspace{5mm}
118:
119: We study anomalous dimensions of (super)conformal Wilson operators at weak and strong
120: coupling making use of the integrability symmetry on both sides of the gauge/string
121: correspondence and elucidate the origin of their single-logarithmic behavior for long
122: operators/strings in the limit of large Lorentz spin. On the gauge theory side, we
123: apply the method of the Baxter $Q-$operator to identify different scaling regimes in
124: the anomalous dimensions in integrable sectors of (supersymmetric) Yang-Mills theory
125: to one-loop order and determine the values of the Lorentz spin at which the logarithmic
126: scaling sets in. We demonstrate that the conventional semiclassical approach based on
127: the analysis of the distribution of Bethe roots breaks down in this domain. We work out
128: an asymptotic expression for the anomalous dimensions which is valid throughout the
129: entire region of variation of the Lorentz spin. On the string theory side, the logarithmic
130: scaling occurs when two most distant points of the folded spinning string approach the
131: boundary of the AdS space. In terms of the spectral curve for the classical string sigma
132: model, the same configuration is described by an elliptic curve with two branching points
133: approaching values determined by the square root of the 't Hooft coupling constant. As a
134: result, the anomalous dimensions cease to obey the BMN scaling and scale logarithmically
135: with the Lorentz spin.
136:
137: \end{titlepage}
138:
139: \setcounter{footnote} 0
140:
141: %\thispagestyle{empty}
142:
143: \newpage
144:
145: \pagestyle{plain} \setcounter{page} 1
146:
147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148: \section{Introduction}
149: \label{Introduction}
150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
151:
152: It is well known that in four-dimensional gauge theories the anomalous dimensions
153: of composite Wilson operators carrying a large Lorentz spin scale (at most)
154: logarithmically with the spin. This result is just one of the facets of a more
155: general Sudakov phenomenon \cite{Col89} and it can be traced back to the
156: existence of massless particles of spin one in the
157: spectrum -- the gauge fields. The logarithmic scaling of anomalous dimensions is
158: a universal feature of all gauge theories ranging from QCD to the maximally
159: supersymmetric $\mathcal{N} = 4$ Yang-Mills (SYM) theory. In particular, in the
160: simplest case of twist-two Wilson operators with large Lorentz spin $N \gg 1$,
161: the anomalous dimension behaves as (in the adjoint representation of the
162: $SU(N_c)$ group)~\cite{Kor88}
163: \be\label{gamma=cusp}
164: \gamma (\lambda) = 2 \Gamma_{\rm cusp}(\lambda) \ln N + \mathcal{O} (N^0) \, ,
165: \ee
166: where $\lambda=g_{\rm \scriptscriptstyle YM}^2 N_c$ is the 't Hooft coupling
167: constant and $\Gamma_{\rm cusp}(\lambda)$ is the so-called cusp anomalous
168: dimension~\cite{Pol80}. $\Gamma_{\rm cusp}(\lambda)$ is not universal however and
169: depends on the theory under consideration. It has numerous applications in
170: phenomenology of strong interactions and its calculation both at weak and strong
171: coupling regimes is one of the long-standing problems in gauge theories. At
172: present, the cusp anomaly is known in perturbation theory to the lowest three
173: orders~\cite{MocVerVog04,BerDixSmi05} and there exists a prediction at strong
174: coupling in the $\mathcal{N}=4$ SYM theory based on the gauge/string
175: correspondence~\cite{GubKlePol03,Kru02,BelGorKor03}.
176:
177: The gauge/string correspondence \cite{Mal97} offers a powerful tool to study the
178: dynamics of four-dimen\-sional Yang-Mills theories at strong coupling. It
179: establishes a correspondence between Wilson operators in $\mathcal{N}=4$ SYM
180: theory and certain string excitations on the AdS${}_5\times$S${}^5$ background
181: \cite{BerMalNas02,GubKlePol03}. For operators carrying large quantum numbers (Lorentz spin,
182: isotopic $R-$charge,...) their scaling dimension at strong coupling can be found
183: as an energy of dual (semi)classical string configurations propagating on the
184: curved space. As was shown in Ref.~\cite{GubKlePol03}, the operators of twist two
185: with large Lorentz spin $N$ are dual to a folded string rotating with the angular
186: momentum $N$ on the AdS${}_3$ part of the target space. The resulting expression
187: for the twist-two anomalous dimension takes the form \re{gamma=cusp} with the
188: cusp anomalous dimension at strong coupling given by
189: \be\label{cusp}
190: \Gamma_{\rm cusp}(\lambda) \stackrel{\lambda
191: \gg 1}{=} \frac{\sqrt{\lambda} }{2\pi} + \mathcal{O}((\sqrt{\lambda})^0)\,.
192: \ee
193: Later, the dual string picture was generalized to Wilson operators of higher
194: twist in the $\mathcal{N}=4$ SYM theory carrying both large Lorentz spin and the
195: $R-$charge~\cite{FroTse03}. For operators built from holomorphic scalar fields
196: carrying a unit isotopic charge, the total $R-$charge equals the twist, $L$. For
197: such operators, in the dual picture the center-of-mass of the string rotates with
198: the angular momentum $L$ along a large circle of S${}^5$.
199:
200: The important difference between the operators of twist two, $L=2$, and of higher
201: twist, $L\ge 3$, is that the latter are not uniquely specified by the total
202: Lorentz spin $N$. More precisely, for $L\ge 3$ there exist several
203: (superconformal) operators with the same $N$. These operators mix under
204: renormalization and the size of the mixing matrix rapidly increases with $L$ and
205: $N$. As a consequence, the anomalous dimensions of Wilson operators of high twist
206: $L\ge 3$ also depend on the integers $\ell = 1,2, \ldots$ which enumerate
207: eigenvalues of the mixing matrix (=anomalous dimensions) for a given Lorentz spin
208: $N$. For fixed $L$ and large $N$, possible values of the anomalous dimension
209: occupy a band~\cite{BelGorKor03}. The lower boundary of the band scales for
210: $N\to\infty$ as in \re{gamma=cusp} while the upper boundary scales as $\sim
211: L\,\Gamma_{\rm cusp}(\lambda) \ln N$. This implies that for operators of twist
212: $L=3,4,\ldots$ the minimal anomalous dimension has the same leading asymptotic
213: behavior for $N\to\infty$ as the twist-two anomalous dimension, Eq.~\re{gamma=cusp}.
214: This result is rather general and it holds true in a generic Yang-Mills theory
215: including QCD and $\mathcal{N}=4$ SYM theory.
216:
217: In the present paper, we study the properties of the minimal anomalous dimensions
218: on both sides of the gauge/string correspondence in the limit of large twist
219: $L$ and Lorentz spin $N$. On the string side, the corresponding single-trace
220: Wilson operators are dual to a folded string spinning with large angular momentum
221: $N$ in the AdS${}_3$ part of the anti-de Sitter space and boosted along a large
222: circle on the sphere with a large angular momentum $L$ \cite{FroTse03}. The
223: energy $E$ of this classical string configuration defines the leading asymptotics
224: of the anomalous dimension $\gamma(\lambda) = E - L - N$ of the dual Wilson
225: operator in the $\mathcal{N}=4$ SYM theory in the strong coupling regime in
226: planar approximation. The string theory provides a definite prediction for
227: $\gamma(\lambda)$ as a function of $L$ and $N$. One finds that $\gamma(\lambda)$
228: takes different forms in three regimes~\cite{GubKlePol03,FroTse03}:
229: \begin{itemize}
230: \item For $N \ll L$, in the ``short'' string limit
231: \be
232: \label{1st} \gamma (\lambda) = \lambda \frac{m^2}{2} \frac{N}{L^2} + \ldots
233: \ee
234: \item For $N \gg L$, in the ``long'' string limit
235: \be\label{3rd}
236: \gamma (\lambda) =\left \{
237: \begin{array}{l}
238: {\displaystyle \frac{\lambda}{2\pi^2} \frac{m^2}{L} \ln^2({N}/{L}) + \ldots \,
239: ,\qquad
240: \mbox{for $\xi_{\rm str} < 1$}\,,} \\[3mm]
241: {\displaystyle \frac{\sqrt{\lambda}}{\pi}m\ln ({N}/\sqrt{\lambda}) +
242: \ldots \, ,\qquad \mbox{for $\xi_{\rm str} \gg 1$}\, ,}
243: \end{array}
244: \right.
245: \ee
246: \end{itemize}
247: with the parameter $\xi_{\rm str}$ defined as $\xi_{\rm str} =
248: \lambda\ln^2(N/L)/L^2$. Here the integer $m$ counts the number of times the
249: string is folded onto itself. The minimal anomalous dimension corresponds to a
250: single-folded string, $m=1$. In that case, for $\xi_{\rm str} \gg 1$, the leading
251: asymptotic behavior of $\gamma(\lambda)$ does not depend on the twist $L$ and is
252: the same as for $L = 2$ operators, Eqs.~\re{gamma=cusp} and \re{cusp}. For
253: $\xi_{\rm str} < 1$, the role of the twist $L$ is to create the ``BMN window'',
254: i.e., a region in the parameter space in which the anomalous dimension has an
255: expansion in powers of the BMN coupling $\lambda^\prime \equiv \lambda/(\pi
256: L)^2$~\cite{BerMalNas02}. It is believed that the first few terms in the
257: expansion of $\gamma(\lambda)$ in powers of $\lambda'$ should match similar
258: expressions for the anomalous dimensions of Wilson operators of twist $L$ and
259: spin $N$ obtained in the $\mathcal{N}=4$ SYM theory in the weak coupling
260: regime~\cite{Tse03}.
261:
262: On the gauge theory side, the calculation of anomalous dimensions of higher twist
263: operators with large Lorentz spin turns out to be an extremely nontrivial task in
264: a generic Yang-Mills theory even to one-loop order due to a large size of the
265: mixing matrix. The problem can be overcome thanks to hidden integrability
266: symmetry of the dilatation operator~\cite{BraDerMan98,Bel99,DerKorMan99}, which
267: maps the one-loop mixing matrix for Wilson operators of twist $L$ belonging to
268: the so-called holomorphic $SL(2)$ sector into a Hamiltonian of the Heisenberg
269: magnet of length $L$ and spin $s$ determined by the conformal spin of the quantum
270: fields (for a review see Ref.~\cite{Bel04}). This observation allows one to
271: calculate the exact eigenspectrum of anomalous dimensions of Wilson operators of
272: arbitrary twist and Lorentz spin in integrable sectors of Yang-Mills theory by
273: means of the Quantum Inverse Scattering Method~\cite{TakFad79}. In the
274: $\mathcal{N}=4$ SYM theory, the minimal anomalous dimension of Wilson operators
275: built from $L$ scalar fields and carrying the Lorentz spin $N$ can be identified
276: to one-loop accuracy as the minimal energy in the eigenspectrum of the $SL(2)$
277: Heisenberg magnet of length $L$ and the total spin $N+Ls$ with $s=\ft12$
278: \cite{BeiFroStaTse03}. The gauge/string correspondence suggests that the minimal
279: anomalous dimension defined in this way should match the relations \re{1st} and
280: \re{3rd} in the thermodynamic limit $L\to \infty$.
281:
282: It follows from \re{1st} and \re{3rd} that the anomalous dimensions of higher
283: twist operator depend on a ``hidden'' parameter $\xi_{\rm str} = \lambda
284: \ln^2(N/L)/L^2$ and their behavior at strong coupling is different for $\xi_{\rm
285: str}< 1$ and $\xi_{\rm str} \gg 1$. For $\xi_{\rm str}\gg 1$ the anomalous
286: dimension does not have a perturbative expansion in the BMN coupling
287: $\lambda^\prime$ and scales as $\sim \ln N$. On the gauge theory side, previous
288: studies~\cite{BeiFroStaTse03} of the Bethe ansatz for the $SL(2)$ spin chain in
289: the thermodynamic limit $L \gg 1$ led to the expression for $\gamma(\lambda)$
290: which coincides with \re{1st} in the limit of short strings and with the first
291: relation in \re{3rd} in the limit of long strings. They did not reveal however
292: neither any trace of the second, logarithmic regime in \re{3rd}, nor appearance
293: of a new parameter similar to $\xi_{\rm str}$. This fact is in contradiction with
294: our expectations that the minimal anomalous dimension of higher twist operators
295: should scale logarithmically to all loops as $N\to\infty$, Eq.~\re{gamma=cusp}.
296: The goal of the present study is to unravel the logarithmic scaling of the
297: anomalous dimension both in the gauge and string theory and to understand the
298: physical meaning of the parameter $\xi_{\rm str}$ and its counter-part $\xi$ on
299: the gauge theory side.
300:
301: We shall revisit the calculation of the energy of the $SL(2)$ Heisenberg magnet
302: of spin $s=\ft12,1,\ft32$ using the method of the Baxter $Q-$operator
303: \cite{Bax72} as a main tool and demonstrate that the one-loop anomalous dimension
304: in integrable sectors of Yang-Mills theory has the following scaling behavior in
305: the thermodynamic limit $L\to\infty$:
306: \begin{itemize}
307: \item For $N \ll L$
308: \be
309: \label{1st-g} \gamma (\lambda) = \lambda \frac{m^2}{4s} \frac{N}{L^2} + \ldots \,
310: ,
311: \ee
312: \item For $N \gg L$
313: \be\label{3rd-g}
314: \gamma (\lambda) =\left \{
315: \begin{array}{l}
316: {\displaystyle \frac{\lambda}{2\pi^2} \frac{m^2}{L} \ln^2({N}/{L}) + \ldots \,
317: ,\qquad
318: \mbox{for $\xi < 1$}\,,} \\[3mm]
319: {\displaystyle \frac{{\lambda}}{2\pi^2}\, m\ln N +
320: \ldots \, ,\qqqquad~ \mbox{for $\xi \gg 1$}\, ,}
321: \end{array}
322: \right.
323: \ee
324: \end{itemize}
325: depending on the parameter $\xi = \ln(N/L)/L$. The minimal anomalous dimension
326: corresponds to $m=1$. Here $s$ equals the conformal spin of the field entering
327: the Wilson operator, i.e., $s=\ft12, 1, \ft32$ for scalar, gaugino and gluon
328: fields, respectively. For scalar operators, Eq.~\re{1st-g} and the first relation
329: in \re{3rd-g} coincide with similar expressions in \re{1st} and \re{3rd},
330: respectively. Notice that the anomalous dimension in \re{3rd-g} does not depend
331: on the spin $s$ for $N\gg L$ which suggests that the two regimes in \re{3rd-g}
332: are universal in all gauge theories. This is indeed the case for $\xi \gg 1$ since
333: the coefficient in front of $2\ln N$ coincides with the cusp anomalous dimension
334: at weak coupling, Eq.~\re{gamma=cusp}.
335:
336: For $N\gg L$ and $\xi < 1$, the one-loop anomalous dimension exhibits a novel
337: double logarithmic behavior \re{3rd-g}. It was first discovered from the string
338: theory considerations~\cite{FroTse03} and was later reproduced on the gauge
339: theory side~\cite{BeiFroStaTse03}. A natural question arises whether similar
340: contributions arise at higher loops and whether they can be resummed to all
341: loops. The gauge/string correspondence suggests that in the $\mathcal{N}=4$ SYM
342: theory the anomalous dimension in the region $N \gg L \gg 1$ admits a BMN-like
343: expansion (with $\lambda'=\lambda/(\pi L)^2$ and $\xi = \ln(N/L)/L$)
344: \be\label{gen-exp}
345: \gamma (\lambda) = L \sum_{n=1}^\infty\lr{ \lambda^\prime \ln^2\frac{N}{L} }^n
346: c_n(\xi)+ \ldots\,,
347: \ee
348: where the coefficient functions $c_n(\xi)$ do not depend on the coupling constant
349: and have the following asymptotics for $\xi\to 0$ and $\xi\to\infty$
350: \be\label{coeff-fun}
351: c_n(\xi)=c_{0,n}+c_{1,n} \xi + \mathcal{O}(\xi^2) \,,\qquad c_n(\xi) =\mathcal{O}
352: (1/\xi^{2n-1} )
353: \ee
354: with $c_{0,n}=(-1)^n(-\ft12)_n/n!$. To one-loop order, Eq.~\re{gen-exp} matches
355: (for $m=1$) both relations in \re{3rd-g}. In addition, for $\xi\to \infty$ the
356: coefficient in front of $\lambda^n$ in the right-hand side of \re{gen-exp} scales
357: as $\sim \ln N$ and determines the $n-$loop correction to the cusp anomalous
358: dimension in the weak coupling regime, Eq.~\re{gamma=cusp}. In the strong
359: coupling regime, upon the substitution $\xi=\xi_{\rm str}^{1/2}/\sqrt{\lambda}$,
360: the perturbative series in \re{gen-exp} can be resummed to all loops into the
361: following expression (for $\lambda\to\infty$ and $\xi_{\rm str}=\lambda\ln^2(N/L)/L^2
362: =\rm fixed$)
363: \be
364: \gamma (\lambda) = L \sum_{n=1}^\infty \lr{ \lambda^\prime \ln^2\frac{N}{L} }^n
365: c_n(0)+ \ldots = L \left[ \sqrt{1 + \lambda^\prime \ln^2\frac{N}{L}} - 1 \right] +
366: \ldots\,,
367: \ee
368: where the ellipsis stands for subleading corrections. For $N\to\infty$ this
369: relation also reproduces the leading asymptotic behavior of the anomalous
370: dimension in the last regime in \re{3rd} and, as a consequence, it leads to the
371: expression for the cusp anomalous dimension at strong coupling, Eq.~\re{cusp}.
372:
373: As was already mentioned, the one-loop anomalous dimension \re{gen-exp} coincides
374: with the energy of the $SL(2)$ spin chain of length $L$ in the thermodynamic
375: limit $L\to\infty$. The latter can be found within the Bethe Ansatz approach by
376: systematically expanding the energy in powers of $1/L$ with a help of known
377: semiclassical methods \cite{Sut95}. To leading order of the semiclassical
378: expansion, the Bethe roots condense on two symmetric intervals on the real axis
379: $[-a,-b]\cup [b,a]$, with the boundaries $a$ and $b$ being functions of $L/N$.
380: This leads~\cite{BeiFroStaTse03} to the expression for the energy given in \re{1st}
381: for $N \ll L$ and in the first relation in \re{3rd} for $N\gg L$. It is believed
382: that subleading corrections to the energy are suppressed by powers of $1/L$ and,
383: therefore, are small. We demonstrate that this assumption is only justified for $\xi =
384: \ln(N/L)/L < 1$, while for $\xi\gg 1$ the semiclassical expansion of the energy
385: becomes divergent indicating the change of asymptotic behavior of the anomalous
386: dimension, Eq.~\re{3rd}. The reason why the semiclassical expansion fails is that
387: the two cuts $[-a,-b]$ and $[b,a]$ collide at the origin, that is $b\to 0$ for
388: $\xi\to\infty$, and the Bethe roots have a nonvanishing distribution at the
389: origin. As a consequence, for $N\gg L$ the semiclassical corrections to the
390: anomalous dimension run in powers of $\xi$. To one-loop order they are described
391: in \re{gen-exp} by the function $c_1(\xi)$. We argue that the semiclassical
392: series for $c_1(\xi)$ is divergent for $\xi>1$ and propose an approach which
393: circumvents this difficulty and allows one to determine this function for
394: arbitrary $\xi$. The resulting expression for the one-loop anomalous dimension is
395: valid in the thermodynamic limit throughout the entire interval of $N$ and
396: reproduces correct logarithmic behavior for $N \gg L$, Eq.~\re{3rd-g}.
397:
398: On the string theory side, in the dual picture of the folded string spinning in
399: the AdS${}_3\times$S${}^1$ part of the target space, the logarithmic behavior of
400: the anomalous dimension at strong coupling is associated with the classical
401: string configuration which has two spikes approaching the boundary of the AdS
402: space. Thanks to classical integrability of the string equations of motion
403: \cite{ManSurWad02}, the same configuration is described by the spectral
404: (elliptic) symmetric curve endowed with a meromorphic differential of
405: quasimomentum possessing a double pole at $x=\pm \sqrt{\lambda'}$ and having a
406: prescribed asymptotic behavior at the origin and infinity~\cite{ZakMik78,Kri94,KazZar04}.
407: The branching points of the curve, $\pm b_{\rm str}$ and $\pm a_{\rm str}$, depend
408: on the ratio $L/N$ and the coupling constant $\lambda'$. We show that for $N \gg L$
409: and $\xi_{\rm str} < 1$ the branching points admit a regular expansion in powers
410: of $\lambda'$ and, as a consequence, the anomalous dimension exhibits the BMN scaling,
411: Eq.~\re{3rd}. For $\xi_{\rm str} \gg 1$, $b_{\rm str}$ approaches its minimal value
412: $\sqrt{\lambda'}$ so that the inner boundaries of two cuts $[-a_{\rm str},-b_{\rm
413: str}]$ and $[b_{\rm str},a_{\rm str}]$ coincide with the position of poles of the
414: momentum differential and cannot collide. This nonanalyticity manifests itself
415: through the appearance of $\sqrt{\lambda}$ prefactor in the logarithmic behavior
416: of the anomalous dimension at strong coupling, Eq.~\re{3rd}.
417:
418: Our consequent presentation is organized as follows. In Sect.~\ref{SectAnDim}, we
419: outline a general framework for analysis of one-loop anomalous dimensions in the
420: thermodynamic limit. It is based on the semiclassical expansion of solutions to
421: the Baxter equation~\cite{PasGau92,Kor95,Smi98}. In Sect.~\ref{ThermodynamicSection},
422: we apply the semiclassical approach to determine the minimal anomalous dimension
423: in the thermodynamic limit and demonstrate that the semiclassical expansion breaks
424: down for $\ln(N/L) \gg L $ due to collision of cuts. Then, we present an approach
425: to go consistently beyond the semiclassical expansion and use it to describe the
426: minimal anomalous dimension for large Lorentz spin. In Sect.~\ref{ConclusionSection},
427: we analyze the asymptotic behavior of the anomalous dimensions at strong coupling
428: based on the string sigma model consideration. Section~\ref{StringSigmaModel}
429: contains concluding remarks. Some technical details of our calculations are
430: summarized in the Appendix.
431:
432: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
433: \section{Anomalous dimensions in gauge theory}
434: \label{SectAnDim}
435: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
436:
437: Let us start with the calculation of one-loop anomalous dimensions of
438: (super)conformal Wilson operators of arbitrary twist $L$ and Lorentz spin $N$
439: belonging to integrable sectors of (supersymmetric) Yang-Mills theories. For
440: quantum fields transforming in the adjoint representation of the gauge group, the
441: operators under considerations have the following generic form
442: \be\label{O-def}
443: \mathcal{O}_{N,L}(0) = \sum_{k_1+\ldots+k_L =N} c_{k_1\ldots k_L} \tr
444: \left\{D_+^{k_1} X(0)D_+^{k_2} X(0)\ldots D_+^{k_L} X(0)\right\},
445: \ee
446: where $X(0)$ stands for the so-called ``good'' component of quantum fields of a
447: definite helicity in the underlying gauge theory, $D_+=D_\mu n^\mu$ is the
448: covariant derivative projected onto the light-cone, $n_\mu^2=0$. The expansion
449: coefficients $c_{k_1\ldots k_L}$ are fixed from the condition for
450: $\mathcal{O}_{N,L}(0)$ to have an autonomous scale dependence, i.e., Eq.\
451: \re{O-def} has to be an eigenstate of the one-loop dilatation operator.
452: Integrability allows one to map the one-loop anomalous dimension of the operators
453: \re{O-def} into energy $\varepsilon$ of the noncompact $SL(2)$ Heisenberg spin
454: chain of length $L$ and the total spin $N + Ls$
455: \cite{BraDerMan98,Bel99,DerKorMan99,Bei04},
456: \be\label{gamma=energy}
457: \gamma(\lambda) = \frac{\lambda}{8\pi^2}\, \varepsilon +
458: \mathcal{O}(\lambda^2)\,.
459: \ee
460: Here the (half-)integer $s$ is given by the conformal spin \cite{Bel04} of the
461: quantum field $X(0)$, that is, $s=1/2$ for scalars, $s=1$ for gaugino fields of
462: helicity $\pm 1/2$ and $s=3/2$ for gauge fields of helicity $\pm 1$.
463:
464: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
465: \subsection{Exact solution}
466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
467:
468: Let us first describe the exact solution for the energy $\varepsilon$ of the
469: $SL(2)$ magnet of length $L$ and single-particle spin $s$ in each site. We shall
470: employ the method of the Baxter $Q-$operator \cite{Bax72} which proves to be
471: convenient for analyzing various semiclassical limits of $\varepsilon
472: =\varepsilon(N,L)$ including the limit of the large spin $N$ and length
473: $L$~\cite{Kor95,Smi98}. The method relies on the existence of an operator $Q(u)$
474: which acts on the Hilbert space of the $SL(2)$ spin chain and is diagonalized by
475: all eigenstates of the magnet for arbitrary complex parameter $u$. Discussing the
476: energy spectrum it suffices to study the eigenvalues of the $Q-$operator that we
477: shall denote by $Q(u)$. The same function $Q(u)$ determines the wave function of
478: the magnet in the representation of Separated Variables \cite{Skl90}
479: \footnote{For an interpretation of the $Q-$operator in string theory see
480: Ref.~\cite{Gor03}.}. This allows one to analyze $Q(u)$ in the semiclassical limit
481: with the help of the WKB machinery well-known from quantum mechanics and, then,
482: determine $\varepsilon(N,L)$.
483:
484: By construction, $Q(u)$ satisfies the second-order finite-difference equation
485: \be\label{Baxter-eq}
486: (u + is)^L Q (u + i) + (u - is)^L Q(u - i) = t_L(u) Q(u) \, ,
487: \ee
488: which can be thought of as a Schr\"odinger equation for a single-particle wave
489: function in the representation of Separated Variables~\cite{Skl90}. Here $t_L(u)$
490: is a polynomial in $u$ of degree $L$ with coefficients given by conserved charges
491: \be\label{t_L}
492: t_L(u) = 2 u^L + q_2 u^{L-2} + \ldots + q_L
493: \ee
494: The lowest integral of motion $q_2$ is related to the total spin of the $SL(2)$
495: chain, $N+Ls$,
496: \be\label{q2}
497: q_2 = -(N + Ls) (N + Ls - 1) + L s (s-1)
498: \ee
499: with $N=0,1,\ldots$.
500:
501: In what follows we shall refer to Eq.\ \re{Baxter-eq} as the Baxter equation.
502: Taken alone, it does not fix the function $Q(u)$ and it has to be supplemented
503: by an additional condition that $Q(u)$ has to be polynomial in $u$ \cite{FadKor94}.
504: Examining the asymptotic behavior of both sides of \re{Baxter-eq} for $u\to\infty$,
505: it is easy to see that the degree of $Q(u)$ is fixed by the total spin $N$ and,
506: therefore, up to an overall normalization, one can write
507: \be\label{Q-polynom}
508: Q(u) = \prod_{k=1}^N (u-\lambda_k)\,.
509: \ee
510: One substitutes this ansatz into \re{Baxter-eq}, takes $u=\lambda_k$ and finds
511: that the roots $\lambda_1,\ldots,\lambda_N$ satisfy the Bethe equations
512: \be\label{Bethe-roots}
513: \lr{\frac{\lambda_k+is}{\lambda_k-is}}^L=\prod_{j=1,j\neq k}^N
514: \frac{\lambda_k-\lambda_j-i}{\lambda_k-\lambda_j+i}\,.
515: \ee
516: Solving the Baxter equation \re{Baxter-eq} supplemented by \re{Q-polynom} one
517: obtains quantized values of the charges $q_3,\ldots,q_L$ and evaluates the
518: corresponding energy and quasimomentum as \cite{FadKor94}
519: \be\label{Energy-Baxter}
520: \varepsilon = i\lr{\ln Q(is)}'-i\lr{\ln Q(-is)}'\,,\qquad \e^{i\theta} =
521: \frac{Q(is)}{Q(-is)}\,.
522: \ee
523: Replacing $Q(u)$ by its expression \re{Q-polynom} one verifies that these relations
524: coincide with those coming from the Algebraic Bethe Ansatz~\cite{TakFad79}
525: \be\label{Energy-ABA}
526: \varepsilon = \sum_{k=1}^N \frac{2s}{\lambda_k^2+s^2}
527: \, ,
528: \qquad \e^{i\theta} = \prod_{k=1}^N \frac{\lambda_k-is}{\lambda_k+is}
529: \, .
530: \ee
531: The cyclic symmetry of the single-trace operators \re{O-def} imposes an
532: additional selection rule for the eigenstates of the spin magnet,
533: $\e^{i\theta}=1$. Equations \re{Energy-Baxter} and \re{Energy-ABA} allow one to
534: calculate the energy of the spin chain and, then, obtain the one-loop anomalous
535: dimension of Wilson operators \re{O-def} with a help of \re{gamma=energy}.
536:
537: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
538: \subsection{Quasiclassical approach}
539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540:
541: Let us examine the Baxter equation \re{Baxter-eq} for $N+Ls \gg 1$. In this
542: limit, the charge $q_2$ takes large negative values and one can apply
543: semiclassical techniques~\cite{Kor95} to construct the solution to
544: \re{Baxter-eq}. To go over to the semiclassical limit, we introduce two scaling
545: parameters
546: \be\label{beta}
547: \eta=(N+Ls)^{-1}\,,\qquad \beta = s L \eta =\frac{s L}{ N+Ls } \,.
548: \ee
549: By definition, $0\le \beta \le 1$ with the boundary values corresponding to
550: \be\label{limits}
551: \beta\stackrel{L\ll N}{\longrightarrow} 0
552: \, , \qqquad
553: \beta\stackrel{L\gg N}{\longrightarrow} 1
554: \, .
555: \ee
556: The parameter $\eta \ll 1$ will play the role of the Planck constant. One
557: rescales the spectral parameter as $u=x/\eta$ and introduces the eikonal phase
558: (the Hamilton-Jacobi ``action'' function) $S(x)$ as
559: \be\label{Q-ansatz}
560: Q(x/\eta) = \eta^{-N}\,\exp\lr{\frac1{\eta} S(x)} \,,\qquad S(x) = \eta
561: \sum_{k=1}^N \ln (x-\eta \lambda_k)\,.
562: \ee
563: The energy and the quasimomentum in Eq.~\re{Energy-Baxter} are then given in
564: terms of the function $S(x)$ by the following expressions
565: \be\label{E-general}
566: \varepsilon = i \left[S'(i\beta/L) - S'(-i\beta/L)\right] \,,\qquad \e^{i\theta}
567: = \exp\left\{\frac1{\eta}[S(i\beta/L) - S(-i\beta/L)]\right\} \,.
568: \ee
569: It also proves convenient to introduce a notation for the ``effective potential''
570: \be\label{tau-expansion}
571: \tau(x) =(\eta/x)^{L}\,t_L(x/\eta) = 2 + \frac{\widehat q_2}{x^2} +
572: \frac{\widehat q_3}{x^3} + \ldots + \frac{\widehat q_L}{x^L}
573: \, ,
574: \ee
575: with $\widehat q_k=q_k \eta^k$.
576:
577: In the semiclassical approach~\cite{PasGau92,Kor95} one assumes that the
578: function $S(x)$ and the integrals of motion $q_k$ (with $k=3,\ldots,L$) admit a
579: systematic expansion in powers of $\eta$
580: \be\label{WKB-ansatz}
581: S(x) = S_0(x) + \eta\, S_1(x) + \ldots\,,\qquad \widehat q_k = \widehat
582: q_k^{\scriptscriptstyle (0)} + \eta \, \widehat q_k^{\scriptscriptstyle (1)} +
583: \ldots\, .
584: \ee
585: It is tacitly assumed that the expansion of $S(x)$ is convergent and each term is
586: uniformly bounded.%
587: \footnote{As we will show in Sect.~\ref{ThermodynamicSection}, this assumption is
588: justified for $\ln(N/L) < L$ and is invalid otherwise.} This leads to the expansion
589: of the effective potential \re{tau-expansion}, $\tau(x)= \tau_0(x) + \eta \tau_1(x)
590: + \ldots$ with
591: \be\label{tau0}
592: \tau_0(x) = 2 - \frac1{x^2}+\frac{\widehat q_3^{\scriptscriptstyle
593: (0)}}{x^3}+\ldots +\frac{\widehat q_L^{\scriptscriptstyle (0)}}{x^L} \,,\qquad
594: \tau_1(x) = \frac{\widehat q_3^{\scriptscriptstyle (1)}}{x^3}+\ldots
595: +\frac{\widehat{q}_L^{\scriptscriptstyle (1)}}{x^L}\,,\qquad \ldots
596: \ee
597: One substitutes \re{WKB-ansatz} into the Baxter equation \re{Baxter-eq} and
598: equates the coefficients in front of powers of $\eta$ to get to leading order
599: \be\label{p-def}
600: 2 \cos p(x) = \tau_0 (x)\,,\qquad p(x) = S_0^\prime (x) + \frac{\beta}x\,.
601: \ee
602: In the finite-gap theory, the function $p(x)$ defines the Bloch-Floquet
603: multiplier in an auxiliary linear problem for the Baker-Akhiezer function and has
604: the meaning of the (quasi)momentum \cite{NovManPitZak84,FadTak87,BabBerTal03}.
605: For the first subleading term in the semiclassical expansion one finds in a
606: similar manner
607: \ba\label{S1-Baxter}
608: S_1^\prime (x) = - \frac{p^\prime (x)}{2} \cot p (x) - \frac{1}{2 \sin p(x)}
609: \left( \tau_1(x) + \frac{\beta (1 - s)}{2 x^2} \tau_0(x) \right) \, .
610: \ea
611: It is straightforward to derive the subleading terms $S_{k\ge 2}'(x)$ but we will
612: not need them for our purposes. The obtained expressions for the action functions
613: $S'_0(x)$ and $S'_1(x)$ depend on yet unknown conserved charges $\widehat
614: q_k^{\scriptscriptstyle (0)}$ and $\widehat q_k^{\scriptscriptstyle (1)}$,
615: respectively. Quantization conditions for these charges follow from the requirement
616: for $Q(x/\eta)$, Eq.~\re{Q-ansatz}, to be a single valued function of $x$.
617:
618: According to \re{tau0}, $\tau_0(x)$ is a polynomial of degree $L$ in $1/x$.
619: Solving \re{p-def}, one finds that the momentum $p(x)$ is, in general, a
620: double-valued function on the complex $x-$plane with the square-root branching
621: points $x_j$ obeying the condition $\tau_0 (x_j)=\pm 2$. It is convenient to
622: introduce the function $y(x)=2 \sin p(x)$ and define a complex curve~\cite{Kor95}
623: \be\label{curve}
624: \Gamma_L: \qquad y^2 = 4-\tau_0^2(x)\,,\qquad \tau_0(x) = 2 - \frac1{x^2} +
625: \frac{\widehat q_3^{\scriptscriptstyle (0)}}{x^3} + \ldots + \frac{\widehat
626: q_L^{\scriptscriptstyle (0)}}{x^L}\,.
627: \ee
628: For arbitrary complex $x$, except the branching points $y(x_j)=0$, the relation
629: \re{curve} defines two values for $y(x)$. Then, $y(x)$ being a double-valued
630: function on the complex $x-$plane, becomes a single-valued function on the
631: hyperelliptic Riemann surface defined by the complex curve $\Gamma_L$. This
632: surface has a genus $L-2$ and is realized by gluing together two copies of the
633: complex $x-$plane along the cuts running between the branching points $x_{2j-1}$
634: and $x_{2j}$.
635:
636: For the $SL(2)$ magnet the Bethe roots verifying Eq.\ \re{Bethe-roots} take real
637: values only, ${\rm Im} \lambda_k =0$. In the semiclassical limit, $\eta \to 0$,
638: they condense on finite intervals on the real axis where the momentum $p(x)$
639: takes purely imaginary values~\cite{Kor95}. In terms of the hyperelliptic curve
640: \re{curve}, this corresponds to $y^2 \le 0$, or
641: \be\label{cuts}
642: \tau_0^2(x) \ge 4\,,\qquad \mbox{for $x\in \mathcal{S}=[x_{2L-2},x_{2L-3}]\cup
643: \ldots \cup [x_4,x_3] \cup [x_2,x_1] $}\,,
644: \ee
645: where $x_1>x_2> \ldots >x_{2L-2}$ and one of the intervals contains the origin.
646: The total number of intervals in \re{cuts} equals $L-1$ and the end points $x_j$
647: are just the branching points of the complex curve \re{curve}, $\tau_0^2(x_j)=4$.
648: As follows from \re{curve}, the curve can be parameterized by the set of $2L-2$
649: real branching points as
650: \be\label{y2-roots}
651: y^2 = \frac4{x^2} \prod_{j=1}^{2L-2} \lr{1-\frac{x_j}{x}}\,.
652: \ee
653: The intervals $\mathcal{S}$ have the meaning of regions where the classical
654: motion of the system occurs in the separated variables\footnote{In the
655: finite-gap theory~\cite{NovManPitZak84}, the same intervals have the meaning of
656: forbidden zones in the auxiliary linear problem for the Baker-Akhiezer function.}.
657: Later on we shall encounter the situation when, say, $j^{\rm th}$ interval shrinks
658: into a point, $x_{2j}=x_{2j-1}$, so that the motion on this interval is frozen at
659: the classical level. In what follows we shall refer to $x_{2j}=x_{2j-1}$ as a
660: double point.
661:
662: The leading term of the semiclassical expansion, $S_0(x)$, is determined by the
663: momentum $p(x)$, Eq.~\re{p-def}. As follows from its definition \re{p-def}
664: \be\label{p-exp}
665: p(x) = 2i \ln \frac{\sqrt{\tau_0(x)+2}-\sqrt{\tau_0(x)-2}}{2}\,.
666: \ee
667: The momentum $p(x)$ takes purely imaginary values on the intervals \re{cuts} and
668: its values at the end point of the $j^{\rm th}$ interval coincide,
669: $p(x_{2j-1})=p(x_{2j})$, with $\e^{i p(x_{2j})}=\pm 1$ for $\tau_0(x_{2j})=\pm
670: 2$, respectively. Continuing $p(x)$ to the complex $x-$plane one finds that
671: $p'(x)$ is an analytical function on the complex plane with cuts running along
672: the intervals \re{cuts}. It defines a meromorphic differential on $\Gamma_L$
673: \be\label{dp}
674: dp = p'(x)\, dx = -\frac{\tau_0'(x)}{\sqrt{4-\tau_0^2(x)}}dx\,.
675: \ee
676: From \re{curve} and \re{p-exp} one finds $\tau_0(x) = 2- 1/x^2 +
677: \mathcal{O}(1/x^3)$ so that $\e^{p(\infty)}=1$ and
678: \be\label{p-infinity}
679: dp \sim\mp \frac{dx}{x^2}\,,\qquad \mbox{for $x\to\infty$}
680: \ee
681: where `$-/+$' correspond to the upper/lower sheet of $\Gamma_L$. According to the
682: definition \re{p-def}, $p(\infty)$ is defined modulo $2\pi$. Choosing the
683: normalization condition $p(\infty) = 0$, one finds that at the end points of the
684: $j^{\rm th}$ interval in \re{cuts}, the momentum takes the values $p(x_{2j})
685: =p(x_{2j-1}) = - \pi j$ for $x_{2j}>0$. As a consequence, the differential $dp$
686: satisfies the normalization conditions~\cite{SmiRes83,NovManPitZak84}
687: \be\label{p-periods}
688: 2 \int^{x_{2j-1}}_{x_{2j}} dx\, p'(x) = - \oint_{\alpha_j} dp = 0\,,\qquad 2
689: \int_{x_{2j-1}}^{\infty} dx\, p'(x) = - \int_{\gamma_j} dp = -2\pi j\,.
690: \ee
691: Here in both equations, the integration in the left-hand side goes over the upper
692: sheet of $\Gamma_L$. In the right-hand side of the first relation, the
693: differential $dp$ is integrated over the $\alpha_j-$cycle encircling the interval
694: $[x_{2j},x_{2j-1}]$ in the anticlockwise direction. The contour $\gamma_j$ in the
695: second relation starts on the upper sheet above $x = \infty$ crosses the same
696: interval and then goes to infinity on the lower sheet (see Fig.\
697: \ref{CyclesFig}).
698:
699: The obtained expressions for $S_0'(x)$ and $S_1'(x)$, Eqs.~\re{p-def} and
700: \re{S1-Baxter}, respectively, depend on the conserved charges $\widehat q_k$, yet
701: to be determined. To work out the quantization conditions for $\widehat q_k$ one
702: examines the first derivative of the eikonal phase \re{Q-ansatz}
703: \be\label{S-prime}
704: S'(x) = \eta \sum_{k=1}^N \frac1{x-\eta \lambda_k}\,.
705: \ee
706: The discontinuity of $S'(x)$ across the cuts \re{cuts} gives the distribution
707: density of rescaled Bethe roots $\eta \lambda_k$. Assuming that the roots $\eta
708: \lambda_k$ take finite values for $\eta\to 0$, one finds the behavior of $S'(x)$
709: at infinity on the upper, physical sheet of $\Gamma_L$ as
710: \be\label{S-res-inf}
711: S'(x) \sim \frac{\eta N}{x}=\frac{1-\beta}{x}\,,\qquad {\rm for}\ x \to \infty\,,
712: \ee
713: with $\eta$ and $\beta$ defined in \re{beta}. Replacing $S(x)$ by its
714: semiclassical expansion \re{WKB-ansatz} and matching the
715: coefficients in front of powers of $\eta$ one obtains%
716: \footnote{Later on we shall consider solutions to the Baxter equation satisfying
717: $Q(u)=Q(-u)$, or equivalently $S(x)=S(-x)$. For such solutions, $S_k'(x)$ are odd
718: functions of $x$ and their asymptotics at infinity involves odd powers of $x$
719: only, that is, $S_{k \ge 1}'(x) \sim 1/{x^3}$.}
720: \be\label{as-infinity}
721: S_0'(x) \sim \frac{1-\beta}{x}\,,\qquad S_{1}'(x) \sim \frac1{x^2} \,,\qquad
722: \ldots
723: \ee
724: %
725: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
726: % Figure
727: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
728: \begin{figure}[t]
729: \begin{center}
730: \mbox{
731: \begin{picture}(0,135)(230,0)
732: \put(0,0){\insertfig{16}{cycles}}
733: \end{picture}
734: }
735: \end{center}
736: \caption{ \label{CyclesFig} The definition of the $\alpha-$cycles and
737: $\gamma-$contours on the Riemann surface $\Gamma_L$. The dashed lines represent
738: the part of the path on the lower sheet of the surface. }
739: \end{figure}%
740: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
741: %
742: According to \re{Q-polynom}, the total number of Bethe roots equals $N$. For a
743: given energy level they are distributed on $L-1$ intervals \re{cuts}. Denoting by
744: $\ell_j$ the number of Bethe roots on the $j^{\rm th}$ interval, one has
745: \be\label{a-period}
746: \frac1{2\pi i}\oint_{\alpha_j} dx\, S'(x) = \eta \ell_j=
747: \frac{\ell_j}{N+Ls}\,,\qquad (j=1,\ldots,L-1)
748: \, .
749: \ee
750: The sum of all $\alpha-$cycles is homologous to zero and, as a consequence, the sum
751: of all $\alpha-$periods is given by the residue of $S'(x)$ at infinity. Together
752: with \re{S-res-inf} this leads to $N=\ell_1 + \ldots+\ell_{L-1}$.
753:
754: Solving the quantization conditions \re{a-period} one can determine quantized
755: values of the conserved charges $\widehat q_k$. According to \re{a-period} they
756: depend both on the scaling parameter $\eta$ and the set of nonnegative integers
757: $\ell_1,\ldots,\ell_{L-1}$. Replacing $\widehat q_k$ in \re{tau0}, \re{p-def} and
758: \re{S1-Baxter} by their quantized values, one constructs semiclassical expression
759: for $S(x)$ and, then, applies \re{E-general} to determine the energy and
760: quasimomentum,
761: \be
762: \widehat q_k = \widehat q_k(\ell_1,\ldots,\ell_{L-1};\eta) \, , \qquad
763: \varepsilon = \varepsilon (\ell_1,\ldots,\ell_{L-1};\eta)\,.
764: \ee
765: Explicit form of these relations for $L=3$ can be found in Refs.\
766: \cite{Kor95,BraDerMan98,Bel99,DerKorMan99}. In particular, the quantized values
767: of the energy and conserved charges exhibit remarkable regularity and form
768: trajectories. The flow parameter along each trajectory is given by the total
769: spin $\eta=(N+Ls)^{-1}$ while the integers $\ell_1,\ldots,\ell_{L-1}$ enumerate
770: the trajectories and encode a nontrivial analytic structure in the eigenspectrum.
771: For given $N$, the total number of trajectories equals the number of partitions
772: of $N$ into the sum of $L-1$ nonnegative integers $\ell_1,\ldots,\ell_{L-1}$.
773: This is in a perfect agreement with the number of irreducible components entering
774: the tensor product of $L$ copies of the $SL(2)$ modules~\cite{BelGorKor03}.
775:
776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
777: \subsection{Minimal energy trajectory}
778: \label{CollisionCuts}
779: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
780:
781: In the semiclassical approach described in the previous section, the energy of
782: the spin magnet, or equivalently the one-loop anomalous dimension of Wilson
783: operators, is parameterized by the complex curve $\Gamma_L$, Eq.~\re{curve}. The
784: genus of the curve $g=L-2$ is defined by the length of the spin chain whereas its
785: moduli are determined by the quantized values of the conserved charges $\widehat
786: q_k$ which depend in their turn on the set of integers $\ell_1,\ldots,\ell_{L-1}$.
787: Going over to different parts of the spectrum amounts to specifying the integers
788: $\ell_1,\ldots,\ell_{L-1}$.
789:
790: In this paper we are interested in the eigenstates possessing the minimal
791: possible energy for a given total spin $N$. Such eigenstates belong to a
792: particular trajectory to which we shall refer as the minimal energy trajectory
793: (see Fig.\ \ref{Fig-roots-energy10} below). To describe it one has first to
794: identify the corresponding integers $\ell_1,\ldots,\ell_{L-1}$. We remind that
795: the energy \re{Energy-ABA} is determined by zeros of the function $Q(u)$, or
796: equivalently the Bethe roots. In the thermodynamical limit they condense on the
797: intervals \re{cuts}. For a given total spin $N$, the minimal energy is realized
798: when the Bethe roots are located on two symmetric cuts most distant from the
799: origin \cite{Sut95}, that is, $\ell_1 = \ell_{L-1}=N/2$ and $\ell_j=0$ for
800: $j=2,\ldots,L-2$. The corresponding $Q(u)$ should be an even function,
801: $Q(u)=Q(-u)$. According to \re{Energy-Baxter}, \re{Q-polynom} and \re{Baxter-eq},
802: such states automatically satisfy the cyclic symmetry condition $\e^{i\theta}=1$
803: and possess the following quantum numbers
804: \be\label{q-odd}
805: N = {\rm even}\,,\qquad q_{2k+1} = 0
806: \, ,
807: \ee
808: with $0 \le k \le (L-1)/2$. The fact that $\ell_j=0$ implies that $j^{\rm th}$
809: interval does not contain Bethe roots and, therefore, it shrinks into a point,
810: $x_{2j-1}=x_{2j}$. From the point of view of separated variables, this means
811: that classically all but two collective degrees of freedom are frozen and the
812: classical motion is confined to the two intervals with $\ell_1 = \ell_{L-1} =N/2$.
813: For the complex curve \re{y2-roots}, this implies that all branching points
814: except four, $x_1=-x_{2L-2}$ and $x_2=-x_{2L-3}$, become the double points,
815: $x_{2j-1}=x_{2j}$, and the curve $\Gamma_L$ reduces to the elliptic curve (see
816: Eq.~\re{curve-reduced} below).
817:
818: Let us apply the semiclassical approach to obtain the expression for the energy
819: for $L\gg 1$ along the minimal energy trajectory as a function of $N$. To begin
820: with, we consider the ground state of the $SL(2)$ spin chain. It has the total
821: spin $N=0$ and is described by a trivial solution to the Baxter equation
822: \re{Baxter-eq}, $Q(u)=1$, or equivalently $S(x)=0$. From \re{Energy-Baxter}, the
823: corresponding energy is $\varepsilon = 0$ and the integrals of motion can be read
824: off from \re{Baxter-eq} upon substituting $Q(u)=1$. To leading order in $\eta=
825: 1/(sL)$ the transfer matrix and momentum, Eq.~\re{p-def}, look like
826: \be
827: \label{N=0-roots}
828: \tau_0(x)=2 \cos(1/x)\,,\qquad p(x) = 1/x \,.
829: \ee
830: Matching these expressions into \re{curve} and \re{y2-roots} one finds that for
831: $L\to \infty$ all branching points of the spectral curve \re{y2-roots} are double
832: points, $x_{2j-1} = x_{2j} = 1/(\pi j)$.
833:
834: Let us now consider the minimal energy eigenstates with $N \gg 1$ and $L \gg 1$
835: and distinguish two limiting cases: (i) $N/L={\rm fixed}$ and (ii) $N/L \gg 1$,
836: corresponding to $0 < \beta < 1$ and $\beta\to 0$, respectively, Eq.~\re{beta}.
837: We recall that the function $Q(u)$ has exactly $N$ roots, Eq.~\re{Q-polynom}.
838: Assuming that the Bethe roots scale as $\lambda_k \sim 1/\eta$, one finds from
839: \re{Energy-ABA} that for $N/L={\rm fixed}$ the energy should behave as
840: $\varepsilon \sim N/L^2\sim 1/L$. Indeed, one can obtain the same scaling of
841: $\varepsilon$ by naively expanding the energy \re{E-general} in powers of $1/L$
842: \be\label{E-naive-expansion}
843: \varepsilon = -\frac{2\beta}{L} S''(0) + \mathcal{O}(\beta^3/L^3) =
844: -\frac{2\beta}{L} \left[ S_0''(0) + \eta S_1''(0)+ \mathcal{O}(\eta^2)\right]
845: \, .
846: \ee
847: Here in the second relation we replaced $S(x)$ by its semiclassical expansion
848: \re{WKB-ansatz}. In a similar manner, one obtains for the quasimomentum
849: \re{E-general}
850: \be
851: \theta = 2s \left[S_0'(0) + \eta S_1'(0) + \mathcal{O}(\eta^2)\right]=0\,,
852: \ee
853: where we took into account that $S_k(x)$ are even functions of $x$ for the
854: minimal energy states
855: \be\label{as-zero}
856: S_0'(0) = 0 \,,\qquad S_1'(0) = 0\,, \quad \ldots
857: \, .
858: \ee
859: Then, one finds from \re{p-def} the asymptotic behavior of the momentum around
860: the origin on the upper (``physical'') sheet of the Riemann surface \re{curve} as
861: \be\label{p-zero}
862: p(x) = \frac{\beta}{x} + \mathcal{O}(x)\,,
863: \ee
864: so that the differential $dp$ has a double pole above $x=0$. It is important to
865: stress that the relations \re{E-naive-expansion} and \re{p-zero} were obtained
866: under the assumption that $S'(x)$ is regular at the origin.
867: %, or equivalently the point $x=0$ does not belong to the cuts \re{cuts}.
868: For the minimal energy eigenstates, all but two intervals in \re{cuts} shrink
869: into the double points and the Bethe roots condense on two symmetric cuts on the
870: real axis that we shall denote as $[-a,-b]$ and $[b,a]$. Then, the spectral curve
871: \re{y2-roots} reduces to
872: \be\label{curve-reduced}
873: y^2 = \tilde y^2 \left[ \frac2{x^3}\prod_j \lr{1-\frac{x^2_{2j}}{x^2}} \right]^2
874: \,,\qquad \tilde y^2 = (x^2-a^2)(x^2-b^2) \,,
875: \ee
876: where we took into account that for the minimal energy states the branching
877: points appear in pairs of opposite sign. The parameters $a$, $b$ and $x_{2j}$ are
878: determined by the condition for $S_0'(x)$, Eq.~\re{p-def}, to be an analytical
879: function on the complex $x-$plane with two symmetric cuts on the real axis and
880: prescribed asymptotic behavior at $x=0$ and $x\to\infty$, Eqs.~\re{as-zero} and
881: \re{as-infinity}, respectively \cite{SmiRes83,NovManPitZak84}. Equation
882: \re{E-naive-expansion} implies that the energy has the BMN scaling in the
883: thermodynamic limit, $\varepsilon \sim 1/L$. This relation is in an apparent
884: contradiction with the well-known fact that for $N \gg L$ the anomalous dimension
885: should scale logarithmically $\varepsilon \sim \ln N$ with the prefactor being
886: $L$ independent. To understand the reason for this discrepancy, one notices that
887: the relation \re{E-naive-expansion} is valid provided that $S'(x)$ is analytical
888: in the vicinity of $x=0$, or equivalently, there is no accumulation of the Bethe
889: roots around the origin. For $N,\, L \to \infty$ with $N/L={\rm fixed}$ this is
890: indeed the case but, as we will argue in the next section, the situation
891: drastically changes for $N\gg L$.
892:
893: As a hint, let us consider the eigenstates with $L={\rm fixed}$, $N\to\infty$ and
894: the charges satisfying \re{q-odd}. In this case, the transfer matrix $\tau_0(x)$,
895: Eq.~\re{tau0}, is an even polynomial of degree $L$ which scales for $x\to 0$ as
896: \be\label{tau-x=0}
897: \tau_0(x) \sim \frac{\widehat q_L^{\scriptscriptstyle (0)}}{x^L}
898: \ee
899: provided that $\widehat q_L^{\scriptscriptstyle (0)}\neq 0$. If $\widehat
900: q_L^{\scriptscriptstyle (0)}$ vanishes, the asymptotics of $\tau_0(x)$ is
901: governed by the first nonvanishing charge $\widehat q_{2k}^{\scriptscriptstyle
902: (0)} \neq 0$ with $2k < L$. Substituting \re{tau-x=0} into \re{dp} one finds the
903: leading asymptotic behavior of the momentum for $x\to 0$ on the upper sheet of
904: \re{curve} as
905: \be
906: dp \sim i L \frac{dx}{x}\,,
907: \ee
908: yielding $p(x) \sim i L \ln x$. According to \re{beta}, $\beta\sim sL/N \to 0$
909: for $N\to\infty$ and one finds from \re{p-def} the asymptotics of $S^\prime_0(x)$
910: above $x=0$ on the physical sheet of \re{curve} as
911: \be\label{S0-log}
912: S_0'(x) \sim i L \ln x\,.
913: \ee
914: Comparing this relation with \re{as-zero}, one concludes that $S_0'(x)$ is no
915: longer analytical at the origin due to accumulation of rescaled Bethe roots at
916: $x=0$. Finally, one substitutes \re{S0-log} into \re{E-general} and obtains the
917: energy
918: \be\label{E-ln}
919: \varepsilon \sim 2 L \ln(L/\beta) \sim 2 L \ln N\,.
920: \ee
921: Notice that the coefficient in front of $\ln N$ in the right-hand side of
922: \re{E-ln} is determined by the leading asymptotic behavior of the transfer matrix
923: \re{tau-x=0} for $x\to 0$. For $\widehat q_L^{\rm (0)}\neq 0$ this coefficient
924: takes the maximal possible value. For $\widehat q_L^{\scriptscriptstyle (0)}
925: =\ldots=\widehat q_{2m+2}^{\scriptscriptstyle (0)} =0$, the transfer matrix
926: \re{tau0} scales as $\tau_0(x)\sim \widehat q_{2m}^{\scriptscriptstyle (0)}
927: /x^{2m}$ (recall that the charges with odd indices vanish, Eq.~\re{q-odd})
928: leading to
929: \be
930: \label{E-m} \varepsilon \sim 4m \ln N \,,\qquad \widehat
931: q_{2m+2}^{\scriptscriptstyle (0)}=\ldots=\widehat q_{L}^{\scriptscriptstyle (0)}
932: = 0\,.
933: \ee
934: The values of the remaining charges $\widehat q_{4}^{\scriptscriptstyle
935: (0)},\ldots,\widehat q_{2m}^{\scriptscriptstyle (0)}$ will be determined below
936: (see Eq.~\re{q-expression}). Throughout this paper we are interested in the
937: eigenstates with the minimal energy for given $N$. Obviously, they correspond to
938: $m=1$, that is,
939: \be\label{Emin-naive}
940: \varepsilon^{(m=1)} \sim 4 \ln N \, , \qquad \widehat q_4^{ (0)} = \ldots =
941: \widehat q_L^{\scriptscriptstyle (0)}=0 \, .
942: \ee
943: Here the superscript `$\scriptstyle (0)$' refers to the leading order
944: approximation of the semiclassical expansion and $\widehat
945: q_{2k}^{\scriptscriptstyle (0)}=0$ does not necessarily imply that $\widehat
946: q_{2k}=0$ but rather $\widehat q_{2k}= \mathcal{O}(\eta)$, or equivalently
947: $q_{2k} \ll (-q_2)^{k}$.
948:
949: The transfer matrix $\tau_0(x)$, Eq.~\re{tau0}, corresponding to the minimal
950: energy state \re{Emin-naive} is given by
951: \be\label{tau-min}
952: \tau^{(m=1)}_{0}(x)=2-\frac1{x^2}
953: \ee
954: and the spectral curve \re{curve} looks like
955: \be\label{Gamma-min}
956: \Gamma_L^{(m=1)}: \qquad y^2 = \lr{x^2-\frac1{4}}\frac4{x^4}\,.
957: \ee
958: It is easy to see that $\Gamma_L^{(m=1)}$ coincides with the complex curve for
959: the spin chain of length $2$, that is $\Gamma_{L=2}$, Eq.~\re{curve}. Indeed,
960: choosing $\widehat q_{2m+1}^{\scriptscriptstyle (0)}=\ldots=\widehat
961: q_{L}^{\scriptscriptstyle (0)} = 0$ in \re{curve} one effectively descends from
962: an infinitely long spin chain (for $L\to\infty$) to the one with finite length
963: $2m$. Comparing \re{Gamma-min} with \re{y2-roots} we conclude that for $m=1$ all
964: but two branching points condense at the origin, $b=x_{2j}=0$, and the two
965: remaining (resolved) branching points are located at $\pm 1/2$. For $m\ge 2$, one
966: can show~\cite{KorKri97} that the complex curve corresponding to \re{E-m} is
967: given by
968: \be\label{Gamma-m}
969: \Gamma_L^{(m\ge 2)}: \qquad y^2=
970: \lr{x^2-\frac1{4m^2}}\frac4{x^4}\prod_{j=1}^{m-1}
971: \lr{1-\frac{x^2_{2j}}{x^2}}^2\,,
972: \ee
973: with $1/x_{2j}=2m\cos\lr{\frac{\pi j}{2m}}$. Comparison with \re{curve-reduced}
974: yields $a=1/(2m)$ and $b=0$, that is, the two cuts $[-a,-b]$ and $[b,a]$ collide
975: at the origin. Notice that $\Gamma_L^{(m\ge 2)}$ has exactly $2m$ double points
976: satisfying $x_{2j}^2 > 1/(2m)^2$ while the remaining double points in
977: \re{curve-reduced} belonging to the interval $[-b,b]$ condensed at $x=0$ as $b\to
978: 0$.
979:
980: The complex curves \re{Gamma-min} and \re{Gamma-m} have genus zero and, as a
981: consequence, the momentum $p'(x)$ is expressed in terms of elementary functions.
982: Indeed, one substitutes \re{Gamma-min} and \re{Gamma-m} into \re{dp} and obtains
983: the following expression for the momentum for arbitrary $m$
984: \be\label{dp-reduced}
985: dp = -\frac{dx}{x\sqrt{x^2-\frac1{(2m)^2}}}\,,\qquad
986: p (x) = \int^x_{\infty} dp =
987: 2 m \arcsin \left( \frac{1}{2m x} \right) \, .
988: \ee
989: Together with \re{p-def} and \re{tau0} this leads to the following expressions
990: for the integrals of motion
991: \be\label{q-expression}
992: \widehat q_{2n}^{\scriptscriptstyle (0)} = \frac{2 (- 1)^{n} m^{1 - 2 n} \Gamma
993: (m + n) }{ \Gamma (2 n + 1) \Gamma (m - n + 1)} \, .
994: \ee
995: We conclude from \re{dp-reduced} that for $m=1$, for the eigenstate with the
996: minimal energy, in the limit $N\to\infty$ and $L=\rm fixed$, the momentum $p'(x)$
997: is an analytical function on the complex plane with the square root cut
998: $[-\ft12,\ft12]$ and a {\sl simple} pole at the origin. This should be compared
999: with the analytical properties of the momentum in the region $N/L={\rm fixed}$ in
1000: which case $p'(x)$ has a {\sl double} pole at $x=0$, Eq.~\re{p-zero}.
1001:
1002: We demonstrated in this section that the analytical properties of the momentum
1003: $p(x)$ and the action function $S'(x)$ are quite different in the two limits
1004: mentioned above. For $0 < \beta < 1$, the Bethe roots condense on two symmetric
1005: intervals $[-a,-b]$ and $[b,a]$ with the end-points $a$ and $b$ depending on
1006: $\beta$. For $\beta\to 0$, one has $b\to 0$ and $a\to \ft12$ so that the two cuts
1007: collide and form a single cut $[-\ft12,\ft12]$ (see Fig.\ \ref{CollisionCutsFig}
1008: (a) and (b)). This leads to different scaling behavior of the energy in these two
1009: limits, Eqs.~\re{E-naive-expansion} and \re{Emin-naive}, respectively. The
1010: question remains what happens in the intermediate region of the parameter
1011: $\beta=Ls/(N+Ls)$, Eq.~\re{beta}, and how important the corrections $\sim \eta$
1012: to the energy \re{E-naive-expansion} are. We shall demonstrate in the next
1013: section that the semiclassical expansion of the energy \re{E-naive-expansion}
1014: becomes divergent for $\beta\to 0$ due to the collision of cuts and work out an
1015: asymptotic expression for the energy valid for small $\beta$.
1016:
1017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1018: % Figure
1019: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1020: \begin{figure}[t]
1021: \begin{center}
1022: \mbox{
1023: \begin{picture}(0,70)(240,0)
1024: \put(0,0){\insertfig{17}{collision}}
1025: \end{picture}
1026: }
1027: \end{center}
1028: \caption{\label{CollisionCutsFig}%
1029: Symmetric two-cut configuration (a) resulting in the BMN scaling of the anomalous
1030: dimension in gauge theory for $\xi < 1$. For $\xi\gg 1$ the two cuts collide at
1031: the origin (b) yielding the logarithmic scaling. The same configuration in string
1032: sigma model (c) for $\xi_{\rm str}\gg 1$ -- the minimal value for the inner end
1033: of the cut is given by the BMN coupling $\sqrt{\lambda^\prime}$ which prevents
1034: the cuts to collide.}
1035: \end{figure}
1036: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1037:
1038: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1039: \section{Symmetric two-cut solution in thermodynamic limit}
1040: \label{ThermodynamicSection}
1041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1042:
1043: In this section, we construct the minimal energy trajectory in the thermodynamic
1044: limit $L \gg 1$ for arbitrary values of the spin $N$. To begin with, we work out
1045: the semiclassical expansion of solutions $Q(u)$ to the Baxter equation
1046: \re{Baxter-eq} for $N/L={\rm fixed}$ and, then, extend the analysis to the region
1047: $N \gg L$.
1048:
1049: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1050: \subsection{Semiclassical expansion at leading order}
1051: \label{WKBloSection}
1052: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1053:
1054: To leading order of the semiclassical expansion \re{WKB-ansatz} and
1055: \re{Q-ansatz}, the function $S_0'(x)$ is determined by the momentum $p(x)$,
1056: Eq.~\re{p-def}. The latter is fixed by the condition that $dp(x)$ should be a
1057: meromorphic differential on the complex curve \re{curve-reduced} with prescribed
1058: asymptotic behavior at infinity, Eq.~\re{p-infinity}, and at the origin,
1059: Eq.~\re{p-zero}, on the upper sheet of $\Gamma_L$
1060: \be
1061: \label{p-asym}
1062: dp \stackrel{x\to\infty}{\sim} -\frac{dx}{x^2}\,,\qqqquad dp \stackrel{x\to
1063: 0}{\sim} dx \left[-\frac{\beta}{x^2} + \mathcal{O}(x^0)\right]\,,
1064: \ee
1065: with $\beta=Ls/(N+Ls)$, Eq.~\re{beta}. Being combined together, these conditions
1066: lead to the following expression~\cite{KazZar04}
1067: \be\label{dp-res}
1068: dp = \frac{-1+\frac{\beta ab}{x^{2}}}{\sqrt{(x^2-a^2)(x^2-b^2)}}\,dx\,.
1069: \ee
1070: It defines $dp$ as a meromorphic differential on the elliptic curve $\tilde y^2 =
1071: (x^2-a^2)(x^2-b^2)$, Eq.~\re{curve-reduced}, which has a pair of double poles
1072: above $x=0$ on both sheets of the corresponding Riemann surface. The values of
1073: the parameters $a$ and $b$ are fixed by the normalization conditions
1074: \re{p-periods}
1075: \be\label{cond}
1076: \int_{b}^a dp = 0\,,\qqqquad \int_a^\infty dp = - \pi m\,.
1077: \ee
1078: Here the integer $m$ defines the position of the interval $[b,a]$ inside
1079: \re{cuts}. We remind that in the limit we are interested in, the complex curve
1080: $\Gamma_L$ gets reduced to the genus one curve \re{curve-reduced}. Namely, all
1081: cuts in \re{cuts} except two, $[-a,-b]$ and $[b,a]$, shrink into points. The
1082: integer $m-1$ counts how many collapsed cuts are situated to the right from the
1083: interval $[b,a]$ on the real axis. The energy depends on $m$ and, as we will see
1084: in a moment, it takes the minimal value for $m=1$, that is, when all shrunken
1085: cuts are located inside the interval $[-b,b]$.
1086:
1087: Solving \re{cond} one finds the positions of the cuts as a function of the parameter
1088: $\beta$, Eq.~\re{beta}
1089: \be\label{ab}
1090: a=\frac1{2m\,\mathbb{E}(\tau)}\,,\qquad b =
1091: \frac{\beta}{2m}\frac1{\mathbb{K}(\tau)} \,,\qquad
1092: \beta=\sqrt{1-\tau}\frac{\mathbb{K}(\tau)}{\mathbb{E}(\tau)}\,,
1093: \ee
1094: where $\mathbb{K}(\tau)$ and $\mathbb{E}(\tau)$ are elliptic integrals of the
1095: first and second kind, respectively, and the modular parameter is defined as (for
1096: $b \le a$)
1097: \be\label{tau}
1098: \tau = 1 -\frac{b^2}{a^2}\,.
1099: \ee
1100: From \re{dp-res} and \re{p-def} one obtains the leading term in the semiclassical
1101: expansion as
1102: \be\label{SO-zero}
1103: S_0''(x) = p'(x) + \frac{\beta}{x^2}
1104: =
1105: \frac{-1+\frac{\beta ab}{x^{2}}}{\sqrt{(x^2-a^2)(x^2-b^2)}} + \frac{\beta}{x^2}
1106: \, ,
1107: \ee
1108: with the parameters $a$ and $b$ given by \re{ab}. One verifies that $S_0''(x)$ has
1109: a regular expansion around $x=0$ which is in agreement with our expectations that
1110: the Bethe roots condense on the intervals $[-a,-b]\cup [b,a]$ and there is no
1111: accumulation of roots at the origin. The energy \re{E-naive-expansion} is determined
1112: by the leading term in the expansion of this expression~\cite{BeiFroStaTse03}
1113: \be\label{Energy-2-cut}
1114: \varepsilon
1115: =
1116: -
1117: \frac{2\beta}{L} \left[{\frac {1}{ab}} - \frac{\beta}2\lr{\frac1{a^2}+\frac1{b^2}} \right]
1118: =
1119: \frac{(2m)^2}{L}
1120: \mathbb{K}(\tau) \left[ (2-\tau) \mathbb{K}(\tau) - 2 \mathbb{E}(\tau)\right]
1121: \, .
1122: \ee
1123: The dependence of $\varepsilon$ on the total spin $N$ and conformal spin $s$
1124: enters into this expression through the parametric dependence of the modular
1125: parameter $\tau$, Eq.~\re{ab}, on the scaling parameter $\beta=Ls/(N+Ls)$,
1126: Eq.~\re{beta}.
1127:
1128: We observe that the minimal energy corresponds to $m=1$, that is, when all double
1129: points $x_{2j}$, Eq.~\re{curve-reduced}, are located inside the interval
1130: $[-b,b]$. It is straightforward to find their position on the real axis with the
1131: help of \re{N=0-roots} and \re{dp-res}. Since the total number of double points
1132: is infinite for $L\to \infty$ and they occupy a compact interval, $x_{2j}$ should
1133: be a smooth function of $j$. Differentiating the second relation in
1134: \re{N=0-roots} with respect to $j$, one finds
1135: \be\label{dx}
1136: \frac{dx_{2j}}{dj} = \frac{\pi} {p'(x_{2j})} = - x_{2j}^2 \frac{\pi} {\beta}
1137: \left[ 1+ \mathcal{O}(x_{2j}^2)\right]\,,
1138: \ee
1139: where we substituted the momentum by its value at the origin \re{p-asym}. It
1140: follows from \re{dx} that for large $j$ the double points $1/x_{2j}$ are
1141: equidistantly distributed on the real axis (see Fig.~\ref{Fig-semi})
1142: \be\label{equid}
1143: \frac{1}{x_{2j}} = \frac{\pi}{\beta} j + \mathcal{O}(1/j)\,.
1144: \ee
1145: We remind that the double points $x_{2j}$ verify the relation $\tau_0(x_{2j})=\pm
1146: 2$. It is instructive to compare \re{equid} with a similar relation $1/{x_{2j}} =
1147: {\pi}j$ for the momentum \re{N=0-roots} corresponding to the state with $N=0$, or
1148: equivalently $\beta=1$. Equation \re{equid} suggests that close to the origin and
1149: away from the cuts, $x^2\ll b^2$, the transfer matrix for the two-cut solution
1150: looks like $\tau_0(x) = 2 \cos p(x) \sim 2 \cos(\beta /x)$. Matching this
1151: relation into the general expression for $\tau_0(x)$, Eq.~\re{tau-expansion}, we
1152: find that the integrals of motion scale as
1153: \be\label{q-scaling}
1154: \widehat q_{2n} \sim 2\frac{(-1)^{n}}{(2n)!}\beta^{2n}\,,\qquad (n \gg 1) \, ,
1155: \ee
1156: where $\widehat q_{2n}=q_{2n}/(N+Ls)^{2n}$. For $\beta=1$, or equivalently $N=0$,
1157: this relation is exact for all $\widehat q_{2k}$. For $\beta\to 0$, or
1158: equivalently $N\to\infty$, it cannot be exact since, by definition, $\widehat q_2
1159: = -1+\mathcal{O}(\eta)$. It instead indicates that higher charges take
1160: anomalously small values $\widehat q_{2k}=\mathcal{O} (\eta^{2k})$ with $\eta\sim
1161: 1/N$. This is in agreement with our expectations that for $\beta=0$ the transfer
1162: matrix reduces to \re{tau-min}.
1163:
1164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1165: % Figure
1166: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1167: \begin{figure}[t]
1168: \begin{center}
1169: \mbox{
1170: \begin{picture}(0,225)(150,0)
1171: \put(0,0){\insertfig{10}{quasi}}
1172: \end{picture}
1173: }
1174: \end{center}
1175: \caption{\label{Fig-semi} The transfer matrix $\cos p(x)=\tau_0(x)/2$ as a
1176: function of $1/x$ for the symmetric two-cut solution for $\beta=3/4$ and $m=1$.
1177: The cuts $[-a,-b] \cup [b,a]$ with $a=0.45$ and $b=0.15$ correspond to $\cos^2
1178: p(x)> 1$. The double points $x_{2j}$ are denoted by crosses, $\cos p(x_{2j})=\pm
1179: 1$ and $1/x_{2j}^2 > 1/b^2$. The ``large'' and ``small'' roots of the transfer
1180: matrix, $\cos p (\delta_n) = 0$, are shown by full and light blobs,
1181: respectively.}
1182: \end{figure}
1183: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1184:
1185: Let us examine the expression for the energy \re{Energy-2-cut} in two limiting
1186: cases $\beta\to 1$ and $\beta\to 0$, or equivalently $N \ll L$ and $N \gg L$,
1187: respectively, Eq.~\re{limits}. According to \re{ab}, the corresponding values of
1188: the modular parameter are $\tau\to 0$ and $\tau\to 1$. For $\tau\to 0$, one finds
1189: from \re{ab}
1190: \be\label{ab-small}
1191: a=\frac1{\pi m}\lr{1+\frac{\tau}4}+...
1192: \,,\qquad
1193: b=\frac1{\pi m}\lr{1-\frac{\tau}4}+...\,,\qquad\beta=1-\frac{\tau^2}{16}+...\,,
1194: \ee
1195: where the ellipses denote subleading terms. Since $a-b\sim \tau \to 0$, the two
1196: cuts shrink into points as $\beta\to 1$. One expands \re{Energy-2-cut} in powers
1197: of $\tau$ and obtains the energy as
1198: \be\label{E-small}
1199: \varepsilon \stackrel{L \gg N}{=} \frac{m^2}{8L} {\pi^2} \tau^2 + \ldots =
1200: \frac{N}{L^2} \frac{2 \pi^2 m^2}{s}+\ldots \,.
1201: \ee
1202: For $\tau\to 1$, one has from \re{ab}
1203: \be\label{ab-large}
1204: a=\frac1{2m}+...\,,\qquad b=\frac{\sqrt{1-\tau}}{2m}+ ... \,,\qquad
1205: \beta= \sqrt{1-\tau}\,\ln\frac{1}{\sqrt{1-\tau}}+ ...
1206: \ee
1207: According to \re{beta}, the limit $\beta\to 0$ corresponds to $\beta=Ls/N$ with
1208: $N \gg L$. Then, making use of \re{Energy-2-cut} one finds for the energy
1209: \be\label{E-large}
1210: \varepsilon \stackrel{L \ll N}{=} \frac{4m^2}{L} \ln^2 \frac{\sqrt{1-\tau}}{4} +
1211: \ldots = \frac{4m^2}{L} \ln^2 \frac{N}{L}+ \ldots \, .
1212: \ee
1213: This relation matches the expression for the energy of long spinning string
1214: folded $m$ times~\cite{FroTse03} and agrees with the thermodynamic Bethe ansatz
1215: analysis \cite{BeiFroStaTse03}.
1216:
1217:
1218: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1219: \subsection{Breakdown of semiclassical expansion}
1220: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1221:
1222: For $m=1$ the expressions \re{Energy-2-cut}, \re{E-small} and \re{E-large}
1223: describe the dependence of the energy on the total spin $N+Ls$ along the minimal
1224: energy trajectory at leading order of the semiclassical expansion for $L\gg 1$
1225: and arbitrary values of $N$ including the regions $N\ll L$ and $N\gg L$. {}From
1226: \re{E-naive-expansion} one expects that the contribution of the subleading terms
1227: to the energy is suppressed by powers of $\eta=1/(L+Ns)$. Let us compare these
1228: predictions with the exact expressions for the minimal energy obtained from
1229: numerical solution to the Baxter equation, Eqs.~\re{Baxter-eq} and
1230: \re{Energy-Baxter}. For $L=10$, the comparison is shown in
1231: Fig.~\ref{Fig-roots-energy10} on the right panel. We observe that the curve
1232: determined by \re{Energy-2-cut} agrees with the exact numbers for $N < L$ and
1233: deviates significantly from them for $N > L$. Most importantly, for $N \gg L$ the
1234: exact energy scales as $\sim 4\ln N$ while the semiclassical formula \re{E-large}
1235: yields a different asymptotic behavior.
1236:
1237: To understand the reason for this discrepancy, let us revisit the calculation of
1238: the energy in the previous section and pay special attention to the position of
1239: the cuts. By construction, the cuts run along two symmetric intervals
1240: $[-a,-b]\cup[b,a]$. According to \re{ab-large}, for $\tau\to 1$ the parameter $a$
1241: approaches the value $a=1/(2m)$ while the parameter $b$ vanishes indicating that
1242: the two cuts collide at the origin. Then, one examines the expression for the
1243: momentum \re{dp-res} and finds that for $b=0$ the differential $dp$ reduces to
1244: \re{dp-reduced}. We argued in Sect.~\ref{CollisionCuts} that Eq.~\re{dp-reduced}
1245: leads to a new, logarithmic scaling of the energy \re{Emin-naive}. This suggests
1246: that the problem with recovering this regime within the semiclassical approach
1247: does not lie in the construction of the momentum $p(x)$ but rather in the expression
1248: for the energy \re{E-naive-expansion}.
1249:
1250: The exact formula for the energy \re{Energy-Baxter} involves $S'(x_\pm)$
1251: evaluated at $x_\pm=\pm i\beta/L$. The relation \re{E-naive-expansion} was
1252: derived under the assumption that $S'(x)$ is analytical at the vicinity of these
1253: points. For the two-cut solution constructed in the previous section this
1254: assumption is justified provided that the inner boundary of the cut $b$ stays
1255: finite in the limit $L\to\infty$. For $x_\pm \to 0$ with $b={\rm fixed}$, the
1256: semiclassical analysis is applicable and one arrives at the asymptotic behavior
1257: \re{E-small} and \re{E-large} depending on the ratio $N/L$. The situation is more
1258: complex when $b$ vanishes in the scaling limit. For $b\to 0$ with $x_\pm ={\rm
1259: fixed}$, the two cuts collide and the very assumption that $S'(x)$ is analytical
1260: at the origin does not hold true anymore. In this case, one expects that the
1261: energy will develop a new asymptotic behavior and the crossover will occur for $b
1262: \sim |x_\pm|$ (see Fig.\ \ref{CollisionCutsFig} (a) and (b)).
1263:
1264: Defining a new parameter $\xi =|x_\pm|/b= {\beta}/{(bL)}$ one anticipates that
1265: the semiclassical approach is applicable for $\xi < 1$. Indeed, for $N\ll L$ one
1266: finds from \re{ab-small} that $\xi \sim 1/L$ and the condition $\xi < 1$ is
1267: satisfied. For $N\gg L$ one gets from \re{ab-small}
1268: \be
1269: \xi \sim \frac1{L} \ln \frac{N}{L}\,.
1270: \ee
1271: One recognizes that similar parameter $\xi_{\rm str}$ naturally arises on the
1272: string side and controls there the transition between two different regimes in
1273: \re{3rd}. For $\xi \gg 1$ one has to take into account that the two cuts collide at the
1274: origin and, as a consequence, the analytical properties of $S'(x)$ are different.
1275: The simplest way to see this is to examine the Taylor expansion of the leading
1276: term $S_0''(x)$, Eq.~\re{SO-zero}, around the origin and retain the terms most
1277: singular for $b\to 0$
1278: \be
1279: S_0''(x) \sim \frac{\beta}{x^2}\left[1 -
1280: \frac1{\sqrt{1-x^2/b^2}}\right]=-\frac{\beta}{2b^2}\left[1+\frac34\,{\frac
1281: {{x}^{2}}{{b}^{2}}}+\frac58\,{\frac {{x}^{4}}{{b}^{4}}}+\ldots
1282: \right]\,.
1283: \ee
1284: As follows from \re{tau-expansion} and \re{E-naive-expansion}, higher order terms
1285: of this expansion contribute to subleading corrections to the energy
1286: $\varepsilon=i\int_{x_-}^{x_+} dx \, S''(x)$. It is easy to see that the series
1287: for $i\int_{x_-}^{x_+} dx \, S_0''(x)$ runs in powers of $\xi=|x_\pm/b|$ and it
1288: is only convergent for $\xi < 1$.
1289:
1290: Based on our analysis we expect that the semiclassical expansion of the energy
1291: \re{E-naive-expansion} will be divergent for $\xi \gg 1$. Let us calculate the
1292: first subleading correction to the energy \re{E-naive-expansion} defined by
1293: $S_1''(0)$. The function $S_1'(x)$ satisfies the relation \re{S1-Baxter} which
1294: involves yet unknown function $\tau_1(x)$. To determine $S_1'(x)$ one requires
1295: that it should have the same analytical properties as the leading term of the
1296: semiclassical expansion $S_0'(x)$, i.e., to be an analytical function in the
1297: complex plane with the cuts $[-a,-b]\cup [b,a]$. Discontinuity of $S_1'(x)$
1298: across the cut defines the correction to the distribution density of the Bethe
1299: roots~\cite{Kor95}. Since $\tau_0(x)$ and $\tau_1(x)$ are entire functions while
1300: $p'(x)$ and $\sin p(x)$ change sign across the cut, $\sin p(x+i0)=-\sin p(x-i0)$,
1301: the function $S_1'(x)$ satisfies the relation
1302: \be
1303: S_1'(x+i0) + S_1'(x-i0) = -p'(x+i0) \cot p(x+i0)\,,
1304: \ee
1305: with $x \in [-a,-b]\cup [b,a]$. In addition, its asymptotic behavior at the
1306: origin and infinity is fixed by Eqs.~\re{as-infinity} and \re{as-zero},
1307: \be
1308: S_1'(0)=0\,,\qqqquad S_1'(x)\stackrel{x\to\infty}{\sim} \frac1{x^3}\,.
1309: \ee
1310: The solution to the resulting Riemann-Hilbert problem looks like (c.f.
1311: \cite{GroKaz05})
1312: \be
1313: \label{S1intermediate} S_1'(x) = \frac{x}{\sqrt{(x^2-a^2)(x^2-b^2)}}\int_b^a
1314: \frac{dz}{\pi}\frac{\sqrt{(a^2-z^2)(z^2-b^2)}}{x^2-z^2} p'(z+i0) \cot p(z+i0)\,,
1315: \ee
1316: where `$+i0$' indicates that the momentum is evaluated on the upper sheet of the
1317: Riemann surface. Replacing $p'(z+i0)$ by its expression \re{dp-res}, we obtain
1318: the following representation for the first subleading correction to
1319: \re{E-naive-expansion}
1320: \be
1321: \varepsilon_1 = -\frac{2\beta}{L}S_1''(0) = \frac{2\beta}{\pi ab L}\int_b^a
1322: \frac{dz}{z^2} \left[1-\beta\frac{ab}{z^2}\right] \coth(ip(z+i0)) \,.
1323: \ee
1324: For our purposes we would like to determine the asymptotic behavior of
1325: $\varepsilon_1$ for $N \gg L$, or equivalently $a\to 1/(2m)$, $b\to 0$ and
1326: $\beta\to 0$, Eq.~\re{ab-large}. Changing the integration variable to $z\to b z$,
1327: one finds in the limit $b\to 0$
1328: \be\label{eps1}
1329: \varepsilon_1 = \frac{2\beta}{\pi ab^2L}\int_1^\infty
1330: \frac{dz}{z^2}\left[1-\frac{\beta a}{b}\frac{1}{z^2}\right] \coth(i p(b z+i0))=
1331: -\frac{2}{3\pi}\frac{\beta^2}{L b^3} + \ldots\, .
1332: \ee
1333: Here in the second relation we took into account that for $\beta\to 0$ the cuts
1334: collide at the origin and the momentum scales as $p(z+i0) \sim 2im \ln z$ for $z
1335: \to 0$, Eq.~\re{dp-reduced}. Comparing \re{eps1} with a similar relation for the
1336: leading term $\varepsilon_0 \sim \beta^2/(L b^2)$, Eq.~\re{Energy-2-cut}, one
1337: finds that
1338: \be
1339: \frac{\eta \varepsilon_1}{\varepsilon_0} = \mathcal{O}(\eta /b)=\mathcal{O}(\xi)
1340: \, ,
1341: \ee
1342: where $\eta/b=1/(b N)=(2m/s) \ln(N/L)/L$ in the limit $N\gg L$, Eqs.~\re{beta}
1343: and \re{ab-large}.
1344:
1345: We conclude that for $N\gg L$ the semiclassical expansion of the energy,
1346: $\varepsilon=\varepsilon_0+\eta\varepsilon_1+\mathcal{O}(\eta^2)$,
1347: Eq.~\re{E-naive-expansion} runs in powers of $\xi$ and it is only convergent for
1348: $\xi <1$. Together with \re{gamma=energy} this leads to the expression for the
1349: one-loop anomalous dimension, Eqs.~\re{gen-exp} and \re{coeff-fun}, with
1350: $c_{1,1}=-2m/(3\pi s)$.
1351:
1352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1353: \subsection{Beyond semiclassical expansion}
1354: \label{BeyondSection}
1355: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1356:
1357: We demonstrated in the previous section that the semiclassical expansion breaks
1358: down for $\xi\gg 1$. The reason for this is that the cuts collide in this limit
1359: and the semiclassical expression for $S(x)$ ceases to be analytical at $x=0$.
1360: Substituting $\beta=0$ into \re{SO-zero}, one finds (for $m=1$)
1361: \be
1362: \label{P-log}
1363: S_0''(x) = p'(x)= -\frac1{x\sqrt{x^2-\frac14}}
1364: \ee
1365: and, therefore, $S_0'(x) \sim \mp 2i \ln x$ for $x\to 0$ on the upper and lower
1366: sheets, respectively. It is easy to see that the corresponding transfer matrix
1367: $\tau_0(x)=2\cos p(x)$ is given by \re{tau-min}. We remind that $\tau_0(x)$ plays
1368: the role of the potential in the Baxter equation \re{Baxter-eq}. The fact that
1369: $\tau_0(x)$ is singular at $x=0$ implies that the semiclassical expansion
1370: \re{WKB-ansatz} breaks down at the origin. In this section, we present an
1371: approach which allows one to construct asymptotic solution to the Baxter equation
1372: \re{Baxter-eq} and evaluate the energy for $N > L$. It is complementary to the
1373: semiclassical approach and takes a full advantage of the above mentioned
1374: singularity of the transfer matrix $\tau_0(x)$ at $x=0$. A detailed account on
1375: this approach can be found in Ref.~\cite{DerKorMan99}.
1376:
1377: To begin with, we rewrite the Baxter equation \re{Baxter-eq} as
1378: \be\label{Baxter-phi}
1379: (u+is)^L \phi(u) + \frac{(u-is)^L}{\phi(u-i)} = t_L(u)
1380: \ee
1381: where the notation was introduced for
1382: \be
1383: \phi(u)=\frac{Q(u+i)}{Q(u)}\,.
1384: \ee
1385: To evaluate the energy \re{Energy-Baxter} one has to construct solutions to
1386: \re{Baxter-phi} for $u\sim \pm is$. Notice that the dressing factors $(u\pm
1387: is)^L$ in \re{Baxter-phi} vanish for $u\to \mp is$ indicating that one of the
1388: terms in the left-hand side of \re{Baxter-phi} becomes anomalously small and can
1389: be neglected. It turns out that in the thermodynamic limit, $L \gg 1$, the same
1390: approximation can be performed not only in the vicinity of $u=\pm is$ but in the
1391: whole region $u=\mathcal{O}\left((N+Ls)^0\right)$. To see this we note that
1392: $t_L(u)$ is given by \re{t_L} with the charges that scale as $q_k \sim (N+Ls)^k$.
1393: This suggests that $|t_L(u)| \gg 1$ for $u=\mathcal{O}\left((N+Ls)^0\right)$.
1394: Indeed, in the semiclassical approach, for $u=x/\eta$, one finds from
1395: \re{tau-expansion}, \re{p-def} and \re{P-log} that the transfer matrix takes the
1396: form
1397: \be\label{t-p}
1398: t_L(x/\eta) = 2 (x/\eta)^L \cos p(x) + \mathcal{O}(\eta)\,.
1399: \ee
1400: For $u=\mathcal{O}(L^0)$, or equivalently $x\sim \eta$, the momentum can be
1401: replaced by its asymptotic behavior at the origin: $p(x) \sim\beta/x$ for
1402: $\xi\ll 1$ and $p(x)\sim 2i\ln x$ for $\xi \gg 1$ leading to $|t_L(x/\eta)|\gg
1403: 1$.
1404:
1405: Let us return to Eq.\ \re{Baxter-phi} and take into account that $t_L(u)$ takes
1406: large values for $u=\mathcal{O}\left((N+Ls)^0\right)$ both for $\xi< 1$ and
1407: $\xi\gg 1$. Requiring the left-hand side of \re{Baxter-phi} to be as large as
1408: $t_L(u)$ one finds that either $\phi(u) \gg 1$, or $\phi(u-i) \ll 1$. In both
1409: cases, one of the terms in the left-hand side of \re{Baxter-phi} can be safely
1410: neglected and one arrives at two different equations
1411: \ba\label{Baxter-as}
1412: (u+is)^L Q_+(u+i) \!\!\!&=&\!\!\! t_L(u) Q_+(u)\,,
1413: \\ \nonumber
1414: (u-is)^L Q_-(u-i) \!\!\!&=&\!\!\! t_L(u) Q_-(u)\,.
1415: \ea
1416: Having solved this system, one can construct an asymptotic solution to the Baxter
1417: equation \re{Baxter-eq} as a linear combination of $Q_+(u)$ and $Q_-(u)$
1418: \be\label{Q-as}
1419: Q^{\rm (as)}(u) = c_+ Q_+(u) + c_- Q_-(u)\,.
1420: \ee
1421: We would like to stress that $Q^{\rm (as)}(u)$ does not satisfy the Baxter
1422: equation \re{Q-polynom}, but asymptotically approaches its solution $Q(u)$ in the
1423: region $u=\mathcal{O}\left((N+Ls)^0\right)$. Equations~\re{Baxter-as} were
1424: obtained under the assumption that $(u+is)^L Q_+(u+i)\gg (u-is)^L Q_+(u-i)$ and
1425: $(u+is)^L Q_-(u+i)\ll (u-is)^L Q_-(u-i)$, respectively. Together with
1426: \re{Baxter-as} it can be expressed as the following relation for the transfer
1427: matrix
1428: \be
1429: t_L\left(u+\ft{i}2\right) t_L\left(u-\ft{i}2\right) \gg \left[u^2 +
1430: \left(\ft12-s\right)^2\right]^L\,.
1431: \ee
1432: For $u=\mathcal{O}\left((N+Ls)^0\right)$ it is equivalent to $|t_L(u)|\gg 1$.
1433:
1434: To solve \re{Baxter-as} one introduces into consideration the roots of the
1435: transfer matrix \re{t_L}
1436: \be\label{t-roots}
1437: t_L(u) = 2 \prod_{n=1}^L (u-\delta_n)\,.
1438: \ee
1439: It is known that for polynomial solutions to the Baxter equation,
1440: Eq.~\re{Q-polynom}, the roots take real values only, $\Im
1441: \delta_k=0$~\cite{Kor95}. Matching \re{t-roots} into \re{t_L} one finds that they
1442: satisfy the sum rules
1443: \be\label{sum_rules}
1444: \sum_{n=1}^L \delta_n = 0 \,,\qquad \sum_{n=1}^L \delta_n^2 = -\frac12q_2
1445: \,,\quad \ldots\, , \quad \prod_{n=1}^L \delta_n = \frac{(-1)^L}{2}q_L \,.
1446: \ee
1447: For even solutions to the Baxter equation, $Q(u)=Q(-u)$, the roots appear in
1448: pairs $\delta_n = -\delta_{L - n + 1}$. Making use of \re{t-roots}, it is
1449: straightforward to verify that the solutions to \re{Baxter-as} are given by
1450: \ba\label{Q+-}
1451: Q_+(u) = 2^{-iu} \prod_{n=1}^L\frac{\Gamma(-iu + i\delta_n)}{\Gamma(-iu +s )} \,
1452: , \qquad Q_-(u) = 2^{iu} \prod_{n=1}^L\frac{\Gamma(iu - i\delta_n)}{\Gamma(iu +s
1453: )} \, .
1454: \ea
1455: To fix the constants $c_\pm$ in \re{Q-as} one examines the relation for the
1456: quasimomentum in Eq.\ \re{Energy-Baxter} and substitutes $Q(u)$ by its asymptotic
1457: expression \re{Q-as}. Taking into account that (for $\varrho \to 0$)
1458: \be\label{Q-vanish}
1459: Q_+(-is + \varrho) \sim \varrho^L \,,\qquad Q_-(is - \varrho) \sim \varrho^L
1460: \ee
1461: one finds from \re{Energy-Baxter} and \re{Q-as}
1462: \be
1463: \e^{i\theta} = \frac{Q^{\rm (as)}(is)}{Q^{\rm (as)}(-is)}
1464: =\frac{c_+}{c_-}\frac{Q_+(is)}{Q_-(-is)}\,.
1465: \ee
1466: Therefore, for cyclically symmetric states $\e^{i\theta}=1$ the asymptotic
1467: solution to the Baxter equation is given up to an overall normalization factor by
1468: \be\label{Q-as-sol}
1469: Q^{\rm (as)}(u) = Q_+(u)Q_-(-is) + Q_-(u)Q_+(is)\,.
1470: \ee
1471: We remind that this relation is only valid in the region
1472: $u=\mathcal{O}\left((N+Ls)^0\right)$. In a similar manner, one uses
1473: \re{Q-vanish} to evaluate the energy \re{Energy-Baxter} as
1474: \be
1475: \varepsilon^{\rm (as)} = i\lr{\ln Q^{\rm (as)}(is)}'-i\lr{\ln Q^{\rm (as)}(-is)}'
1476: = i\lr{\ln Q_+(is)}'-i\lr{\ln Q_-(-is)}' \, .
1477: \ee
1478: Replacing $Q_\pm(u)$ by their expressions \re{Q+-} one finds the following
1479: remarkable expression for the energy in terms of the roots of the transfer matrix
1480: \re{t-roots}~\cite{Kor95,BraDerMan98,Bel99,DerKorMan99}
1481: \be\label{energy-as}
1482: \varepsilon^{\rm (as)} = 2 \ln 2 + \sum_{n=1}^L \left[\psi(s+
1483: i\delta_n)+\psi(s-i\delta_n) -2 \psi(2s) \right] \,,
1484: \ee
1485: where $\psi(x)=d\ln\Gamma(x)/dx$. The interpretation of \re{energy-as} in terms
1486: of classical $SL(2)$ spin magnet and properties of the Wilson operators \re{O-def}
1487: in gauge theory can be found in Ref.~\cite{BelGorKor03}.
1488:
1489: Since the roots $\delta_n$ are functions of the conserved charges,
1490: Eq.~\re{sum_rules}, the relation \re{energy-as} establishes the dependence of the
1491: energy on $q_2,\ldots,q_L$. To check \re{energy-as}, we compared the asymptotic
1492: ``dispersion curve'' $\varepsilon^{\rm (as)}(q_2,\ldots,q_L)$ with the exact one
1493: $\varepsilon^{\rm (ex)}(q_2,\ldots,q_L)$ coming from the Bethe Ansatz solution
1494: \re{Energy-ABA}. We found that for $L=10$ and $s=1/2$ the accuracy of
1495: \re{energy-as}, $\varepsilon^{\rm (as)}/\varepsilon^{\rm (ex)}-1 $, increases
1496: from $-4.6 \times 10^{-5}$ for $N=2$, to $-2.6 \times 10^{-6}$ for $N=10$ and to
1497: $1.9 \times 10^{-10}$ for $N=100$. We conclude that Eq.~\re{energy-as} describes
1498: the exact eigenspectrum with a high accuracy throughout the whole interval of $N$
1499: including the region $N\sim L$ in which the semiclassical approach is applicable.
1500: Indeed, it is straightforward to show that the relation \re{energy-as} coincides
1501: with the semiclassical expression \re{E-naive-expansion} for $N\sim L$ (see
1502: Appendix).
1503:
1504: The charges $q_3,\ldots,q_L$ have to satisfy quantization conditions. In the
1505: method of Baxter $Q-$operator they follow from the requirement for $Q(u)$ to be
1506: polynomial solutions to the Baxter equation, Eqs.~\re{Baxter-eq} and
1507: \re{Q-polynom}. The roots of $Q(u)$ scale in the thermodynamic limit as
1508: $u=\lambda_j \sim L$ and, therefore, they lie outside the applicability range of
1509: \re{Q-as-sol}. This does not allow us to impose the polynomiality condition for
1510: $Q^{\rm (as)}(u)$. We shall require instead that $Q^{\rm (as)}(u)$, given by
1511: \re{Q-as-sol}, should be regular on the real $u-$axis inside the region
1512: $u=\mathcal{O}\left((N+Ls)^0\right)$. This property is not warranted since both
1513: $Q_+(u)$ and $Q_-(u)$ develop poles on the real $u-$axis originating from the
1514: product of $\Gamma-$functions in the numerator in the right-hand side of
1515: Eq.~\re{Q+-}. The poles are located at zeros of the transfer matrix \re{t-roots},
1516: $u=\delta_j$ with $j=1,\ldots,L$. Requiring for $Q^{\rm (as)}(u)$ to have zero
1517: residues at $u=\delta_n$ with $\delta_n=\mathcal{O} \left((N+Ls)^0\right)$, one
1518: gets from \re{Q-as-sol} and \re{Q+-}
1519: \be\label{quant-cond}
1520: 2^{-2i\delta_n} \prod_{j=1, j\neq n}^L
1521: \frac{\Gamma(-i\delta_n+i\delta_j)}{\Gamma(i\delta_n-i\delta_j)}=
1522: \left[\frac{\Gamma(s-i\delta_n)}{\Gamma(s+i\delta_n)}\right]^L \prod_{j=1}^L
1523: \frac{\Gamma(s+i\delta_j)}{\Gamma(s-i\delta_j)}\,.
1524: \ee
1525: The total number of roots of the transfer matrix \re{t-roots} equals the length
1526: of the spin chain $L$. As we will see in a moment, in the thermodynamic limit $L
1527: \gg 1$ and $N \gg L$ all roots can be separated into two different groups
1528: depending on their scaling: ``small'' roots $\delta_n=\mathcal{O}(N^0)$ and
1529: ``large'' roots $\delta_n=\mathcal{O}(N)$. It is important to realize that the
1530: quantization conditions \re{quant-cond} should only hold for the small roots
1531: while the product over $j$ entering both sides of \re{quant-cond} involves {\sl
1532: all} roots.
1533:
1534: The total number of small roots depends on the value of the parameter
1535: $\beta=Ls/(N+Ls)$. For $0 < \beta \le 1$, that is, within the applicability range
1536: of the semiclassical expansion, the integrals of motion scale as in
1537: Eq.~\re{q-scaling}. Together with \re{sum_rules} this implies that the roots
1538: behave as $\delta_n = \mathcal{O}(N)$ and, therefore, there are no small roots.
1539: For $\beta\to 0$ higher charges $q_L, q_{L-1}, \ldots $ take anomalously small
1540: values \re{q-scaling} indicating that the small roots are there. Let us
1541: demonstrate that their total number equals $L-2m$ with positive integer $m$
1542: entering the right-hand side of \re{cond}. One makes use of \re{t-p} to deduce
1543: that the large roots of the transfer matrix satisfy $\cos p(\delta_j\eta)=0 +
1544: \mathcal{O}(\eta)$. Recall that the momentum verifies $(\cos p(x))^2 \ge 1$ on
1545: the cuts (for $b^2 \le x^2 \le a^2$) and takes the values $\cos p(x)=\pm 1$ at
1546: the double points $x=x_{2j}$, Eq.~\re{equid}. Since $\cos p(x_{2j})=-\cos
1547: p(x_{2j-2})$, the roots of the transfer matrix \re{t-p} should lie on the real
1548: axis in between the branching points, $x_{2k} < \delta_j\eta < x_{2j-2}$, away
1549: from the cuts $[-a,-b]\cup [b,a]$ (see Fig.~\ref{Fig-semi}). Going over to the
1550: limit $\beta =Ls/(N+Ls)\to 0$, or equivalently $N \gg L$, one takes into account
1551: that the edges of the cuts scale as $a = 1/{(2m)}$ and $b=\beta/(-2m\ln\beta)\to
1552: 0$, Eq.~\re{ab-large}. As a result, all roots of the transfer matrix
1553: $\delta_j\eta\sim \delta_j/N$ ``trapped'' inside the interval $[-b,b]$ are to be
1554: small while the roots belonging to the intervals $(-\infty,-a]\cup [a,\infty)$
1555: are to be large
1556: \be
1557: \label{roots-as} |\delta^{\rm (large)}| > \frac{N}{2m} \,,\qqqquad |\delta^{\rm
1558: (small)}| < \frac{1}{4m\xi}\,,
1559: \ee
1560: with $\xi=\ln (N/L)/L$. Thus, the total number of small roots, $L-2m$, equals the
1561: number of double points on the interval $[-b,b]$ and, as a consequence, the total
1562: number of the large roots is $2m$.
1563:
1564: Let us consider the minimal energy state with $m=1$. For $N \gg L$, the transfer
1565: matrix has two large roots $\delta_1=-\delta_{L}$ and $L-2$ small roots
1566: $\delta_k=-\delta_{L-k+1}$ with $k=2,\ldots,L-1$. It follows from \re{sum_rules}
1567: that the large root is given by
1568: \be\label{delta_1}
1569: \delta_1 = {(-q_2/2)^{1/2}} \left[ 1+ \mathcal{O}(1/L)\right]
1570: \sim
1571: \frac{N+Ls}{\sqrt{2}}
1572: \ee
1573: with $q_2$ defined in \re{q2}. The small roots satisfy the quantization
1574: conditions \re{quant-cond}. Separating in \re{quant-cond} the contribution of the
1575: large roots with $j=1$ and $j=L$, one can rewrite \re{quant-cond} in terms of the
1576: small roots only (for $n=2,\ldots,L-1$)
1577: \be\label{system}
1578: (-q_2)^{-2i\delta_n} = (-1)^{L-1}
1579: \left[\frac{\Gamma(s-i\delta_n)}{\Gamma(s+i\delta_n)}
1580: \right]^{L}\prod_{j=2}^{L-1}
1581: \frac{\Gamma(1+i\delta_n-i\delta_j)}{\Gamma(1-i\delta_n+i\delta_j)} \,.
1582: \ee
1583: Here we took into account that the product over $j$ in the right-hand side of
1584: \re{quant-cond} equals $1$ due to pairing of the roots
1585: $\delta_n=-\delta_{L-n+1}$. Taking the logarithm of both sides in \re{system},
1586: one rewrites the quantization conditions as
1587: \be\label{system-arg}
1588: \delta_n \ln(-q_2) + L \,{\rm arg}\, \Gamma(s-i\delta_n) + \sum_{j=2}^{L-1} {\rm
1589: arg}\, \Gamma(1+i\delta_n-i\delta_j) = \frac{\pi}2 k_n
1590: \ee
1591: where integers $k_2> k_3> \ldots$ count the branches of the logarithms and
1592: satisfy $k_n=-k_{L-n+1}$. In addition, they have a parity opposite to that of
1593: $L$, that is $k_n={\rm even/odd}$ for $L={\rm odd/even}$. Notice that in
1594: distinction with the Bethe ansatz equations \re{Bethe-roots}, the number of
1595: relations in \re{system-arg} does not depend on the total spin $N$ but only on
1596: the length of the spin chain $L$.
1597:
1598: To evaluate the energy \re{energy-as} one has to solve the system \re{system-arg}
1599: and determine the set of small roots. The resulting expression for
1600: $\varepsilon^{\rm (as)}$, Eq.~\re{energy-as}, depends on integers $k_n$ and as we
1601: show below (see Eq.~\re{Ek}) it takes minimal value for the occupation numbers
1602: $k_n=L+1-2n$. A systematic analysis of the quantization conditions
1603: \re{system-arg} will be given elsewhere. The system of equations \re{system-arg}
1604: can be easily solved in two limits $\xi \gg 1$ and $\xi < 1$.
1605:
1606: For $\xi \gg 1$ (or $\ln(N/L) \gg L $) one deduces from \re{roots-as} that
1607: solutions to \re{system-arg} have to satisfy $|\delta_n| \ll 1$. In this case,
1608: one expands the $\Gamma-$functions in \re{system-arg} in powers of $\delta$'s and
1609: one finds after some algebra
1610: \be\label{delta-fin}
1611: \delta_n = \frac{\pi k_n/2}{\ln(-q_2) + (L-2)\psi(1)-L\psi(s)}+\ldots
1612: \ee
1613: with $n=2,\ldots,L-1$. One verifies a posteriori that this relation is in
1614: agreement with \re{roots-as}.
1615:
1616: For $\xi < 1$ (or $1 \ll \ln(N/L) < L $), one obtains from \re{roots-as} that
1617: $|\delta_n| < 1/(2\xi)$ and, therefore, not all roots verify the relation
1618: $|\delta_n| \ll 1$. Still, the relation \re{delta-fin} is valid for the roots
1619: with small absolute value $|\delta_n| \ll 1$ (with $n\sim L/2$). For roots
1620: $|\delta_n| \sim 1/(2\xi)$ (with $n \sim L^0$) one expands the $\Gamma-$functions
1621: in \re{system-arg} in inverse powers of $\delta$'s and obtains
1622: \be\label{dev}
1623: \delta_n\ln(-q_2) - \delta_n \ln \delta_n^2 +\ldots =\frac{\pi}{2} k_n\,,
1624: \ee
1625: with $k_n = \mathcal{O}(L)$. From this relation one obtains the relation
1626: $\delta_n \sim \pi k_n/(4 \ln(N/k_n))$ which is in agreement with \re{roots-as}.
1627:
1628: To test the quantization conditions \re{system-arg} we compared solutions to
1629: \re{system-arg} for $k_n=L+1-2n$ with the exact, Bethe ansatz expressions for the
1630: roots of the transfer matrix corresponding to the minimal energy state with the
1631: quantum numbers $s=\ft 12$, $L=10$ and $N=100$ and observed a good agreement (see
1632: Table \ref{tab:WKB}).
1633: %
1634: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1635: \begin{table}[t]
1636: \begin{center}
1637: \begin{tabular}{|c||c|c|c|c|c||c|}
1638: \hline & $\delta_1$ & $\delta_2$ & $\delta_3$ &
1639: $\delta_4$ & $\delta_5$ &
1640: $E$
1641: \\
1642: \hline exact & $73.897289$ & $0.5297596$ & $0.3443964$ & $0.1937491$ &
1643: $0.06253897$ & $7.3790455$
1644: \\
1645: \hline asym & $73.900271$ & $0.5297573$ & $0.3443955$ & $0.1937487$ &
1646: $0.06253886$ & $7.3791719$
1647: \\
1648: \hline
1649: \end{tabular}
1650: \end{center}
1651: \caption{Comparison of the exact roots, $\delta_n=-\delta_{L-n+1}$, and the
1652: energy, $E$, for $s=1/2$, $L=10$ and $N=100$ with the asymptotic expressions
1653: obtained from Eqs.~\re{system} and \re{energy-as}, respectively.} \label{tab:WKB}
1654: \end{table}
1655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1656: %
1657: In agreement with \re{delta-fin} and \re{dev}, the small roots $\delta_n$ vary
1658: linearly with $n$ close to the origin and deviate from the linear behavior close
1659: to the end points. Moreover, for $\xi \gg 1$ all small roots scale linearly with
1660: $n$ in agreement with \re{delta-fin} (see Fig.~\ref{Fig-roots-energy10}).
1661:
1662: In a similar manner, we solved quantization conditions \re{system-arg} for
1663: $s=\ft12$, $L=10$ and $0 \le N \le 100$ and compared the resulting expression for
1664: the energy \re{energy-as} with the exact expression \re{Energy-ABA} as shown in
1665: Table~\ref{Fig:energy}.
1666: %
1667: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1668: \begin{table}[t]
1669: \begin{center}
1670: \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|c|}
1671: \hline $N$ & $0$ & $10$ & $20$ &
1672: $30$ & $40$ &
1673: $50$ & $60$ & $70$ & $80$ & $90$
1674: \\
1675: \hline exact & $0$ & $2.2766$ & $3.5069$ & $4.3644$ & $5.0266$ & $5.5678$ &
1676: $6.0262$ & $6.4243$ & $6.7765$ & $ 7.0924$
1677: \\
1678: \hline asym & $0.0907$ & $2.2851$ & $3.5097$ & $4.3657$ & $5.0274$ & $5.5683$ &
1679: $6.0265$ & $6.4246$ & $6.7767$ & $7.0926$
1680: \\
1681: \hline
1682: \end{tabular}
1683: \end{center}
1684: \caption{Comparison of the exact energy, $\varepsilon$, with the asymptotic
1685: expression obtained from Eqs.~\re{system} and \re{energy-as} for $s=\ft12$,
1686: $L=10$ and different total spin $N$.} \label{Fig:energy}
1687: \end{table}%
1688: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1689: We found that for $N \ge L$ the asymptotic expression \re{energy-as} approximates
1690: the exact result with an accuracy better than $0.002\%$ while the semiclassical
1691: approach significantly overestimates the value of the energy (see
1692: Fig.~\ref{Fig-roots-energy10}). Remember that the minimal energy states carry
1693: even Lorentz spin only, Eq.~\re{q-odd}.
1694:
1695: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1696: % Figure
1697: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1698: \begin{figure}[t]
1699: \begin{center}
1700: \mbox{
1701: \begin{picture}(0,230)(230,0)
1702: \put(-10,0){\insertfig{8}{rootsL10N100-10to10}} \put(-10,116){$\delta_n$}
1703: \put(110,-5){$n$} \put(180,50){$\scriptstyle{N = 10^2}$}
1704: \put(180,100){$\scriptstyle{N = 10^{10}}$}
1705: \put(250,5){\insertfig{7.65}{energyL10N100}} \put(245,116){$\varepsilon$}
1706: \put(360,-5){$N$}
1707: \end{picture}
1708: }
1709: \end{center}
1710: \caption{ \label{Fig-roots-energy10} Left panel: ``small'' roots of the transfer
1711: matrix $\delta_n$ (with $n=2,\ldots,L-1$) for $s=\ft12$, $L=10$ and two values of
1712: the spin $N = 10^2$ and $N=10^{10}$. Crosses stand for the solutions to the
1713: quantization condition \re{system-arg} and the lines correspond to \re{delta-fin}
1714: for $k_n=L+1-2n$. The roots tend to approach the line as $\xi=\ln(N/L)/L$
1715: increases from $\xi = 0.23$ to $\xi=2.07$. Right panel: the minimal energy for
1716: $s=\ft12$, $L=10$ and the total spin $0 \le N \le 100$. Crosses denote the exact
1717: values, Eq.~\re{Energy-ABA}, while the solid line stands for the semiclassical
1718: expression \re{Energy-2-cut}. We do not display the data for the asymptotic
1719: energy \re{energy-as} with roots deduced from the quantization conditions
1720: \re{system-arg}, since they are not distinguishable from the exact spectrum.}
1721: \end{figure}
1722: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1723:
1724: Let us examine the expression for the energy \re{energy-as} in the limit $\xi \gg
1725: 1$, or $\ln(N/L) \gg L $. We remind that the semiclassical expansion breaks down
1726: in this region. Equation~\re{energy-as} involves the sum over the roots of the
1727: transfer matrix \re{t-roots}. As before, we separate them into two groups.
1728: According to \re{roots-as}, the large and small roots scale as $\delta_j \sim N$
1729: and $\delta_j\sim 1/\xi$, respectively. This allows one to replace $\psi(s\pm
1730: i\delta_j)$ in \re{energy-as} by its asymptotic behavior at infinity and at the
1731: origin, respectively. In this way we obtain %for $\xi\gg 1$
1732: \be
1733: \varepsilon = 2\ln 2 - 2L \psi(2s)+ 2 \sum_{\rm large} \ln |\delta_j| +
1734: \sum_{\rm small} \left[2\psi(s) - \psi''(s)\,\delta_j^2\right] + \ldots\,,
1735: \ee
1736: where the ellipsis denotes subleading terms. The number of large roots equals $2
1737: m$, $\delta_j=-\delta_{L-j+1}$ (with $j=1,\ldots,m$) and, therefore, the leading
1738: asymptotic behavior of the energy for $\xi\gg 1$ is~\cite{BelGorKor03}
1739: \be
1740: \varepsilon = 4 \ln (\delta_1 \delta_2 \ldots \delta_m) + \mathcal{O}(N^0) = 2
1741: \ln |q_{2m}| + \mathcal{O}(N^0)\,,
1742: \ee
1743: where we took into account the relations \re{sum_rules} and $q_{2m}= \widehat
1744: q_{2m}^{\scriptscriptstyle (0)} N^{2m}$ was defined in \re{q-expression}. In this
1745: way we obtain the relation
1746: \be\label{SingleLogAsymptotics}
1747: \varepsilon = 4 m \ln N + \mathcal{O}(N^0)\,,
1748: \ee
1749: which is in an agreement with \re{E-m}. The minimal energy corresponds to $m=1$.
1750:
1751: For $m=1$ one has just two large roots $\delta_1=-\delta_{L}$ and $L-2$ small
1752: roots $\delta_n=-\delta_{L-n+1}$ (with $n=2,\ldots,L-1$) defined in
1753: Eqs.~\re{delta_1} and \re{delta-fin}, respectively. Therefore, for $\xi \gg 1$
1754: one gets
1755: \be\label{Ek}
1756: \varepsilon = 4\ln N + 2 L \left[ \psi(s)-\psi(2s)\right] + \frac{ (-\psi''(s))
1757: \,\pi^2 }{16\ln^2 N }\sum_{n=2}^{L-1}{k_n^2} +\ldots
1758: \ee
1759: This expression depends on the set of integers $k_2> k_3> \ldots$ (with
1760: $k_n=-k_{L-n+1}$) defined in \re{system-arg}. Since $\psi''(s)<0$ for $s>0$, the
1761: minimal value of $\varepsilon$ corresponds to $k_n=L+1-2n$
1762: \be\label{e-minimal}
1763: \varepsilon = 4\ln N + 2 L \left[ \psi(s)-\psi(2s)\right] + \frac{L^3
1764: (-\psi''(s)) \,\pi^2}{48\ln^2 N } +\ldots
1765: \ee
1766: This relation defines the minimal energy for $\xi \gg 1$. The energy of the
1767: excited states is described by \re{Ek} with another set of $k-$integers but of
1768: the same parity. The lowest lying excited state has the same $k$'s as the minimal
1769: energy state except $k_2=-k_{L-1}=L-1$. It is separated from the latter by the
1770: distance
1771: \be
1772: \Delta \varepsilon = \frac{ \pi^2 L (-\psi''(s))}{2\ln^2 N}\sim \frac{L}{\ln^2 N}
1773: \,.
1774: \ee
1775: This relation defines the level spacing in the spectrum of anomalous dimension of
1776: the Wilson operators \re{O-def} close to the minimal anomalous dimension
1777: trajectory.
1778:
1779: So far our discussion was limited to one loop. Remarkably enough the logarithmic
1780: behavior of the anomalous dimension \re{gamma=energy} persists to all orders of
1781: perturbation theory. Higher order corrections to \re{gamma=energy} merely modify
1782: the coefficient in front of $\ln N$ replacing $\lambda$ by an infinite series in
1783: the coupling constant,
1784: \be
1785: \gamma(\lambda) = \frac{\lambda}{8\pi^2}\, \left[4\ln N + \mathcal{O}(N^0)\right]
1786: + \mathcal{O}(\lambda^2)= 2\Gamma_{\rm cusp} (\lambda) \ln N +
1787: \mathcal{O}(N^0)\,,
1788: \ee
1789: with $\Gamma_{\rm cusp} (\lambda)$ being the cusp anomalous dimension. This
1790: relation holds true in a generic Yang-Mills theory for arbitrary values of the
1791: coupling constant $\lambda$ including the strong coupling regime. In the latter
1792: case, one can apply the gauge/string correspondence to obtain a prediction for
1793: the cusp anomalous dimension at strong coupling in the $\mathcal{N}=4$ SYM
1794: theory. In the next section, we shall trace the origin of logarithmic scaling of
1795: anomalous dimension in the dual picture of folded string rotating on the AdS
1796: space.
1797:
1798: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1799: \section{Classical strings in AdS${}_3 \times$S$^1$}
1800: \label{StringSigmaModel}
1801: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1802:
1803: Let us turn to the analysis of anomalous dimensions of composite operators
1804: \re{O-def} on the string side. According to the gauge/string correspondence
1805: the Wilson operators are mapped into certain string states whose energy is
1806: identified with the scaling dimensions of the former. It is known that in
1807: the $\mathcal{N}=4$ SYM theory, Wilson operators with the minimal anomalous
1808: dimension discussed in the preceding sections are dual to a single-folded
1809: string rotating in the AdS$_3 \times$S$^1$ sector of the target space of the
1810: type IIB string theory \cite{GubKlePol03,FroTse03}.
1811:
1812: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1813: \subsection{Folded rotating string}
1814: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1815:
1816: For the folded closed string spinning both in AdS${}_3$ and S${}^1$, the
1817: relevant bosonic part of the superstring action reads
1818: \be\label{NG}
1819: S = - \frac{\sqrt{\lambda}}{2\pi} \int d^2 \xi \sqrt{- \det\| G_{MN}(X)\,
1820: \partial_a X^M \partial_b X^N\|} \, ,
1821: \ee
1822: where the 't Hooft coupling $\lambda=g^2 N_c$ is related to the product of the
1823: radius of anti-de Sitter space $R$ and the string tension $1/\alpha^\prime$,
1824: $\sqrt{\lambda} = R^2/\alpha^\prime$, and the metric in the target space looks
1825: like
1826: \be
1827: ds^2 \equiv G_{MN} d X^M d X^N = - \cosh^2 \rho \, dt^2 + \sinh^2 \rho\,
1828: d\theta^2 + d\rho^2 + d\varphi^2 \, .
1829: \ee
1830: Here $t$, $\rho$ are $\theta$ are the global time, the radial coordinate and
1831: the angle, respectively, on the AdS${}_3$ space and $\varphi$ is the angle on
1832: a large circle of S${}^5$. The worldsheet coordinates of the string $\xi^a =
1833: (\xi^0,\xi^1)$ are chosen in such a way that
1834: \be\label{para}
1835: t = \xi^0 \, , \qquad \theta = \omega \xi^0 + \xi^1 \, , \qquad \varphi = \nu
1836: \xi^0 \, .
1837: \ee
1838: Here $\omega$ is the angular velocity on AdS and $\nu$ is a boost parameter of
1839: the center-of-mass of the folded string on S${}^5$. For the rigid folded string
1840: the radial variable $\rho$ does not depend on $\xi^0$ and is periodic in $\xi^1$,
1841: that is, $\rho (\xi^1 + 2 \pi) = \rho (\xi^1)$. The dependence $\rho=\rho(\xi^1)$
1842: is determined by the classical string equations of motion
1843: \cite{GubKlePol03,FroTse03,Kru05}. Their solution describes a folded string
1844: configuration which is sewed out of four segments running along the radial
1845: direction from the center of AdS space, $\rho=0$, towards its boundary by a
1846: distance $\rho_0$, which depends on the angular velocities entering \re{para}
1847: \be
1848: \coth^2 \rho_0 = \frac{\omega^2 - \nu^2}{1 - \nu^2}\ge 1\,.
1849: \ee
1850: The string has two spikes located at $\rho=\rho_0$ which are responsible for the
1851: logarithmic scaling of the anomalous dimension of Wilson operators at strong
1852: coupling~\cite{BelGorKor03,Kru05}.
1853:
1854: Within the gauge/string correspondence, the anomalous dimension of the operators
1855: \re{O-def} is related to the energy $E$ of the classical rotating string as
1856: \be\label{E-gamma}
1857: E = N + L + \gamma(\lambda)
1858: \, ,
1859: \ee
1860: where the Lorentz spin $N$ and twist $L$ are translated into the angular momenta
1861: on the AdS$_3$ and S$^1$ spaces, respectively. From the Nambu-Goto action
1862: \re{NG}, one finds these charges as
1863: \be\label{Efolded}
1864: E = \frac{2 \sqrt{\lambda}}{\pi} \frac{\sqrt{- \chi}}{\sqrt{1 - \nu^2}}
1865: \mathbb{E} (\chi)\,,\qquad N=\frac{2 \sqrt{\lambda}}{\pi} \frac{\omega \sqrt{-
1866: \chi}}{\sqrt{1 - \nu^2}} \left[ \mathbb{E} (\chi) - \mathbb{K} (\chi) \right] \,
1867: ,\qquad L=\frac{2 \sqrt{\lambda}}{\pi} \frac{\nu \sqrt{- \chi}}{\sqrt{1 - \nu^2}}
1868: \mathbb{K} (\chi)\,,
1869: \ee
1870: where $\mathbb{K}(\chi)$ and $\mathbb{E}(\chi)$ are the elliptic functions of
1871: the first and second kind, respectively, and the auxiliary (negative valued)
1872: parameter $\chi$ is related to the distance $\rho_0$ from the center of anti-de
1873: Sitter space to the end of the string extending to its boundary,
1874: \be\label{rho0}
1875: \chi = -\sinh^2 \rho_0 \,.
1876: \ee
1877: Excluding $\omega$ and $\nu$ in favor of $\chi$, Eqs.~\re{Efolded} can be cast
1878: into the parametric form \cite{BeiFroStaTse03}
1879: \be
1880: \label{FrolovTseytlin} \left( \frac{E}{\mathbb{E} (\chi)} \right)^2 - \left(
1881: \frac{L}{\mathbb{K} (\chi)} \right)^2 = - \frac{4\lambda}{\pi^2}\chi \, , \qquad
1882: \left( \frac{N}{\mathbb{E} (\chi) - \mathbb{K} (\chi)} \right)^2 - \left(
1883: \frac{L}{\mathbb{K} (\chi)} \right)^2 = \frac{4\lambda}{\pi^2} ( 1 - \chi ) \, .
1884: \ee
1885: Using these relations, one can analyze the anomalous dimension $\gamma=E-N-L$ at
1886: strong coupling in the semiclassical limit of large angular momenta $L$, $N \to
1887: \infty$ with $N/\sqrt{\lambda}$ and $L/\sqrt{\lambda}$ kept fixed. In this way
1888: one finds from \re{FrolovTseytlin} that the two limiting cases $N \ll L$ and $N
1889: \gg L$ correspond to the short and long strings, $\rho_0\ll 1$ and $\rho_0\gg 1$,
1890: respectively.
1891:
1892: It is well-known~\cite{GubKlePol03,FroTse03} that in the short string limit,
1893: $\rho_0\ll 1$, or equivalently $(-\chi)\ll 1$, the anomalous dimension exhibits
1894: the BMN scaling
1895: \be
1896: \label{AnomalousDimensionPertExp} \gamma (\lambda) = L\left[ \lambda'
1897: \gamma^{(0)}+ {(\lambda')}^2\gamma^{(1)}+ \ldots \right] \, ,
1898: \ee
1899: with $\gamma^{(0)}, \gamma^{(1)},\ldots$ being functions of the ratio $N/L$ and
1900: $\lambda'=\lambda/(\pi L)^2$ determining the BMN coupling constant. Substituting
1901: $\chi = \chi_0 + \lambda^\prime \chi_1 + \dots$ into \re{FrolovTseytlin} and
1902: matching the coefficients in the expansion of both sides of \re{FrolovTseytlin}
1903: in powers of $\lambda'$, one finds for $N/L \ll
1904: 1$~\cite{BerMalNas02,FroTse03,BeiFroStaTse03}
1905: \be
1906: \gamma^{(0)}=\frac{\pi^2}{2}\frac{N}{L}+\ldots \,,\qquad
1907: \gamma^{(1)}=-\frac{\pi^4}{8}\frac{N}{L}+\ldots\,.
1908: \ee
1909: The lowest order term in \re{AnomalousDimensionPertExp} matches the one-loop
1910: result on the gauge theory side, Eq.~\re{E-small}, for $s=\ft12$ and $m=1$
1911: \be\label{match}
1912: \gamma^{(0)} = \frac{L}{8} \varepsilon\,.
1913: \ee
1914: In a similar manner, in the long string limit, $\rho_0 \to \infty$, or
1915: equivalently $(-\chi)\to\infty$, assuming that $\chi$ has a regular expansion in
1916: powers of $\lambda'$, one finds from Eq.\ \re{FrolovTseytlin} that the leading
1917: order parameter $\chi_0$ scales for $N/L\gg 1$ as
1918: \be
1919: \chi_0 = - \frac{N}{2 L} \ln \frac{N}{L} \, .
1920: \ee
1921: This leads to the following expressions for the coefficients of the BMN expansion
1922: of the anomalous dimension \re{AnomalousDimensionPertExp} for $N/L\gg 1$
1923: \ba
1924: \label{AsyPertExp} \gamma^{(0)} \!\!\!&=&\!\!\! \ft12 \ln^2 (- \chi_0)+\ldots
1925: =\ft12 \ln^2 (N/L)+\ldots\, , \\ \nonumber \gamma^{(1)} \!\!\!&=&\!\!\!
1926: -\ft{1}{8} \ln^4 (- \chi_0) +\ldots = -\ft18 \ln^4 (N/L)+\ldots\, ,
1927: \ea
1928: where ellipses denote terms subleading for $(-\chi_0)\to\infty$. In agreement
1929: with \re{match}, the relation \re{AsyPertExp} matches the one-loop expression for
1930: the anomalous dimension on the gauge theory side, Eq.~\re{E-large}. Moreover, one
1931: can show~\cite{BeiFroStaTse03} that the relation \re{match} (upon the Landen
1932: transformation defined below in Eq.\ \re{LandenGauss}) holds for an arbitrary
1933: value of $N/L$ with $\gamma^{(0)}$ given by \re{FrolovTseytlin} and
1934: \re{AnomalousDimensionPertExp} and $\varepsilon$ defined in \re{Energy-2-cut} and
1935: \re{ab}.
1936:
1937: So far we see no trace of expected scaling of the anomalous dimension $\gamma
1938: (\lambda) \sim \ln N$ for $N\gg L$. We recall that on the gauge theory side the
1939: semiclassical expansion for the one-loop energy $\varepsilon$ was divergent for
1940: $N\gg L$ and in order to recover the logarithmic scaling $\varepsilon\sim \ln N$
1941: we had to resum the entire semiclassical series in powers of the parameter
1942: $\xi=\ln(N/L)/L$. It turns out that a similar phenomenon happens for the series
1943: \re{AnomalousDimensionPertExp} on the string side although the parameter of the
1944: semiclassical expansion is different and equals
1945: \be\label{xi-str}
1946: \xi_{\rm str}=\frac{\lambda}{L^2} \ln^2\frac{N}{L}=\lambda\, \xi^2\,.
1947: \ee
1948: It can be easily identified by comparing the contribution to
1949: \re{AnomalousDimensionPertExp} from $\gamma^{(0)}$ and $\gamma^{(1)}$,
1950: Eq.~\re{AsyPertExp}.
1951:
1952: To sum up the infinite series in \re{AnomalousDimensionPertExp} for $\xi_{\rm
1953: str}\gg 1$ we return to \re{FrolovTseytlin} and rewrite the first relation as
1954: \be\label{gamma-sqrt}
1955: \gamma(\lambda) = L\left[\sqrt{1- {4\lambda'} {\chi}{\mathbb{K}^2(\chi)}}-1+
1956: \Delta\gamma\right]
1957: \ee
1958: with $\lambda'=\lambda/(\pi L)^2$ and
1959: \be
1960: \Delta\gamma=-
1961: \frac{{4\lambda'}\,\mathbb{K}(\chi)[\mathbb{E}(\chi)-\mathbb{K}(\chi)]}{\sqrt{1-{4\lambda'}
1962: \chi {\mathbb{K}^2(\chi)} }+ \sqrt{1-{4\lambda'} (\chi-1){\mathbb{K}^2(\chi)}
1963: }}\,.
1964: \ee
1965: In the limit of a long string, for $(-\chi)\to\infty$, the expression for
1966: $\gamma(\lambda)$ can be expanded in powers of $4\lambda' (-\chi)
1967: \mathbb{K}^2(\chi)\sim \lambda'\ln^2(-\chi)$. Examining the expression for
1968: $\Delta\gamma$ one finds that its contribution to $\gamma(\lambda)$ is suppressed
1969: by the factor $[\mathbb{E}(\chi)-\mathbb{K}(\chi)]/(-\chi\mathbb{K}(\chi))\sim
1970: 1/\ln(-\chi)$ compared to the first term in the square brackets in
1971: \re{gamma-sqrt}. This implies that in the limit $(-\chi)\to\infty$ with
1972: $\lambda'\ln^2(-\chi)= {\rm fixed}$ the leading asymptotic behavior of
1973: \re{gamma-sqrt} reads
1974: \be\label{for}
1975: \gamma(\lambda) = L\left[\sqrt{1+{\lambda'} \ln^2(-\chi)}-1\right]+\ldots\,.
1976: \ee
1977: This relation resums all double-logarithmic corrections $\sim L[{\lambda'}
1978: \ln^2(-\chi)]^n$ to the anomalous dimension $\gamma(\lambda)$ to all orders $n$
1979: at strong coupling. In particular, for $n=1$ and $n=2$ it reproduces
1980: \re{AsyPertExp}. For ${\lambda'} \ln^2(-\chi)\ll 1$ one expands the square-root
1981: in \re{for} and arrives at a BMN-like expansion \re{AnomalousDimensionPertExp}.
1982: At the same time, for ${\lambda'} \ln^2(-\chi)\gg 1$ the relation \re{for} leads
1983: to the expression
1984: \be
1985: \gamma(\lambda) = L \sqrt{\lambda'} \ln (-\chi) + \ldots =
1986: \frac{\sqrt{\lambda}}{\pi} \ln(-\chi) + \ldots \,.
1987: \ee
1988: For $(-\chi)\to\infty$, the dependence of $\chi$ on the ratio $N/L$ and the BMN
1989: coupling $\lambda'$ follows from the second relation in \re{FrolovTseytlin}
1990: \be\label{xi-eq}
1991: \frac{1}{4\chi^2} =\frac{L^2}{N^2} \left[ \frac{1}{\ln^2 (- \chi)} +
1992: \lambda^\prime \right]\, .
1993: \ee
1994: For $\xi_{\rm srt}< 1$ and $N\gg L$ one finds from \re{xi-eq} and \re{xi-str}
1995: that $\chi\sim\chi_0 = - \frac{N}{2 L} \ln \frac{N}{L}$ and, therefore,
1996: \be\label{gamma-str-dlog}
1997: \gamma(\lambda)= L\left[\sqrt{1+\frac{\lambda}{(\pi L)^2} \ln^2(N/L)}-1\right] +
1998: \ldots
1999: \ee
2000: For $\xi_{\rm srt}\gg 1$ one finds from \re{xi-eq} and \re{xi-str} that
2001: $(-\chi)\sim N/(2L\sqrt{\lambda'})\sim N/\sqrt{\lambda}$ and, therefore,
2002: \be\label{gamma-str-log}
2003: \gamma(\lambda)=\frac{\sqrt{\lambda}}{\pi} \ln(N/\sqrt{\lambda}) + \ldots
2004: \ee
2005: This relation reproduces the correct asymptotic behavior of the anomalous
2006: dimension in the regime \re{3rd} and it is in agreement with the results
2007: of Refs.~\cite{GubKlePol03,FroTse03,Kru05}.
2008:
2009: The coefficient in front of $\ln N$ in \re{gamma-str-log} determines the leading
2010: asymptotics of the cusp anomalous dimension in the strong coupling regime in the
2011: $\mathcal{N}=4$ SYM theory, Eqs.~\re{gamma=cusp} and \re{cusp}. It follows from
2012: the relations \re{gamma-str-dlog} and \re{gamma-str-log} that quite remarkably
2013: the cusp anomaly at strong coupling can be obtained by resumming double
2014: logarithmic terms $\sim L [\lambda\ln^2(N/L)/L^2]^k$ in the expansion of
2015: anomalous dimensions of operators of higher twist $L$ and large Lorentz spin
2016: $N\gg L$.
2017:
2018: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2019: \subsection{Two-cut solution in string sigma model}
2020: \label{StringTwoCut}
2021: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2022:
2023: In the previous section, we observed that for $N\gg L$ the dependence of the
2024: anomalous dimension $\gamma (\lambda)$ on the coupling constant is different for
2025: $\xi_{\rm str}< 1$ and $\xi_{\rm str}\gg 1$. In the former case, $\gamma
2026: (\lambda)$ has a BMN-like expansion in powers of $\lambda/L^2$, while in the
2027: latter case $\gamma(\lambda)$ is not analytical in $\lambda$ and scales
2028: logarithmically $\sim \sqrt{\lambda}\ln N$. The reason for this non-analyticity
2029: is that for $\xi_{\rm str}\gg 1$ the end-points of the rotating string approach
2030: the boundary of the AdS space and the dominant contribution to the energy of the
2031: string, or equivalently the anomalous dimension \re{E-gamma}, comes from the
2032: vicinity of two spikes, $\gamma(\lambda) \sim (\sqrt{\lambda}/\pi) \cdot 2
2033: \rho_0$. Here the radial coordinate of the spikes scales for $\xi_{\rm str}\gg 1$
2034: as $\rho_0\sim \ft12\ln (-\chi)\sim \ft12\ln(N/\sqrt{\lambda})$, Eqs.~\re{rho0}
2035: and \re{xi-eq}. The phenomenon is rather general~\cite{BelGorKor03,Kru05} and it
2036: holds true for classical string configurations with an arbitrary number of spikes
2037: $n$. In that case, each spike provides a logarithmic contribution to the energy
2038: and the anomalous dimension scales for $\xi_{\rm str}\gg 1$ as $\gamma(\lambda)=
2039: {n\sqrt{\lambda}}/{(2\pi)} \ln(N/\sqrt{\lambda})$. At first glance, this
2040: mechanism is rather different from the one in gauge theory. We recall that in
2041: gauge theory, to one-loop order, the logarithmic scaling of anomalous dimension
2042: for $N\gg L$ arises due to collision of cuts for the spectral curve of the spin
2043: chain, Eq.~\re{curve}. It is known that the classical equations of motion for the
2044: string on the AdS${}_5\times$S${}^5$ background are completely
2045: integrable~\cite{ZakMik78,ManSurWad02} and their solutions are parameterized by
2046: the spectral curves. Moreover, for the strings on the AdS${}_3\times$S${}^1$ part
2047: of the target space the spectral curve can be identified as a complex
2048: hyperelliptic curve~\cite{Kri94}. For the folded rotating closed string discussed
2049: in the previous section, it is given by the elliptic curve with symmetric
2050: branching points on the real axis~\cite{KazZar04}
2051: \be\label{curve-str}
2052: \Gamma_{\rm str}: \qquad y^2=(x^2-a_{\rm str}^2)(x^2-b_{\rm str}^2)\,,
2053: \ee
2054: with $a_{\rm str}$ and $b_{\rm str}$ taking positive values, $b_{\rm str}< a_{\rm
2055: str}$. In this section, we shall translate different asymptotic behavior of the
2056: anomalous dimensions for $N\gg L$, Eqs.~\re{gamma-str-dlog} and
2057: \re{gamma-str-log}, into properties of the curve \re{curve-str} and reveal the
2058: mechanism responsible for the logarithmic scaling \re{gamma-str-log}.
2059:
2060: Similar to the classical $SL(2)$ spin chain, the classical string equations of
2061: motion admit the Lax representation and they can be solved exactly by
2062: constructing the Baker-Akhiezer function~\cite{Kri94,ZakMik78}. As before, the
2063: Bloch-Floquet multiplier for this function gives rise to the (quasi)momentum
2064: $p(x)$ which is the generating function for the conserved charge including the
2065: energy. For the folded rotating string configuration, $p'(x)$ is an analytical
2066: function in the complex plane with two symmetric cuts $[-a_{\rm str},-b_{\rm
2067: str}]\cup[b_{\rm str}, a_{\rm str}]$. It is uniquely fixed by the requirement
2068: that $dp=p'(x)dx$ should be a meromorphic differential on the complex curve
2069: \re{curve-str} satisfying the following conditions on the upper sheet of
2070: $\Gamma_{\rm str}$~\cite{KazZar04}:
2071: \begin{itemize}
2072:
2073: \item Prescribed asymptotics at infinity and at the origin
2074: \be\label{p-as}
2075: dp \stackrel{x\to\infty}{\sim} -2\frac{dx}{x^2} \frac{E+N}{L}\,,\qquad dp
2076: \stackrel{x\to 0}{\sim} -dx\,\frac{2}{\lambda'} \frac{E-N}{L}
2077: \ee
2078:
2079: \item Single-valuedness condition
2080: \be\label{p-periodsSigma}
2081: \int_{b_{\rm str}}^{a_{\rm str}} dp = 0\,,\qquad \int_{a_{\rm str}}^{\infty} dp =
2082: -\pi m\,,
2083: \ee
2084:
2085: \item Double poles at $x=\pm\sqrt{\lambda}/(\pi L)\equiv \sqrt{\lambda'}$
2086: \be\label{poles}
2087: dp \sim dx
2088: \left[
2089: - \frac{1}{(x\pm \sqrt{\lambda'})^2}
2090: +
2091: \mathcal{O} ((x\pm \sqrt{\lambda'})^0) \right]
2092: \, .
2093: \ee
2094:
2095: \end{itemize}
2096: The resulting expression for the differential $dp$ takes the form
2097: \be\label{p-ans}
2098: dp = \frac{dx}{y}
2099: \left[
2100: \frac{y_+}{(x-\sqrt{\lambda'})^2}+\frac{y_+'}{x-\sqrt{\lambda'}}
2101: +
2102: \frac{y_+}{(x+\sqrt{\lambda'})^2}-\frac{y_+'}{x+\sqrt{\lambda'}}
2103: +
2104: C
2105: \right]\,,
2106: \ee
2107: where $y=y(x)$ is defined in \re{curve-str}, $y_+ = y(\sqrt{\lambda'})$ and
2108: $y_+'=y'(\sqrt{\lambda'})$.
2109:
2110: Equation \re{p-ans} depends on three parameters ${a_{\rm str}}$, ${b_{\rm str}}$ and
2111: $C$. They are fixed by the normalization conditions \re{p-as} and \re{p-periodsSigma}
2112: as
2113: \ba\label{b-str}
2114: b_{\rm str}
2115: \!\!\!&=&\!\!\!
2116: \frac{1}{m\mathbb{K}(\tau)} \left[
2117: \left(1-\frac{\lambda^\prime}{a_{\rm str}^2}\right)
2118: \left(1 - \frac{\lambda^\prime}{b_{\rm str}^2}\right)
2119: \right]^{-1/2}
2120: \\[3mm]
2121: C
2122: \!\!\!&=&\!\!\!
2123: -\frac{m a_{\rm str}}{2}
2124: \left[
2125: \mathbb{E}(\tau) -\frac{\lambda^\prime}{a_{\rm str}^2} \mathbb{K}(\tau) \right]
2126: \, , \nonumber
2127: \ea
2128: with the modular parameter $\tau = 1-{b_{\rm str}^2}/{a_{\rm str}^2}$. In
2129: addition, one finds from \re{p-as} the following expressions for the ratio
2130: $N/L$ and for the anomalous dimension $\gamma(\lambda)=E-N-L$
2131: \ba\label{N/L}
2132: {N}/{L} \!\!\!&=&\!\!\!
2133: \frac{m}{2}
2134: \left[
2135: \mathbb{E}(\tau)\lr{ a_{\rm str} + \frac{{\lambda'}}{b_{\rm str}} }
2136: -
2137: \mathbb{K}(\tau) \lr{ b_{\rm str} + \frac{{\lambda'}}{a_{\rm str}} }
2138: \right]
2139: \, , \\[3mm]
2140: \gamma(\lambda)/L \!\!\!&=&\!\!\!
2141: m \left[ \mathbb{K}(\tau)\, b_{\rm str} - \mathbb{E}(\tau) \frac{{\lambda'}}{b_{\rm str}} \right]
2142: - 1
2143: \, , \nonumber
2144: \ea
2145: with $b_{\rm str}$ defined in \re{b-str} and $a_{\rm str}=b_{\rm str}/\sqrt{1-\tau}$.
2146:
2147: Let us examine the dependence of the anomalous dimension on the coupling constant
2148: $\lambda'=\lambda/(\pi L)^2$. Assuming that $b_{\rm str}$ and $\tau$ both admit a
2149: regular expansion in powers of $\lambda'$, one substitutes into \re{b-str} and
2150: \re{N/L}
2151: \be\label{b-exp}
2152: b_{\rm str} = b_{\rm str}^{(0)} + \lambda^\prime \, b_{\rm str}^{(1)} + + \ldots
2153: \, , \qquad \tau = \tau^{(0)} + \lambda^\prime \,\tau^{(1)} + \ldots \, ,
2154: \ee
2155: and matches the coefficients in front of powers of $\lambda'$. In this way, one
2156: obtains to leading order
2157: \be\label{b0}
2158: b_{\rm str}^{(0)} = \frac{1}{m\mathbb{K}(\tau^{(0)})}
2159: \,,\qquad
2160: \frac{N}{L} =
2161: \frac12\left[\frac{\mathbb{E}(\tau^{(0)})}{\sqrt{1-\tau^{(0)}}\mathbb{K}(\tau^{(0)})}
2162: -1\right]\,,
2163: \ee
2164: and all subleading corrections to \re{b-exp} are expressed in terms of the leading
2165: terms. Then, the first few coefficients of the BMN series for anomalous dimension
2166: \re{AnomalousDimensionPertExp} are given by
2167: \ba
2168: \label{gamma1Diff} \gamma^{(0)} \!\!\!&=&\!\!\! \frac{m^2}2 \mathbb{K}
2169: (\tau^{(0)}) \left[ (2 - \tau^{(0)}) \mathbb{K} (\tau^{(0)}) - 2 \mathbb{E}
2170: (\tau^{(0)}) \right] \, ,
2171: \\
2172: \label{gamma2Diff} \gamma^{(1)} \!\!\!&=&\!\!\! \frac{m^4}{8} \mathbb{K}^{3}
2173: (\tau^{(0)}) \left[ \lr{ 4 (2 - \tau^{(0)}) \sqrt{1 - \tau^{(0)}} - (\tau^{(0)})^2}
2174: \mathbb{K} (\tau^{(0)}) - 8 \sqrt{1 - \tau^{(0)}} \mathbb{E} (\tau^{(0)}) \right]
2175: \, .
2176: \ea
2177: Together with the second relation in \re{b0}, they define the parametric dependence
2178: of the anomalous dimension \re{AnomalousDimensionPertExp} on $N/L$.
2179:
2180: Equations \re{b0} determine perturbative corrections to the anomalous dimension of
2181: long scalar operators in the $\mathcal{N}=4$ SYM theory. According to \re{match},
2182: it should match at one-loop order a similar asymptotic expression \re{Energy-2-cut}
2183: obtained on the gauge theory side within the semiclassical approach. The conformal
2184: spin of scalars equals $s=1/2$ and the scaling parameter $\beta=L/(L+2N)$, Eq.\
2185: \re{beta}, is given by $\beta = \sqrt{1 -\tau^{(0)}}\,{\mathbb{K}(\tau^{(0)})}/
2186: {\mathbb{E}(\tau^{(0)})}$ in agreement with \re{ab}. Then, one observes that
2187: $\gamma^{(0)}$ and $\varepsilon$ given by \re{gamma1Diff} and \re{Energy-2-cut},
2188: respectively, verify the relation \re{match}. The expressions for the branching
2189: points $b_{\rm str}^{(0)}$ and $b$, defined in \re{b0} and \re{ab}, respectively,
2190: are different, $b/b_{\rm str}^{(0)} = \beta/2$, but the agreement can be restored
2191: through the rescaling of the local complex parameter $x$ in the definition of the
2192: curve \re{curve-str}, $x \to x \beta/2$.
2193:
2194: The functional form of the anomalous dimensions \re{N/L} is different compared to
2195: the ones found in the previous section, Eq.~\re{FrolovTseytlin}. The agreement is
2196: achieved by means of the Landen transformation of the modular parameters
2197: \cite{BeiFroStaTse03}
2198: \be
2199: \label{LandenGauss}
2200: \chi^{(0)} = - \frac{\left( 1 - \sqrt{1 - \tau^{(0)}} \right)^2}{4 \sqrt{1 -
2201: \tau^{(0)}}} \, ,
2202: \ee
2203: upon which the relations \re{N/L} and \re{FrolovTseytlin} coincide provided that
2204: $m=1$. Remember that $(-\chi)$ depends on the radial coordinate of the spike
2205: $\rho_0$, Eq.~\re{rho0}, so that $\sqrt{1-\tau}=\e^{-2\rho_0}$ and the long
2206: string limit $\rho_0\to\infty$ corresponds to $\tau\to 1$. We have demonstrated
2207: in Sect.~\ref{CollisionCuts} that on the gauge theory side the limit $\tau\to 1$
2208: corresponds to $a\to 1/(2m)$ and $b\to 0$, Eq.~\re{ab-large}. As a consequence,
2209: the two cuts $[-a,-b]$ and $[b,a]$ collide at the origin yielding the logarithmic
2210: scaling of the one-loop anomalous dimension \re{E-m}. Let us examine the limit
2211: $\tau\to 1$ of the obtained stringy expressions \re{b-str} and \re{N/L}.
2212:
2213: Since $\mathbb{K}(\tau)\sim -\ft12 \ln(1-\tau)$ for $\tau\to 1$, one would expect
2214: from \re{b-str} that $b_{\rm str}$ should vanish in this limit. Indeed, this is
2215: the case for $\lambda'=0$ while for $\lambda'\neq 0$ one deduces from \re{b-str}
2216: that the reality condition for $b_{\rm str}$ implies that its minimal value is
2217: bounded as $b_{\rm str}\ge \sqrt{\lambda'}$. Carefully examining \re{b-str} for
2218: $\tau\to 1$ one finds
2219: \be\label{barrier}
2220: b_{\rm str}=\sqrt{{\lambda'}+\frac{1}{m^2\ln^2 \sqrt{1-\tau}}}\,.
2221: \ee
2222: For $\lambda'=0$ this relation coincides with the one-loop expression
2223: $b^{(0)}_{\rm str}=2 b/\beta$, Eq.~\re{ab-large}. Matching \re{barrier} into
2224: \re{b-exp} we conclude that higher order corrections to $b_{\rm str}$ push its
2225: minimal possible value away from the origin and, therefore, prevent the two cuts
2226: $[-a_{\rm str},-b_{\rm str}]$ and $[b_{\rm str},a_{\rm str}]$ to collide. From
2227: \re{N/L} one finds in the limit $\tau\to 1$
2228: \be\label{gamma-str}
2229: \beta\approx\frac{L}{2N}
2230: =
2231: \frac{\sqrt{1-\tau}}{m\, b_{\rm str}}
2232: + \ldots
2233: \,,\qquad
2234: \gamma(\lambda) = L \left[- m \, b_{\rm str} \, \ln{\sqrt{1-\tau}} - 1\right]
2235: +
2236: \ldots
2237: \, .
2238: \ee
2239: The asymptotic behavior of these expressions depends on the value of the
2240: parameter $\xi_{\rm str}=\lambda'\ln^2(N/L)$, Eq.~\re{xi-str}.
2241:
2242: For $\xi_{\rm str}< 1$ the expression for $b_{\rm str}$, Eq.~\re{barrier}, admits
2243: a series expansion in $\lambda'$ and leads together with \re{gamma-str} to
2244: \be\label{ab-str}
2245: b_{\rm str}=\frac{1}{m\ln\frac{N}{L}} + \ldots \,,\qquad a_{\rm
2246: str}=\frac{N}{2m{L}} + \ldots \,,\qquad \sqrt{1-\tau}=\frac{L/N}{2 \ln
2247: \frac{N}{L}} + \ldots \,.
2248: \ee
2249: Substitution of \re{barrier} into \re{gamma-str} yields an expression for the
2250: anomalous dimension $\gamma(\lambda)$ which coincides with \re{gamma-str-dlog}
2251: for $m=1$. It is instructive to compare the positions of the cuts in gauge
2252: theory, $a$ and $b$, and on the string side, $\widehat a=\beta a_{\rm str}/2$ and
2253: $\widehat b=\beta b_{\rm str}/2$. Here the additional factor $\beta/2$ appears
2254: due to a different definition of the local parameter $x$ on the spectral curves
2255: \re{curve-reduced} and \re{curve-str}. In this way we find $ \widehat a
2256: =1/{(2m)}+\ldots$ and $\widehat b={(L/N)}/{(4m\ln(N/L))}+\ldots$ which coincides
2257: with the similar relation \re{ab-large} in gauge theory to one-loop order.
2258:
2259: For $\xi_{\rm str}\gg 1$ the expression for $b_{\rm str}$, Eq.~\re{barrier}, is
2260: not analytical in the BMN coupling $\lambda'=\lambda/(\pi L)^2$
2261: \be
2262: b_{\rm str}=\sqrt{\lambda'} + \ldots
2263: \,,\qquad
2264: a_{\rm
2265: str}=\frac{N}{2m{L}} + \ldots \,,\qquad \sqrt{1-\tau}=\frac{Lm}{2N}
2266: \sqrt{\lambda'} + \ldots \,,
2267: \ee
2268: while the expression for $a_{\rm str}$ is the same as in \re{ab-str}. This
2269: suggests that for $\xi_{\rm str}\gg 1$ higher order corrections only modify the
2270: lower edge of the cut. One finds from \re{gamma-str} that the anomalous dimension
2271: scales as $\gamma(\lambda)\sim m\sqrt{\lambda}\ln (N/(m\sqrt{\lambda}))$ and
2272: matches \re{gamma-str-log} for $m=1$. We conclude that the logarithmic scaling of
2273: the anomalous dimension for $N\gg L$ on the string side is realized when the
2274: inner boundary of the cut $b_{\rm str}$ approaches its minimal possible value
2275: $\sqrt{\lambda'}$ (see Fig.~\ref{CollisionCutsFig}c) which coincides with the
2276: position of the double pole of the differential \re{poles}.
2277:
2278: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2279: \section{Conclusion}
2280: \label{ConclusionSection}
2281: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2282:
2283: In the present paper we have studied the properties of anomalous dimensions
2284: of Wilson operators of higher twist $L$ and large Lorentz spin $N$ in the
2285: weak and strong coupling regimes by making use of the remarkable integrability
2286: symmetry on both sides of the gauge/string correspondence. We concentrated on
2287: operators which have the minimal anomalous dimension for given Lorentz spin
2288: and put a special emphasis on the appearance of the single-logarithmic behavior,
2289: $\gamma(\lambda) \sim \ln N$, in the thermodynamic limit $L\to\infty$.
2290:
2291: On the gauge theory side, we applied the method of the Baxter $Q-$operator to
2292: identify different regimes of the minimal anomalous dimension in integrable
2293: sectors of (supersymmetric) Yang-Mills theory to one-loop order. We argued that
2294: for $N \gg L$ the asymptotic behavior of $\gamma(\lambda)$ is controlled by the
2295: parameter $\xi=\ln(N/L)/L$. For $\xi < 1$ the anomalous dimension possesses the
2296: BMN scaling $\gamma(\lambda)\sim \lambda/L$, while for $\xi \gg 1$ it scales
2297: logarithmically with $N$. Transition to the second, logarithmic regime manifests
2298: itself through the divergence of the semiclassical expansion for $\gamma(\lambda)$
2299: as $\xi \sim 1$. The anomalous dimension is uniquely determined by the configuration
2300: of Bethe roots which condense in the thermodynamic limit on two symmetric cuts. We
2301: demonstrated that the semiclassical approach breaks down for $\xi\sim 1$ due to
2302: the collision of cuts at the origin and worked out an asymptotic expression for
2303: anomalous dimensions which is valid throughout the entire region of $\xi$.
2304:
2305: On the string theory side, we used the identification of the minimal anomalous
2306: dimension of scalar operators in the $\mathcal{N}=4$ SYM at strong coupling as
2307: the energy of folded string rotating on AdS${}_3\times$S${}^1$ part of the target
2308: space. Similar to the previous case, the anomalous dimension has different
2309: behavior for $N\gg L$ depending on the value of the parameter $\xi_{\rm str} =
2310: \lambda \ln^2(N/L)/L^2$. For $\xi_{\rm str}<1$ the anomalous dimension has a
2311: regular expansion in powers of the BMN coupling and its lowest term matches the
2312: one-loop expression for $\gamma(\lambda)$ at weak coupling for $\xi<1$. For
2313: $\xi_{\rm str}\gg 1$ the anomalous dimension scales logarithmically but its
2314: dependence on the 't Hooft coupling is not analytical anymore. We described the
2315: latter regime using two different (although equivalent) configurations. In terms
2316: of the folded rotating string the logarithmic scaling occurs when two most
2317: distant points of the string (two spikes) approach the boundary of the AdS space.
2318: In terms of the spectral curve for the classical string sigma model, the same
2319: configuration is described by the elliptic curve with symmetric branching points.
2320: Different regimes of $\gamma(\lambda)$ arise depending on the position of the
2321: branching points. In the logarithmic regime, the inner branching points approach
2322: the minimal possible value $\pm \sqrt{\lambda^\prime}$ so that the anomalous
2323: dimension ceases to obey the BMN scaling.
2324:
2325: Integrability played a key role in our analysis. In generic (supersymmetric)
2326: Yang-Mills theory it holds to one-loop order for Wilson operators belonging to
2327: special, holomorphic sectors only. We would like to stress that logarithmic
2328: behavior of anomalous dimensions is not tied to integrability. In non-integrable
2329: sectors the mixing matrix for Wilson operators contains additional terms which
2330: break integrability symmetry. They do not affect however the logarithmic scaling
2331: of anomalous dimension for $N\to\infty$. The reason for this is that the
2332: logarithmic scaling of anomalous dimension can be associated with the
2333: contribution of soft gluons (i.e., gauge field quanta). Soft gluon radiation is
2334: not sensitive to the quantum numbers of Wilson operators (except of the total
2335: Lorentz spin) and, therefore, it provides the same logarithmic contribution to
2336: the anomalous dimension of Wilson operators in all sectors. Integrability allows
2337: one to identify various regimes of the asymptotic behavior of anomalous
2338: dimensions and to determine the ``critical'' value of $\ln(N/L) \sim L $ at which
2339: the logarithmic scaling sets in.
2340:
2341: As a function of the total Lorentz spin $N$, the anomalous dimensions of
2342: twist$-L$ operators occupy a band. Our discussion was restricted to the minimal
2343: anomalous dimensions belonging to the lower edge of the band. It would be
2344: interesting to extend the above analysis to excited states
2345: \re{SingleLogAsymptotics} with $m > 1$ and describe the band structure occupied
2346: by the anomalous dimensions as it arises form the Baxter equation on the gauge
2347: theory side. In string theory, there are two different classical configurations
2348: yielding the same logarithmic asymptotics of anomalous dimensions at strong
2349: coupling --- the multiple folded string \cite{FroTse03} and the spiky string
2350: \cite{Kru05}. The coefficient in front of the logarithm, $m$, is twice the number
2351: of foldings of the string on itself, in the former case and it is the number of
2352: spikes, in the latter one. It would be interesting to construct a generic
2353: configuration which interpolates between both solutions.
2354:
2355: \vspace{0.5cm}
2356:
2357: \noindent {\bf Note added}: After the work has been completed we were informed by
2358: Yuji Satoh that he, in collaboration with Kazuhiro Sakai, analyzed the spectrum
2359: of anomalous dimensions with a special emphasis on the appearance of logarithmic
2360: scaling in the large spin limit. For the two-cut solution, their findings are in
2361: agreement with the analysis presented in sections \ref{WKBloSection} and
2362: \ref{StringTwoCut} of this paper. In particular, they also came to the conclusion
2363: that, in the thermodynamical limit $L\to\infty$, the semiclassical approach to
2364: the Bethe Ansatz equations is only applicable for $\ln (N/L) < L$ and one can not
2365: reproduce $\sim \ln N$ behavior of anomalous dimensions unless the finite size
2366: corrections in $1/L$ are included.
2367:
2368: \vspace{0.5cm}
2369:
2370: We would like to thank to thank S.~Derkachov, Yu.~Makeenko, A.~Manashov, Y.~Satoh,
2371: F.~Smirnov, E.~Sokatchev, B.~Stefanski and A.~Tseytlin for very useful discussions
2372: and correspondence. A.G.\ is grateful to Laboratoire de Physique Th\'eorique (Orsay)
2373: for hospitality extended to him during his stay which was partially supported by
2374: the Russian-French Exchange Program. This work was supported by the U.S.\ National
2375: Science Foundation under grant no.\ PHY-0456520 (A.B.), by the grant CRDF-RUP2-261-MO-04
2376: from the U.S.\ Civilian Research and Development Foundation and Russian Foundation
2377: for Basic Research under contract RFBR-04-011-00646 (A.G.).
2378:
2379: \vspace{0.5cm}
2380:
2381: \appendix
2382: \setcounter{section}{0}
2383: \setcounter{equation}{0}
2384: \renewcommand{\theequation}{A.\arabic{equation}}
2385:
2386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2387: \section*{Appendix \ \ Asymptotic expression for the energy}
2388: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2389:
2390: The asymptotic expression for the energy \re{energy-as} involves roots of the
2391: transfer matrix and is not well suited for performing the thermodynamical limit
2392: $L \to \infty$. Let us rewrite the energy in terms of the transfer matrix itself.
2393: We notice from \re{t-roots} that
2394: \be
2395: \frac{d}{du} \ln t_L(u) = \sum_{k=1}^L \frac1{u-\delta_k}
2396: \ee
2397: and replace the $\psi-$functions in \re{energy-as} by series representation
2398: \be
2399: \psi(s+i\delta_k) + \psi(s-i\delta_k) -2 \psi(2s) = \sum_{k=0}^\infty
2400: \left[\frac{i}{i(s+k)-\delta_k}+\frac{-i}{-i(s+k)-\delta_k} -\frac{2}{2s+k}
2401: \right]\,.
2402: \ee
2403: Then, the energy \re{energy-as} can be written as an infinite sum
2404: \be\label{E-main}
2405: E = 2\ln 2-\sum_{k=0}^\infty \frac{d}{dk}f(k)\,,\qquad
2406: %
2407: f(x)=\ln \left[ \frac{t_L(i(s+x))t_L(-i(s+x))}{(2s+x)^{2L}} \right]\,,
2408: \ee
2409: where $f(x)\to 2\ln 2$ for $x\to\infty$. The sum can be evaluated with a help of
2410: the Euler-Maclaurin summation formula
2411: \be\label{E-ex}
2412: E = f(0)-\frac1{2} f'(0)+\sum_{k=1}^\infty \frac{B_{2k}}{(2k)!}f^{(2k)}(0) =
2413: \frac{\textrm{d}}{\e^\textrm{d}-1} f(0)
2414: \ee
2415: with $B_{2k}$ being Bernoulli numbers. Taking $u=\pm i(s+x)$ in the Baxter
2416: equation \re{Baxter-eq}, one obtains $t_L(\pm i(s+x))$ as the ratio of
2417: $Q-$functions and finds for $x\to 0$
2418: \be
2419: f(x)=\ln \left[\frac{Q(i(s+x+1))}{Q(i(s+x))}
2420: \frac{Q(-i(s+x+1))}{Q(-i(s+x))}\right] + \mathcal{O}\lr{ x^{L}}\,.
2421: \ee
2422: Then, one uses the WKB ansatz $Q(u) \sim \exp(\eta^{-1}S(u\eta))$ to get
2423: \be
2424: f(x)= \frac1{\eta}\big[S\lr{i{(s+x+1)}{\eta}}-S\lr{i{(s+x)}{\eta}}
2425: +S\lr{-i{(s+x+1)}{\eta}}-S\lr{-i{(s+x)}{\eta}} \big]+ \mathcal{O}\lr{ x^{L}}
2426: \, .
2427: \ee
2428: Substituting this expression into \re{E-ex} one finally obtains the relation
2429: \be\label{E-app}
2430: E=f(0)-\frac1{2} f'(0)+ \ldots = -{2s}{\eta} S''(0)+\ldots= -\frac{2\beta}{L}
2431: S''(0)+\ldots\,,
2432: \ee
2433: which coincides with the semiclassical expression \re{E-naive-expansion}. It is
2434: important to keep in mind that Eq.~\re{E-app} was obtained under the assumption
2435: that contribution of terms with higher derivatives is small, $f''(0) \ll f'(0)$,
2436: or equivalently $ \eta S'''(0)\ll S''(0)$. This relation is not satisfied if the
2437: distribution of Bethe roots scales at the origin as $\sim \ln x$,
2438: Eq.~\re{S0-log}. In that case, one has to rely on the formula \re{E-main} which
2439: resums all singular terms.
2440:
2441: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2442: \begin{thebibliography}{99}
2443: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2444: %
2445: \bibitem{Col89}
2446: J.C. Collins,
2447: Adv. Ser. Direct. High Ener. Phys. 5 (1989) 573.
2448: %%CITATION = HEP-PH 0312336;%%
2449: %
2450: \bibitem{Kor88}
2451: G.P. Korchemsky,
2452: Mod. Phys. Lett. A 4 (1989) 1257; \\
2453: %%CITATION = MPLAE,A4,1257;%%
2454: %
2455: G.P. Korchemsky, G. Marchesini,
2456: Nucl. Phys. B 406 (1993) 225.
2457: %%CITATION = HEP-PH 9210281;%%
2458: %
2459: \bibitem{Pol80}
2460: A.M. Polyakov,
2461: Nucl. Phys. B 164 (1980) 171.
2462: %%CITATION = NUPHA,B164,171;%%
2463: %
2464: \bibitem{MocVerVog04}
2465: S. Moch, J.A.M. Vermaseren, A. Vogt,
2466: Nucl. Phys. B 688 (2004) 101.
2467: %%CITATION = HEP-PH 0403192;%%
2468: %
2469: \bibitem{BerDixSmi05}
2470: Z. Bern, L.J. Dixon, V.A. Smirnov,
2471: Phys. Rev. D 72 (2005) 085001.
2472: %%CITATION = HEP-TH 0505205;%%
2473: %
2474: \bibitem{GubKlePol03}
2475: S.S. Gubser, I.R. Klebanov, A.M. Polyakov,
2476: Nucl. Phys. B 636 (2002) 99.
2477: %%CITATION = HEP-TH 0204051;%%
2478: %
2479: \bibitem{Kru02}
2480: M. Kruczenski,
2481: J. High Ener. Phys. 0212 (2002) 024; \\
2482: %%CITATION = HEP-TH 0210115;%%
2483: %
2484: Yu. Makeenko,
2485: J. High Ener. Phys. 0301 (2003) 007.
2486: %%CITATION = HEP-TH 0210256;%%
2487: %
2488: \bibitem{BelGorKor03}
2489: A.V. Belitsky, A.S. Gorsky, G.P. Korchemsky,
2490: Nucl. Phys. B 667 (2003) 3.
2491: %%CITATION = HEP-TH 0304028;%%
2492: %
2493: \bibitem{Mal97}
2494: J.M. Maldacena, Adv. Theor. Math. Phys. 2 (1998) 231;\\
2495: %%CITATION = HEP-TH 9711200;%%
2496: %
2497: S.S. Gubser, I.R. Klebanov, A.M. Polyakov,
2498: Phys. Lett. B 428 (1998) 105;\\
2499: %%CITATION = HEP-TH 9802109;%%
2500: %
2501: E. Witten, Adv. Theor. Math. Phys. 2 (1998) 253.
2502: %%CITATION = HEP-TH 9802150;%%
2503: %
2504: \bibitem{BerMalNas02}
2505: D. Berenstein, J.M. Maldacena, H. Nastase,
2506: J. High Ener. Phys. 0204 (2002) 013.
2507: %%CITATION = HEP-TH 0202021;%%
2508: %
2509: \bibitem{FroTse03}
2510: S. Frolov, A.A. Tseytlin,
2511: J. High Ener. Phys. 0206 (2002) 007.
2512: %%CITATION = HEP-TH 0204226;%%
2513: %
2514: \bibitem{Tse03}
2515: A.A. Tseytlin, in Ian Kogan Memorial Volume, {\sl From Fields to Stings:
2516: Circumnavigating Theoretical Physics}, eds.\ M. Shifman, A. Vainshtein,
2517: J. Wheater, World Scientific (Singapore, 2004) vol.\ 2, pp. 1648--1707,
2518: hep-th/0311139.
2519: %%CITATION = HEP-TH 0311139;%%
2520: %
2521: \bibitem{BraDerMan98}
2522: V.M. Braun, S.E. Derkachov, A.N. Manashov,
2523: Phys. Rev. Lett. 81 (1998) 2020; \\
2524: %%CITATION = HEP-PH 9805225;%%
2525: %
2526: V.M. Braun, S.E. Derkachov, G.P. Korchemsky, A.N. Manashov,
2527: Nucl. Phys. B 553 (1999) 355.
2528: %%CITATION = HEP-PH 9902375;%%
2529: %
2530: \bibitem{Bel99}
2531: A.V. Belitsky,
2532: Phys. Lett. B 453 (1999) 59;
2533: %%CITATION = HEP-PH 9902361;%%
2534: %
2535: Nucl. Phys. B 558 (1999) 259;
2536: %%CITATION = HEP-PH 9903512;%%
2537: %
2538: Nucl. Phys. B 574 (2000) 407.
2539: %%CITATION = HEP-PH 9907420;%%
2540: %
2541: \bibitem{DerKorMan99}
2542: S.E. Derkachov, G.P. Korchemsky, A.N. Manashov,
2543: Nucl. Phys. B 566 (2000) 203.
2544: %%CITATION = HEP-PH 9909539;%%
2545: %
2546: \bibitem{Bel04}
2547: A.V. Belitsky, V.M. Braun, A.S. Gorsky, G.P. Korchemsky,
2548: Int. J. Mod. Phys. A 19 (2004) 4715.
2549: %%CITATION = HEP-TH 0407232;%%
2550: %
2551: \bibitem{TakFad79}
2552: L.A. Takhtajan, L.D. Faddeev,
2553: Russ. Math. Survey 34 (1979) 11; \\
2554: %%CITATION = RMSUA,34,11;%%
2555: %
2556: V.E. Korepin, N.M. Bogoliubov, A.G. Izergin,
2557: {\sl Quantum inverse scattering method and correlation functions},
2558: (Cambridge Univ. Press, 1993); \\
2559: %
2560: L.D. Faddeev, Int. J. Mod. Phys. A 10 (1995) 1845; Les Houches
2561: Lectures (1995), {\sl How algebraic Bethe ansatz works for integrable
2562: models}, hep-th/9605187.
2563: %%CITATION = HEP-TH 9605187;%%
2564: %
2565: \bibitem{BeiFroStaTse03}
2566: N. Beisert, S. Frolov, M. Staudacher, A.A. Tseytlin,
2567: J. High Ener. Phys. 0310 (2003) 037.
2568: %%CITATION = HEP-TH 0308117;%%
2569: %
2570: \bibitem{Bax72}
2571: R.J. Baxter,
2572: Annals Phys. 70 (1972) 193;
2573: %%CITATION = APNYA,70,193;%%
2574: %
2575: {\sl Exactly Solved Models in Statistical Mechanics},
2576: Academic Press (London, 1982).
2577: %
2578: \bibitem{Sut95}
2579: B. Sutherland,
2580: Lect. Notes Phys. 242 (1985) 1;
2581: Phys. Rev. Lett. 74 (1995) 816.
2582: %
2583: \bibitem{ManSurWad02}
2584: G. Mandal, N.V. Suryanarayana, S.R. Wadia,
2585: Phys. Lett. B 543 (2002) 81; \\
2586: %%CITATION = HEP-TH 0206103;%%
2587: %
2588: I. Bena, J. Polchinski, R. Roiban,
2589: Phys. Rev. D 69 (2004) 046002; \\
2590: %%CITATION = HEP-TH 0305116;%%
2591: %
2592: A.M. Polyakov,
2593: Mod. Phys. Lett. A 19 (2004) 1649.
2594: %%CITATION = HEP-TH 0405106;%%
2595: %
2596: \bibitem{ZakMik78}
2597: V.E. Zakharov, A.V. Mikhailov, Sov. Phys. JETP 47 (1978) 1017.
2598: %%CITATION = SPHJA,47,1017;%%
2599: %
2600: \bibitem{Kri94}
2601: I.M. Krichever, Funct. Anal. Appl. 28 (1994) 21.
2602: %
2603: \bibitem{KazZar04}
2604: V.A. Kazakov, K. Zarembo, J. High Ener. Phys. 0410 (2004) 060.
2605: %%CITATION = HEP-TH 0410105;%%
2606: %
2607: \bibitem{PasGau92}
2608: M. Gaudin, V. Pasquier, J. Phys. A 25 (1992) 5243.
2609: %%CITATION = JPAGB,A25,5243;%%
2610: %
2611: \bibitem{Kor95}
2612: G.P. Korchemsky,
2613: Nucl. Phys. B 462 (1996) 333;
2614: %%CITATION = HEP-TH 9508025;%%
2615: %
2616: Nucl. Phys. B 498 (1997) 68.
2617: %%CITATION = HEP-TH 9609123;%%
2618: %
2619: \bibitem{Smi98}
2620: F.A. Smirnov,
2621: Amer. Math. Soc. Trans. 201 (2000) 283.
2622: %%CITATION = HEP-TH 9802132;%%
2623: %
2624: \bibitem{Bei04}
2625: N. Beisert, Phys. Rept. 407 (2004) 1.
2626: %%CITATION = HEP-TH 0407277;%%
2627: %
2628: \bibitem{Skl90}
2629: E.K. Sklyanin,
2630: Lect. Notes Phys. 226 (1985) 196;
2631: %
2632: {\sl Quantum inverse scattering method. Selected topics\/},
2633: ``Quantum Group and Quantum Integrable Systems,''
2634: ed.\ Mo-Lin Ge, World Scientific, (Singapore, 1992) pp.\ 63--97, hep-th/9211111;
2635: %%CITATION = HEP-TH 9211111;%%
2636: %
2637: Progr. Theor. Phys. Suppl. 118 (1995) 35.
2638: %%CITATION = SOLV-INT 9504001;%%
2639: %
2640: \bibitem{Gor03}
2641: A.S. Gorsky,
2642: Theor. Math. Phys. 142 (2005) 153.
2643: %%CITATION = HEP-TH 0308182;%%
2644: %
2645: \bibitem{FadKor94}
2646: G.P. Korchemsky,
2647: Nucl. Phys. B 443 (1995) 255.
2648: %%CITATION = HEP-PH 9501232;%%
2649: %
2650: \bibitem{NovManPitZak84}
2651: S.P. Novikov, S.V. Manakov, L.P. Pitaevsky, V.E. Zakharov,
2652: {\sl Theory of soliton: the inverse scattering method},
2653: Consultants Bureau (New York, 1984).
2654: %
2655: \bibitem{FadTak87}
2656: L.D. Faddeev, L.A. Takhtajan,
2657: {\sl Hamiltonian methods in the theory of solitons},
2658: Springer-Verlag (Berlin, 1987).
2659: %
2660: \bibitem{BabBerTal03}
2661: O. Babelon, D. Bernard, M. Talon, {\sl Introduction to classical integrable
2662: systems}, (Cambridge University Press, 2003).
2663: %
2664: \bibitem{SmiRes83}
2665: N.Yu. Reshetikhin, F.A. Smirnov,
2666: Zap. Nauch. Sem. LOMI 131 (1983) 128.
2667: %
2668: \bibitem{KorKri97}
2669: G.P. Korchemsky, I.M. Krichever,
2670: Nucl. Phys. B 505 (1997) 387.
2671: %%CITATION = HEP-TH 9704079;%%
2672: %
2673: \bibitem{GroKaz05}
2674: N. Gromov, V. Kazakov,
2675: Nucl. Phys. B 736 (2006) 199.
2676: %%CITATION = HEP-TH 0510194;%%
2677: %
2678: \bibitem{Kru05}
2679: M. Kruczenski,
2680: J. High Ener. Phys. 0508 (2005) 014.
2681: %%CITATION = HEP-TH 0410226;%%
2682: %
2683: \end{thebibliography}
2684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2685:
2686: \end{document}
2687: