cond-mat0503313/art.tex
1: % FINAL VERSION AFTER CORRECTIONS IN PROOF 2/5/05
2: \documentclass[12pt]{article}
3: \usepackage{graphicx}
4: \setlength{\textwidth}{16cm}
5: %\setlength{\textwidth}{17cm}
6: %\setlength{\textheight}{24cm}
7: \setlength{\textheight}{22cm}
8: \setlength{\topmargin}{0cm}
9: \setlength{\oddsidemargin}{0cm}
10: \setlength{\parindent}{0cm}
11: 
12: \def\ffrac#1#2{\textstyle{#1\over#2}\displaystyle}
13: 
14: 
15: \begin{document}
16: \pagestyle{myheadings}
17: \parindent 0mm
18: \parskip 6pt
19: \markright{SLE for Theoretical Physicists}
20: 
21: \title{SLE for Theoretical Physicists}
22: 
23: \author{John Cardy\\
24: Rudolf Peierls Centre for Theoretical Physics\\
25:          1 Keble Road, Oxford OX1 3NP,
26:          U.K.\footnote{Address for correspondence.}\\
27: and All Souls College, Oxford.}
28: %
29: \date{May 2005}
30: %
31: \maketitle
32: %
33: \begin{abstract}
34: This article provides an introduction to
35: Schramm(stochastic)-Loewner evolution (SLE) and to its connection
36: with conformal field theory, from the point of view of its
37: application to two-dimensional critical behaviour. The emphasis is
38: on the conceptual ideas rather than rigorous proofs.
39: \end{abstract}
40: %
41: \newpage
42: %
43: 
44: \section{Introduction}
45: %
46: \subsection{Historical overview}
47: %
48: The study of critical phenomena has long been a breeding ground
49: for new ideas in theoretical physics. Such behaviour is
50: characterised by a diverging correlation length and cannot easily
51: be approximated by considering small systems with only a few
52: degrees of freedom. Initially, it appeared that the analytic study
53: of such problems was a hopeless task, although self-consistent
54: approaches such as mean field theory were often successful in
55: providing a semi-quantitative description.
56: 
57: Following Onsager's calculation of the free energy of the square
58: lattice Ising model in 1944, steady progress was made in the exact
59: solution of an ever-increasing number of lattice models in two
60: dimensions. While many of these are physically relevant, and the
61: techniques used have spawned numerous spin-offs in the theory of
62: integrable systems, it is fair to say that these methods have not
63: cast much light on the general nature of the critical state. In
64: addition, virtually no progress has been made from this direction
65: in finding analytic solutions to such simple and important
66: problems as percolation.
67: 
68: An important breakthrough occurred in the late 1960's, with the
69: development of renormalisation group (RG) ideas by Wilson and
70: others. The fundamental realisation was that, in the scaling limit
71: where both the correlation length and all other macroscopic length
72: scales are much larger than that of the microscopic interactions,
73: classical critical systems are equivalent to renormalisable
74: quantum field theories in euclidean space-time. Since there is
75: often only a finite or a denumerable set of such field theories
76: with given symmetries, at a stroke this explained the observed
77: phenomenon of universality: systems with very different
78: constituents and microscopic interactions nevertheless exhibit the
79: same critical behaviour in the scaling limit. This single idea has
80: led to a remarkable unification of the theoretical bases of
81: particle physics, statistical mechanics and condensed matter
82: theory, and has led to extensive cross-fertilisation between these
83: disciplines. These days, a typical paper using the ideas and
84: methods of quantum field theory is as likely to appear in a
85: condensed matter physics journal as in a particle physics
86: publication (although there seems to be a considerable degree of
87: conservatism among the writers of field theory text books in
88: recognising this fact.)
89: 
90: Two important examples of this interdisciplinary flow were the
91: development of lattice gauge theories in particle physics, and the
92: application of conformal field theory (CFT), first developed as a
93: tool in string theory, to statistical mechanics and condensed
94: matter physics. As will be explained later, in two-dimensional
95: classical systems and quantum systems in 1+1 dimensions conformal
96: symmetry is extremely powerful, and has led to a cornucopia of new
97: exact results. Essentially, the RG programme of classifying all
98: suitable renormalisable quantum field theories in two dimensions
99: has been carried through to its conclusion in many cases,
100: providing exact expressions for critical exponents, correlation
101: functions, and other universal quantities. However the
102: geometrical, as opposed to the algebraic, aspects of conformal
103: symmetry are not apparent in this approach.
104: 
105: One minor but nevertheless theoretically influential prediction of
106: these methods was the conjectured crossing formula
107: \cite{JCcrossing} for the probability that, in critical
108: percolation, a cluster should exist which spans between two
109: disjoint segments of the boundary of some simply connected region
110: (a more detailed account of this problem will be given later.)
111: With this result, the simmering unease that mathematicians felt
112: about these methods came to the surface (see, for example, the
113: comments in \cite{Lang}.) What exactly are these renormalised
114: local operators whose correlation functions the field theorists so
115: happily manipulate, according to rules that sometimes seem to be a
116: matter of cultural convention rather than any rigorous logic? What
117: does conformal symmetry really mean? Exactly which object is
118: conformally invariant? And so on. Aside from these deep concerns,
119: there was perhaps also the territorial feeling that percolation
120: theory, in particular, is a branch of probability theory, and
121: should be understood from that point of view, not merely as a
122: by-product of quantum field theory.
123: 
124: Thus it was that a number of pure mathematicians, versed in the
125: methods of probability theory, stochastic analysis and conformal
126: mapping theory, attacked this problem. Instead of trying make
127: rigorous the notions of field theory about local operators, they
128: focused on the random curves which form the boundaries of clusters
129: on the lattice, and on what should be the properties of the
130: measure on such curves in the continuum limit as the lattice
131: spacing approaches zero. The idea of thinking about lattice models
132: this way was not new: in particular in the 1980s it led to the
133: very successful but non-rigorous Coulomb gas approach
134: \cite{Nienhuis} to two-dimensional critical behaviour, whose
135: results parallel and complement those of CFT. However, the new
136: approach focused on the properties of a single such curve,
137: conditioned to start at the boundary of the domain, in the
138: background of all the others. This leads to a very specific and
139: physically clear notion of conformal invariance. Moreover, it was
140: shown by Loewner\cite{Loewner} in the 1920s that any such curve in
141: the plane which does not cross itself can be described by a
142: dynamical process called Loewner evolution, in which the curve is
143: imagined to be grown in a continuous fashion. Instead of
144: describing this process directly, Loewner considered the evolution
145: of the analytic function which conformally maps the region outside
146: the curve into a standard domain. This evolution, and therefore
147: the curve itself, turns out to be completely determined by a real
148: continuous function $a_t$. For random curves, $a_t$ itself is
149: random. (The notation $a_t$ is used rather than $a(t)$ to conform
150: to standard usage in the case when it is a stochastic variable.)
151: Schramm\cite{Schramm} argued that, if the measure on the curve is
152: to be conformally invariant in the precise sense referred to
153: above, the only possibility is that $a_t$ be one-dimensional
154: Brownian motion, with only a single parameter left undetermined,
155: namely the diffusion constant $\kappa$. This leads to stochastic-,
156: or Schramm-, Loewner evolution (SLE). (In the original papers by
157: Schramm et al. the term `stochastic' was used. However in the
158: subsequent literature the `S' has often been taken to stand for
159: Schramm in recognition of his contribution.) It should apply to
160: any critical statistical mechanics model in which it is possible
161: to identify these non-crossing paths on the lattice, as long as
162: their continuum limits obey the underlying conformal invariance
163: property. For only a few cases, including percolation, has it been
164: proved that this property holds, but it is believed to be true for
165: suitably defined curves in a whole class of systems known as
166: O$(n)$ models. Special cases, apart from percolation, include the
167: Ising model, Potts models, the XY model, and self-avoiding walks.
168: They each correspond to a particular choice of $\kappa$.
169: 
170: Starting from the assumption that SLE describes such a single
171: curve in one of these systems, many properties, such as the values
172: of many of the critical exponents, as well as the crossing formula
173: mentioned above, have been rigorously derived in a brilliant
174: series of papers by Lawler, Schramm and Werner (LSW)\cite{LSW}.
175: Together with Smirnov's proof \cite{Smirnov} of the conformal
176: invariance property for the continuum limit of site percolation on
177: the triangular lattice, they give a \em rigorous \em derivation
178: \cite{SmirnovWerner} of the values of the critical exponents for
179: two-dimensional percolation. This represents a paradigm shift in
180: rigorous statistical mechanics, in that results are now being
181: derived directly in the continuum for models for which the
182: traditional lattice methods have, so far, failed.
183: 
184: However, from the point of view of theoretical physics, these
185: advances are  perhaps not so important for being rigorous, as for
186: the new light they throw on the nature of the critical state, and
187: on conformal field theory. In the CFT of the O$(n)$ model, the
188: point where a random curve hits the boundary corresponds to the
189: insertion of a local operator which has a particularly simple
190: property: its correlation functions satisfy linear second-order
191: differential equations\cite{JCsurf}. These equations turn out to
192: be directly related to the Fokker-Planck type equations one gets
193: from the Brownian process which drives SLE. Thus there is a close
194: connection, at least at an operational level, between CFT and SLE.
195: This has been made explicit in a series of papers by Bauer and
196: Bernard\cite{BB} (see also \cite{FriedrichWerner}.) Other
197: fundamental concepts of CFT, such as the central charge $c$, have
198: their equivalence in SLE. This is a rapidly advancing subject, and
199: some of the more recent directions will be mentioned in the
200: concluding section of this article.
201: %
202: \subsection{Aims of this article}
203: The original papers on SLE are mostly both long and difficult,
204: using, moreover, concepts and methods foreign to most theoretical
205: physicists. There are reviews, in particular those by
206: Werner\cite{Wreview} and by Lawler\cite{Lreview} which cover much
207: of the important material in the original papers. These are
208: however written for mathematicians. A more recent review by Kager
209: and Nienhuis\cite{KNreview} describes some of the mathematics in
210: those papers in way more accessible to theoretical physicists, and
211: should be essential reading for any reader who wants then to
212: tackle the mathematical literature. A complete bibliography up to
213: 2003 appears in \cite{KadGruz}.
214: 
215: However, the aims of the present article are more modest. First,
216: it does not claim to be a thorough review, but rather a
217: semi-pedagogical introduction. In fact some of the material,
218: presenting some of the existing results from a slightly different,
219: and hopefully clearer, point of view, has not appeared before in
220: print. The article is directed at the theoretical physicist
221: familiar with the basic concepts of quantum field theory and
222: critical behaviour at the level of a standard graduate textbook,
223: and with a theoretical physicist's knowledge of conformal mappings
224: and stochastic processes. It is not the purpose to prove anything,
225: but rather to describe the concepts and methods of SLE, to relate
226: them to other ideas in theoretical physics, in particular CFT, and
227: to illustrate them with a few simple computations, which, however,
228: will be presented in a thoroughly non-rigorous manner. Thus, this
229: review is most definitely not for mathematicians interested in
230: learning about SLE, who will no doubt cringe at the lack of
231: preciseness in some of the arguments and perhaps be puzzled by the
232: particular choice of material. The notation used will be that of
233: theoretical physics, for example $\langle\cdots\rangle$ for
234: expectation value, and so will the terminology. The word
235: `martingale' has just made its only appearance. Perhaps the
236: largest omission is any account of the central arguments of
237: LSW\cite{LSW} which relate SLE to various aspects of Brownian
238: motion and thus allow for the direct computation of many critical
239: exponents. These methods are in fact related to two-dimensional
240: quantum gravity, whose role in this is already the subject of a
241: recent long article by Duplantier\cite{Dupreview}.
242: 
243: \subsubsection*{Acknowledgments.}
244: One may go only so far in learning a subject like this by reading
245: the original literature, and much of my knowledge, such as it is,
246: comes from seminars by, and informal discussions with, a large
247: number of people. To them I give thanks, as well as apologies if I
248: have occasionally misrepresented or oversimplified their ideas:
249: M.~Aizenman, M.~Bauer, R.~Bauer, V.~Beffara, J.~Dub\'edat,
250: D.~Bernard, M.~den Nijs, B.~Doyon, B.~Duplantier, R.~Friedrich,
251: I.~Gruzberg, M.~Hastings, J.~Jacobsen, L.~Kadanoff, W.~Kager,
252: R.~Kenyon, H.~Kesten, P.~Kleban, J.~Kondev, R.~Langlands,
253: G.~Lawler, B.~Nienhuis, S.~Rohde, Y.~Saint-Aubin, H.~Saleur,
254: O.~Schramm, S.~Sheffield, S.~Smirnov, W.~Werner, D.~Wilson. I also
255: thank B.~Doyon and V.~Riva for comments on the manuscript. This
256: work was begun while the author was a Member of the Institute for
257: Advanced Study, supported by the Ellentuck Fund.
258: %
259: \newpage
260: %
261: \section{Random curves and lattice models.}
262: \label{sec:lattice}
263: \subsection{The Ising and percolation models}
264: \label{sec:ising} In this section, we introduce the lattice models
265: which can be interpreted in terms of random non-intersecting paths
266: on the lattice whose continuum limit will be described by SLE.
267: 
268: The prototype is the Ising model. It is most easily realised on a
269: honeycomb lattice (see Fig.~\ref{fig:hex}).
270: %
271: \begin{figure}[h]
272: \centering
273: \includegraphics[width=8cm]{hex.eps}
274: \caption{\label{fig:hex}\small
275: Ising model on the honeycomb lattice, with loops corresponding to a
276: term in the expansion of (\ref{ZI}). Alternatively, these may be
277: thought of as domain walls of an Ising model on the dual triangular
278: lattice.}
279: \end{figure}
280: %
281: At each site $r$ is an Ising `spin' $s(r)$ which takes the
282: values $\pm1$. The partition function is
283: \begin{equation}
284: \label{ZI} Z_{\rm Ising}={\rm Tr}\,e^{\beta J\sum_{rr'}s(r)s(r')}
285: \propto{\rm Tr}\,\prod_{rr'}\left(1+xs(r)s(r')\right)\,,
286: \end{equation}
287: where $x=\tanh\beta J$, and the sum and product are over all edges
288: joining nearest neighbour pairs of sites. The trace operation is
289: defined as ${\rm Tr}=\prod_r\big(\frac12\sum_{s(r)}\big)$, so that
290: ${\rm Tr}\,s(r)^n=1$ if $n$ is even, and $0$ if it is odd.
291: 
292: At high temperatures ($\beta J\ll 1$) the spins are disordered,
293: and their correlations decay exponentially fast, while at low
294: temperatures ($\beta J\gg1$) there is long-range order: if the
295: spins on the boundary are fixed say, to the value $+1$, then
296: $\langle s(r)\rangle\not=0$ even in the infinite volume limit. In
297: between, there is a critical point. The conventional approach to
298: the Ising model focuses on the behaviour of the correlation
299: functions of the spins. In the scaling limit, they become local
300: operators in a quantum field theory (QFT). Their correlations are
301: power-law behaved at the critical point, which corresponds to a
302: massless QFT, that is a conformal field theory (CFT). From this
303: point of view (as well as exact lattice calculations) it is found
304: that correlation functions like $\langle s(r_1)s(r_2)\rangle$
305: decay at large separations according to power laws
306: $|r_1-r_2|^{-2x}$: one of the aims of the theory is to obtain the
307: values of the exponents $x$ as well as to compute, for example,
308: correlators depending on more than two points.
309: 
310: However, there is an alternative way of thinking about the
311: partition function (\ref{ZI}), as follows: imagine expanding out
312: the product to obtain $2^N$ terms, where $N$ is the total number
313: of edges. Each term may be represented by a subset of edges, or
314: graph $\cal G$, on the lattice, in which, if the term $xs(r)s(r')$
315: is chosen, the corresponding edge $(rr')$ is included in $\cal G$,
316: otherwise it is not. Each site $r$ has either 0, 1, 2 or 3 edges
317: in $\cal G$. The trace over $s(r)$ gives 1 if this number is even,
318: and 0 if it is odd. Each surviving graph is then the union of
319: non-intersecting closed loops (see Fig.~\ref{fig:hex}). In
320: addition, there can be open paths beginning and ending at a
321: boundary. For the time being, we suppress these by imposing `free'
322: boundary conditions, summing over the spins on the boundary. The
323: partition function is then
324: \begin{equation}
325: Z_{\rm Ising}=\sum_{\cal G}\,x^{\rm length}\,,
326: \end{equation}
327: where the length is the total of all the loops in $\cal G$.
328: When $x$ is small, the mean length of a single loop is small. The
329: critical point $x_c$ is signalled by a divergence of this quantity.
330: The low-temperature phase corresponds to $x>x_c$. While in this phase
331: the Ising spins are ordered, and their connected correlation functions\
332: decay exponentially, the loop gas is in fact still critical, in that, for
333: example, the probability that two points lie on the same loop has
334: a power-law dependence on their separation. This is the
335: \em dense \em phase.
336: 
337: The loops in $\cal G$ may be viewed in another way: as \em domain
338: walls \em for another Ising model on the dual lattice, which is a
339: triangular lattice whose sites $R$ lie at the centres of the
340: hexagons of the honeycomb lattice (see Fig.~\ref{fig:hex}). If the
341: corresponding interaction strength of this dual Ising model is
342: $(\beta J)^*$, then the Boltzmann weight for creating a segment of
343: domain wall is $e^{-2(\beta J)^*}$. This should be equated to
344: $x=\tanh(\beta J)$ above. Thus we see that the high-temperature
345: regime of the dual model corresponds to low temperature in the
346: original model, and vice versa. Infinite temperature in the dual
347: model ($(\beta J)^*=0$) means that the dual Ising spins are
348: independent random variables. If we colour each dual site with
349: $s(R)=+1$ black, and white if $s(R)=-1$, we have the problem of
350: \em site percolation \em on the triangular lattice, critical
351: because $p_c=\frac12$ for that problem. Thus the curves with $x=1$
352: correspond to percolation cluster boundaries. (In fact in the
353: scaling limit this is believed to be true throughout the dense
354: phase $x>x_c$.)
355: 
356: So far we have discussed only closed loops. Consider the spin-spin
357: correlation function
358: \begin{equation}
359: \langle s(r_1)s(r_2)\rangle= {{\rm
360: Tr}\,s(r_1)s(r_2)\prod_{rr'}\left(1+xs(r)s(r')\right) \over{\rm
361: Tr}\,\prod_{rr'}\left(1+xs(r)s(r')\right)}\,,
362: \end{equation}
363: where the sites $r_1$ and $r_2$ lie on the boundary. Expanding out
364: as before, we see that the surviving graphs in the numerator each
365: have a single edge coming into $r_1$ and $r_2$. There is therefore
366: a single open path $\gamma$ connecting these points on the
367: boundary (which does not intersect itself nor any of the closed
368: loops.) In terms of the dual variables, such a single open curve
369: may be realised by specifying the spins $s(R)$ on all the dual
370: sites on the boundary to be $+1$ on the part of the boundary
371: between $r_1$ and $r_2$ (going clockwise) and $-1$ on the
372: remainder. There is then a single domain wall connecting $r_1$ to
373: $r_2$. {\it SLE describes the continuum limit of such a curve
374: $\gamma$.}
375: 
376: Note that we could also choose $r_2$ to lie in the interior. The
377: continuum limit of such curves is then described by radial SLE
378: (Sec.~\ref{sec:radialsle}).
379: 
380: \subsubsection{Exploration process} \label{sec:exploration}
381: An important property of the ensemble of curves $\gamma$ on the
382: lattice is that, instead of generating a configuration of all the
383: $s(R)$ and then identifying the curve, it may be constructed
384: step-by-step as follows (see Fig.~\ref{fig:exploration}).
385: %
386: \begin{figure}
387: \centering
388: \includegraphics[width=8cm]{explore.eps}
389: \caption{\label{fig:exploration}\small
390: The exploration process for the Ising model. At each step the walk
391: turns L or R according to the value of the spin in front of it. The
392: relative probabilities are determined by the expectation value of this
393: spin given the fixed spins either side of the walk up to this time. The
394: walk never crosses itself and never gets trapped.}
395: \end{figure}
396: %
397: Starting
398: from $r_1$, at the next step it should turn R or L according to
399: whether the spin in front of it is $+1$ or $-1$. For independent
400: percolation, the probability of either event is $\frac12$, but for
401: $x<1$ it depends on the values of the spins on the boundary.
402: Proceeding like this, the curve will grow, with all the dual sites
403: on its immediate left taking the value $+1$, and those its right
404: the value $-1$. The relative probabilities of the path turning R
405: or L at a given step depend on the expectation value of the spin
406: on the site $R$ immediately in front of it, given the values of
407: the spins already determined, that is, given the path up to that
408: point. Thus the relative probabilities that the path turns R or L
409: are completely determined by the domain and the path up to that
410: point. This implies the crucial
411: 
412: {\bf Property 2.1} (lattice version). Let $\gamma_1$ be the part
413: of the total path covered after a certain number of steps. Then
414: the conditional probability distribution of the remaining part of
415: the curve, given $\gamma_1$, is the same as the unconditional
416: distribution of a whole curve, starting at the tip and ending at
417: $r_2$ in the domain ${\cal D}\setminus\gamma_1$.
418: 
419: In the Ising model, for example, if we already know part of the domain wall,
420: the rest of it can be considered as a complete domain
421: wall in a new region in which the left and right sides of the existing part
422: form part of the boundary.
423: This means the path is a history-dependent random walk. It can
424: be seen (Fig.~\ref{fig:exploration}) that when the growing tip
425: $\tau$
426: approaches an earlier section of the path, it must always turn away
427: from it: the tip never gets trapped.
428: There is always at least one path on the lattice from
429: the tip $\tau$ to the final point $r_2$.
430: 
431: \subsection{O$(n)$ model}\label{sec:Onmodel}
432: The loop gas picture of the Ising and percolation models may
433: simply be generalised by counting each closed loop with a fugacity
434: $n$:
435: \begin{equation}\label{On1}
436: Z_{{\rm O}(n)}=\sum_{\cal G}\,x^{\rm length}\,n^{\rm number\ of\
437: loops}\,.
438: \end{equation}
439: This is called the O$(n)$ model, for the reason that it gives the
440: partition function for $n$-component spins
441: ${\bf s}(r)=(s_1(r),\ldots s_n(r))$
442: with
443: \begin{equation}\label{On2}
444: Z_{{\rm O}(n)}={\rm Tr}\,\prod_{rr'}\left(1+ x{\bf s}(r)\cdot{\bf
445: s}(r')\right)\,,
446: \end{equation}
447: where ${\rm Tr}\,s_a(r)s_b(r)=\delta_{ab}$. Following the same
448: procedure as before we obtain the same set of closed loops (and
449: open paths) except that, on summing over the last spin in each
450: closed loop, we get a factor $n$. The model is called O$(n)$
451: because of its symmetry under rotations of the spins. The version
452: (\ref{On2}) makes sense only when $n$ is a positive integer (and
453: note that the form of the partition function is different from
454: that of the conventional O$(n)$ model, where the second term is
455: exponentiated.) The form in (\ref{On1}) is valid for general
456: values of $n$, and it gives a probability measure on the loop gas
457: for real $n\geq0$. However, the dual picture is useful only for
458: $n=1$ and $n=2$ (see below.) As for the case $n=1$, there is a
459: critical value $x_c(n)$ at which the mean loop length diverges.
460: Beyond this, there is a dense phase.
461: 
462: Apart from $n=1$, other important physical values of $n$ are:
463: \begin{itemize}
464: \item $n=2$. In this case we can view each loop as being oriented
465: in either a clockwise or anti-clockwise sense, giving it an overall
466: weight 2. Each loop configuration then corresponds to a
467: configuration of integer valued\ \em height \em variables $h(R)$
468: on the dual lattice, with the convention that the nearest
469: neighbour difference $h(R')-h(R)$ takes the values 0, $+1$ or $-1$
470: according to whether the edge crossed by $RR'$ is unoccupied,
471: occupied by an edge oriented at $90^{\circ}$ to $RR'$, or at
472: $-90^{\circ}$. (That is, the current running around each loop is
473: the lattice curl of $h$.) The variables $h(R)$ may be pictured as
474: the local height of a crystal surface. In the low-temperature
475: phase (small $x$) the surface is smooth: fluctuations of the
476: height differences decay exponentially with separation. In the
477: high-temperature phase it is rough: they grow logarithmically. In
478: between is a roughening transition. It is believed that relaxing
479: the above restriction on the height difference $h(R')-h(R)$ does
480: not change the universality class, as long as large values of this
481: difference are suppressed, for example using the weighting
482: $\exp\big[-\beta\big(h(R')-h(R)\big)^2\big]$. This is the discrete
483: Gaussian model. It is dual to a model of 2-component spins with
484: O$(2)$ symmetry called the XY model. \item $n=0$. In this case,
485: closed loops are completely suppressed, and we have a single
486: non-self-intersecting path connecting $r_1$ and $r_2$, weighted by
487: its length. Thus, all paths of the same length are counted with
488: equal weight. This is the self-avoiding walk problem, which is
489: supposed to describe the behaviour of long flexible polymer
490: chains. As $x\to x_c-$, the mean length diverges. The region
491: $x>x_c$ is the dense phase, corresponding to a long polymer whose
492: length is of the order of the area of the box, so that it has
493: finite density. \item $n=-2$ corresponds to the loop-erased random
494: walk. This is an ordinary random walk in which every loop, once it
495: is formed, is erased. Taking $n=-2$ in the $O(n)$ model of
496: non-intersecting loops has this effect.
497: \end{itemize}
498: %
499: \subsection{Potts model.}
500: Another important model which may described in terms of random
501: curves in the $Q$-state Potts model. This is most easily considered
502: on square lattice, at each site of which is a variable $s(r)$ which can take
503: $Q$ (initially a positive integer) different values. The partition function
504: is
505: \begin{equation}
506: Z_{\rm Potts}={\rm Tr}\,e^{\beta J\sum_{rr'}\delta_{s(r),s(r')}}
507: \propto{\rm
508: Tr}\,\prod_{rr'}\left(1-p+p\delta_{s(r),s(r')}\right)\,,
509: \end{equation}
510: with $e^{\beta J}=(1-p)^{-1}$. The product may be expanded in a
511: similar way to the case of the Ising model. All possible graphs
512: $\cal G$ will appear. Within each connected component of $\cal G$
513: the Potts spins must be equal, giving rise to a factor $Q$ when
514: the trace is performed. The result is
515: \begin{equation}
516: Z_{\rm Potts}=\sum_{\cal G}\,p^{|{\cal G}|}(1-p)^{|\overline{\cal
517: G}|} Q^{||{\cal G}||}\,,
518: \end{equation}
519: where $|{\cal G}|$ is the number of edges in $\cal G$,
520: $|\overline{\cal G}|$ the number in its complement, and $||{\cal
521: G}||$ is the number of connected components of $\cal G$, which are
522: called Fortuin-Kasteleyn (FK) clusters. This is the \em random
523: cluster \em representation of the Potts model. When $p$ is small,
524: the mean cluster size is small. As $p\to p_c$, it diverges, and
525: for $p>p_c$ there is an infinite cluster which contains a finite
526: fraction of all the sites in the lattice. It should be noted that
527: these FK clusters are \em not \em the same as the spin clusters
528: within which the original Potts spins all take the same value.
529: 
530: The limit $Q\to1$ gives another realisation of percolation -- this
531: time bond percolation on the square lattice. For $Q\to0$ there is
532: only one cluster. If at the same time $x\to0$ suitably, all loops
533: are suppressed and the only graphs $\cal G$ which contribute are
534: \em spanning trees\em, which contain every site of the lattice. In
535: the Potts partition function each possible spanning tree is
536: counted with the same weight, corresponding to the problem of
537: uniform spanning trees (UST). The ensemble of paths on USTs
538: connecting two points $r_1$ and $r_2$ turns out to be be that of
539: loop-erased random walks.
540: %
541: \begin{figure}
542: \centering
543: \includegraphics[width=8cm]{medial.eps}
544: \caption{\label{fig:medial}\small Example of FK clusters (heavy
545: lines) in the random cluster representation of the Potts model,
546: and the corresponding set of dense loops (medium heavy) on the
547: medial lattice. The loops never cross the edges connecting sites
548: in the same cluster.}
549: \end{figure}
550: %
551: 
552: The random cluster model may be realised as a gas of dense loops
553: in the way illustrated in Fig.~\ref{fig:medial}. These loops lie
554: on the medial lattice, which is also square but has twice the
555: number of sites. It may be shown that, at $p_c$, the weights for
556: the clusters are equivalent to counting each loop with a fugacity
557: $\sqrt Q$. Thus the boundaries of the critical FK clusters in the
558: $Q$-state Potts model are the same in the scaling limit (if it
559: exists) as the closed loops of the dense phase of the O$(n)$
560: model, with $n=\sqrt Q$.
561: 
562: To generate an open path in the random cluster model
563: connecting sites $r_1$ and $r_2$ on the boundary we must choose
564: `wired' boundary conditions, in which $p
565: =1$ on all the edges
566: parallel to the boundary, from $r_1$ to $r_2$, and free boundary
567: conditions, with $p=0$, along the remainder.
568: 
569: \subsection{Coulomb gas methods}\label{sec:coulomb}
570: Many important results concerning the O$(n)$ model can be derived
571: in a non-rigorous fashion using so-called Coulomb gas methods. For
572: the purposes of comparison with later results from SLE, we now
573: summarise these methods and collect a few relevant formulae. A
574: much more complete discussion may be found in the review by
575: Nienhuis\cite{Nienhuis}.
576: 
577: We assume that the boundary conditions on the O$(n)$ spins are free,
578: so that the partition function is a sum over closed loops only.
579: First orient each loop at random. Rather than giving clockwise and
580: anti-clockwise orientations the same weight $n/2$, give them complex
581: weights $e^{\pm 6i\chi}$, where $n=e^{6i\chi}+e^{-6i\chi}
582: =2\cos6\chi$. These may be taken into account, on the honeycomb lattice,
583: by assigning a weight $e^{\pm i\chi}$ at each vertex where an oriented
584: loop turns R (resp.~L). This transforms the non-local factors of $n$
585: into local (albeit complex) weights depending only on the local
586: configuration at each vertex.
587: 
588: Next transform to the height variables described above. By convention,
589: the heights are taken to be integer multiples of $\pi$. The local weights
590: at each vertex now depend only on the differences of the three adjacent
591: heights. The crucial assumption of the Coulomb gas approach is that,
592: under the RG, this model flows to one in which the lattice can be replaced
593: by a continuum, and the heights go over into a gaussian free field, with
594: partition function $Z=\int e^{-S[h]}[dh]$, where
595: \begin{equation}\label{gff}
596: S=(g/4\pi)\int(\nabla h)^2d^2r\,.
597: \end{equation}
598: As it stands, this is a simple free field theory.
599: The height fluctuations grow logarithmically:
600: $\langle\big(h(r_1)-h(r_2)\big)^2\rangle\sim(2/g)\ln|r_1-r_2|$,
601: and the correlators of exponentials of the height decay with power
602: laws:
603: \begin{equation}
604: \langle e^{iqh(r_1)}\,e^{-iqh(r_2)}\rangle\sim
605: |r_1-r_2|^{-2x_q}\,,
606: \end{equation}
607: where $x_q=q^2/2g$. All the subtleties come from the combined
608: effects of the phase factors and the boundaries or the topology.
609: This is particularly easy to see if we consider the model on a
610: cylinder of circumference $\ell$ and length $L\gg\ell$. In the
611: simple gaussian model (\ref{gff}) the correlation function between
612: two points a distance $L$ apart along the cylinder decays as
613: $\exp\big(-2\pi x_qL/\ell\big)$. However, if $\chi\not=0$, loops
614: which wrap around the cylinder are not counted correctly by the
615: above prescription, because the total number of left turns minus
616: right turns is then zero. We may arrange the correct factors by
617: inserting $e^{\pm 6i\chi h/\pi}$ at either end of the cylinder.
618: This has the effect of modifying the partition function: one finds
619: $\ln Z\sim (\pi c/6)(L/\ell)$ with
620: \begin{equation}\label{eqn:c6}
621: c=1-6{(6\chi/\pi)^2\over g}\,.
622: \end{equation}
623: This dependence of the partition function is one way of
624: determining the so-called central charge of the corresponding CFT
625: (Sec.~\ref{sec:cft}). The charges at each end of the cylinder also
626: modify the scaling dimension $x_q$ to
627: $(1/2g)\big((q-6i\chi/\pi)^2- (6i\chi/\pi)^2\big)$.
628: 
629: The value of $g$ may be fixed \cite{Kondev} in terms of the
630: original discreteness of the height variables as follows: adding a
631: term $-\lambda\int\cos2h\,d^2r$ to $S$ in (\ref{gff}) ensures
632: that, in the limit $\lambda\to\infty$, $h$ will be an integer
633: multiple of $\pi$. For this deformation not to affect the critical
634: behaviour, it must be marginal in the RG sense, which means that
635: it must have scaling dimension $x_2=2$. This condition then
636: determines $g=1-6\chi/\pi$.
637: 
638: \subsubsection{Winding angle distribution}\label{sec:winding1}
639: A simple property which can be inferred from the Coulomb gas
640: formulation is the winding angle distribution. Consider a cylinder
641: of circumference $2\pi$ and a path that winds around it. What the
642: probability that it winds through an angle $\theta$ around the
643: cylinder while it moves a distance $L\gg1$ along the axis? This
644: will correspond to a height difference $\Delta h=\pi(\theta/2\pi)$
645: between the ends of the cylinder, and therefore an additional free
646: energy $(g/4\pi)(2\pi L)(\theta/2L)^2$. The probability density is
647: therefore
648: \begin{equation}
649: P(\theta)\propto \exp(-g\theta^2/8L)\,,
650: \end{equation}
651: so that $\theta$ is normally distributed with variance $(4/g)L$.
652: This result will be useful later (Sec.~\ref{sec:radialsle}) for
653: comparison with SLE.
654: 
655: \subsubsection{$N$-leg exponent}\label{Nleg1}
656: As a final simple exponent prediction, consider the correlation
657: function $\langle\Phi_N(r_1)\Phi_N(r_2)\rangle$ of the $N$-leg
658: operator, which in the language of the O$(n)$ model is
659: $\Phi_N=s_{a_1}\cdots s_{a_N}$, where none of the indices are
660: equal. It gives the probability that $N$ mutually non-intersecting
661: curves connect the two points. Taking them a distance $L\gg\ell$
662: apart along the cylinder, we can choose to orient them all in the
663: same sense, corresponding to a discontinuity in $h$ of $N\pi$ in
664: going around the cylinder. Thus we can write $h=\pi Nv/\ell+\tilde
665: h$, where $0\leq v<\ell$ is the coordinate around the cylinder,
666: and $\tilde h(v+\ell)=\tilde h(v)$. This gives
667: \begin{equation}\label{eqn:Nleg}
668: \langle\Phi_N(r_1)\Phi_N(r_2)\rangle\sim
669: e^{-(g/4\pi)(N\pi/\ell)^2L+(\pi L/6\ell)-(\pi cL/6\ell)}\,.
670: \end{equation}
671: The second term in the exponent comes from the integral over the
672: fluctuations $\tilde h$, and the last from the partition function.
673: They differ because in the numerator, once there are curves
674: spanning the length of the cylinder, loops around it, which give
675: the correction term in (\ref{eqn:c6}), are forbidden.
676: (\ref{eqn:Nleg}) then gives
677: \begin{equation}
678: x_N=(gN^2/8)-(g-1)^2/2g\,.
679: \end{equation}
680: %
681: \newpage
682: %
683: \section{SLE}\label{sec:sle}
684: \subsection{The postulates of SLE.}\label{sec:postulates}
685: SLE gives a description of the continuum limit of the lattice
686: curves connecting two points on the boundary of a domain $\cal D$
687: which were introduced in Sec.~\ref{sec:lattice}. The idea is to
688: define a \em measure \em $\mu(\gamma;{\cal D},r_1,r_2))$ on these
689: continuous curves. (Note that the notion of a probability density
690: of such objects does not make sense, but the more general concept
691: of a measure does.)
692: 
693: There are two basic properties of this
694: continuum limit which must either be assumed, or, better, proven to hold
695: for a particular lattice model.
696: The first is the continuum version of Property 1:
697: 
698: {\bf Property 3.1} (continuum version). Denote the curve by
699: $\gamma$, and divide it into two disjoint parts: $\gamma_1$ from
700: $r_1$ to $\tau$, and $\gamma_2$ from $\tau$ to $r_2$. Then the
701: conditional measure $\mu(\gamma_2|\gamma_1;{\cal D},r_1,r_2)$ is
702: the same as $\mu(\gamma_2;{\cal D}\setminus\gamma_1,\tau,r_2)$.
703: 
704: This property we expect to be true for the scaling limit
705: of all such curves in the
706: O$(n)$ model (at least for $n\geq0$), even away from the critical
707: point. However the second property encodes the notion of conformal
708: invariance, and it should be valid, if at all, only at $x=x_c$
709: and, separately, for $x>x_c$.
710: 
711: {\bf Property 3.2} (conformal invariance.) Let $\Phi$ be a
712: conformal mapping of the interior of the domain $\cal D$ onto the
713: interior of ${\cal D}'$, so that the points $(r_1,r_2)$ on the
714: boundary of $\cal D$ are mapped to points $(r_1',r_2')$ on the
715: boundary of ${\cal D}'$. The measure $\mu$ on curves in $\cal D$
716: induces a measure $\Phi*\mu$ on the image curves in ${\cal D}'$.
717: The conformal invariance property states that this is the same as
718: the measure which would be obtained as the continuum limit of
719: lattice curves from $r_1'$ to $r_2'$ in ${\cal D}'$. That is
720: \begin{equation}
721: (\Phi*\mu)(\gamma;{\cal D},r_1,r_2)= \mu(\Phi(\gamma);{\cal
722: D}',r_1',r_2')\,.
723: \end{equation}
724: %
725: \subsection{Loewner's equation}\label{sec:loewner}
726: We have seen that, on the lattice, the curves $\gamma$ may be
727: `grown' through a discrete exploration process. The Loewner
728: process is the continuum version of this. Because of Property 2 it
729: suffices to describe this in a standard domain $\cal D$, which is
730: taken to be the upper half plane $\bf H$, with the points $r_1$
731: and $r_2$ being the origin and infinity respectively.
732: 
733: The first thing to notice is that, although on the honeycomb
734: lattice the growing path does not intersect itself, in the
735: continuum limit it might (although it still should not cross
736: itself.) This means that there may be regions enclosed by the path
737: which are not on the path but nevertheless are not reachable from
738: infinity without crossing it. We call the union of the set of such
739: points, together with the curve itself, up to time $t$, the \em
740: hull \em $K_t$. (This is a slightly different usage of this term
741: from that in the physics percolation literature.) It is the
742: complement of the connected component of the half plane which
743: includes $\infty$, itself denoted by ${\bf H}\setminus K_t$. See
744: Fig.~\ref{fig:hull}.
745: %
746: \begin{figure}
747: \centering
748: \includegraphics[width=6cm]{hull.eps}
749: \caption{\label{fig:hull}\small Schematic view of a trace and its
750: hull.}
751: \end{figure}
752: %
753: 
754: Another property which often holds in the half-plane is that of
755: \em reflection invariance\em: the distribution of lattice paths
756: starting from the origin and ending at $\infty$ is invariant under
757: $x\to-x$. For the lattice paths in the $O(n)$ model discussed in
758: Sec.~\ref{sec:Onmodel} this follows from the symmetry of the
759: underlying weights, but for the boundaries of the FK clusters in
760: the Potts model it is a consequence of duality. Not all simple
761: curves in lattice models have this property. For example if we
762: consider the 3-state Potts model in which the spins on the
763: negative and positive real axes are fixed to different values,
764: there is a simple lattice curve which forms the outer boundary of
765: the spin cluster containing the positive real axis. This is not
766: the same as the boundary of the spin cluster containing the
767: negative real axis, and it is not in general symmetric under
768: reflections.
769: 
770: Since ${\bf H}\setminus K_t$ is simply connected, by the Riemann
771: mapping theorem it can be mapped into the standard domain $\bf H$
772: by an analytic function $g_t(z)$. Because this preserves the real
773: axis outside $K_t$ it is in fact real analytic. It is not unique,
774: but can be made so by imposing the behaviour as $z\to\infty$
775: \begin{equation}\label{eqn:asymp}
776: g_t(z)\sim z+O(1/z)\,.
777: \end{equation}
778: It can be shown that, as the path grows, the coefficient of $1/z$
779: is monotonic increasing (essentially it is the electric dipole
780: moment of $K_t$ and its mirror image in the real axis.) Therefore
781: we may reparametrise time so that this coefficient is $2t$. (The
782: factor 2 is conventional.) Note that the length of the curve is
783: not be a useful parametrisation in the continuum limit, since the
784: curve is a fractal.
785: 
786: The function $g_t(z)$ maps the whole boundary of $K_t$ onto part
787: of the real axis. In particular, it maps the growing tip $\tau_t$
788: to a real point $a_t$. Any point on the real axis outside $K_t$
789: remains on the real axis. As the path grows, the point $a_t$ moves
790: on the real axis. For the path to describe a curve, it must always
791: grow only at its tip, and this means that the function $a_t$ must
792: be continuous, but not necessarily differentiable.
793: 
794: A simple but instructive example is when $\gamma$ is a straight
795: line growing vertically upwards from a fixed point $a$. In this
796: case
797: \begin{equation}\label{eqn:example}
798: g_t(z)=a+\big((z-a)^2+4t\big)^{1/2}\,.
799: \end{equation}
800: This satisfies (\ref{eqn:asymp}), and $\tau_t=2i\sqrt t$. More
801: complicated deterministic examples can be found\cite{KadKager}. In
802: particular, $a_t\propto t^{1/2}$ describes a straight line growing
803: at a fixed angle to the real axis.
804: 
805: Loewner's idea \cite{Loewner} was to describe the path $\gamma$
806: and the evolution of the tip $\tau_t$ in terms of the evolution of
807: the conformal mapping $g_t(z)$. It turns out that the equation of
808: motion for $g_t(z)$ is simple:
809: \begin{equation}\label{eqn:loewner}
810: {dg_t(z)\over dt}={2\over g_t(z)-a_t}\,.
811: \end{equation}
812: This is Loewner's equation. The idea of the proof is
813: straightforward. Imagine evolving the path for a time $t$, and
814: then for a further short time $\delta t$. The image of
815: $K_{t+\delta t}$ under $g_t$ is a short vertical line above the
816: point $a_t$ on the real axis. Thus we can write, using
817: (\ref{eqn:example})
818: \begin{equation}
819: g_{t+\delta t}(z)\approx a_t+\big((g_t(z)-a_t)^2+4\delta
820: t\big)^{1/2}\,.
821: \end{equation}
822: Differentiating with respect to $\delta t$ and then letting
823: $\delta t\to0$, we obtain (\ref{eqn:loewner}).
824: 
825: Note that, even if $a_t$ is not differentiable (as is the case for
826: SLE), (\ref{eqn:loewner}) gives for each point $z_0$ a solution
827: $g_t(z_0)$ which is differentiable with respect to $t$, up to the
828: time when $g_t(z_0)=a_t$. This is the time when $z_0$ is first
829: included in $K_t$. However, it is sometimes (see
830: Sec.~\ref{sec:cft}) useful to normalise the Loewner mapping
831: differently, defining $\hat g_t(z)=g_t(z)-a_t$, which always maps
832: the growing tip $\tau_t$ to the origin. If $a_t$ is not
833: differentiable, neither is $\hat g_t$, and the Loewner equation
834: should be written in differential form as $d\hat g_t=(2dt/\hat
835: g_t)-da_t$.
836: 
837: Given a growing path, we can determine the hull $K_t$ and hence,
838: in principle, the function $g_t(z)$ and thereby $a_t=g_t(\tau_t)$.
839: Conversely, given $a_t$ we can integrate (\ref{eqn:loewner}) to
840: find $g_t(z)$ and hence in determine the curve (although proving
841: that this inverse problem gives a curve is not easy.)
842: 
843: \subsection{Schramm-Loewner Evolution}\label{sec:schramm}
844: In the case that we are interested in, $\gamma$ is a random curve,
845: so that $a_t$ is a random continuous function. What is the measure
846: on $a_t$? This is answered by the following remarkable result, due
847: to Schramm\cite{Schramm}:
848: 
849: {\bf Theorem.} \em If Properties 3.1 and 3.2 hold, together with
850: reflection symmetry, then $a_t$ is proportional to a standard
851: Brownian motion.\em
852: 
853: That is
854: \begin{equation}
855: a_t=\sqrt\kappa B_t\,,
856: \end{equation}
857: so that $\langle a_t\rangle=0$,
858: $\langle\big(a_{t_1}-a_{t_2}\big)^2\rangle=\kappa|t_1-t_2|$. The
859: only undetermined parameter is $\kappa$, the diffusion constant.
860: It will turn out that different values of $\kappa$ correspond to
861: different universality classes of critical behaviour.
862: 
863: The idea behind the proof is once again simple. As before,
864: consider growing the curve for a time $t_1$, giving $\gamma_1$,
865: and denote the remainder $\gamma\setminus\gamma_1=\gamma_2$.
866: Property 3.1 tells us that the conditional measure on $\gamma_2$
867: given $\gamma_1$ is the same as the measure on $\gamma_2$ in the
868: domain ${\bf H}\setminus K_{t_1}$, which, by Property 3.2, induces
869: the same measure on $g_{t_1}(\gamma_2)$ in the domain ${\bf H}$,
870: shifted by $a_{t_1}$ (see Fig.~\ref{fig:loewner}).
871: %
872: \begin{figure}
873: \centering
874: \includegraphics[width=8cm]{loewner.eps}
875: \caption{\label{fig:loewner}\small A hull evolved from $a_0$ for
876: time $t_1$, then to infinity. The measure on the image of the rest
877: of the curve under $g_{t_1}$ is the same, according to the
878: postulates of SLE, as a hull evolved from $a_{t_1}$ to $\infty$.}
879: \end{figure}
880: %
881: In terms of the function $a_t$ this means that the probability law
882: of $a_t-a_{t_1}$, for $t>t_1$, is the same as the law of
883: $a_{t-t_1}$. This implies that all the increments $\Delta_n \equiv
884: a_{(n+1)\delta t}-a_{n\delta t}$ are independent identically
885: distributed random variables, for all $\delta t>0$. The only
886: process that satisfies this is Brownian motion with a possible
887: drift term: $a_t=\sqrt\kappa B_t+\alpha t$. Reflection symmetry
888: then implies that $\alpha=0$.
889: 
890: \subsection{Simple Properties of SLE}\label{sec:simpleproperties}
891: \subsubsection{Phases of SLE}\label{sec:phases}
892: Many of the results discussed in this section have been proved by
893: Rohde and Schramm\cite{RS}. First we address the question of how
894: the trace (the trajectory of $\tau_t$) looks for different values
895: of $\kappa$. For $\kappa=0$, it is a vertical straight line. As
896: $\kappa$ increases, the trace should randomly turn to the L or R
897: more frequently. However, it turns out that there are qualitative
898: differences at critical values of $\kappa$. To see this, let us
899: first study the process on the real axis. Let
900: $x_t=g_t\big(x_0\big)-a_t$ be the distance between the image at
901: time $t$ of a point which starts at $x_0$ and the image $a_t$ of
902: the tip. It obeys the stochastic equation
903: \begin{equation}\label{eqn:bessel}
904: dx_t={2dt\over x_t}-\sqrt\kappa dB_t\,.
905: \end{equation}
906: Physicists often write such an equation as $\dot x=(2/x)-\eta_t$
907: where $\eta_t$ is `white noise' of strength $\kappa$. Of course
908: this does not make sense since $x_t$ is not differentiable. Such
909: equations are always to be interpreted in the `Ito sense', that
910: is, as the limit as $\delta t\to0$ of the forward difference
911: equation $x_{t+\delta t}\approx x_t+(2\delta
912: t/x_t)+\int_t^{t+\delta t}\eta_{t'}dt'$.
913: 
914: (\ref{eqn:bessel}) is known as the Bessel process. (If we set
915: $R_t=(D-1)^{1/2}x_t/2$ and $\kappa^2=4/(D-1)$ it describes the
916: distance $R_t$ from the origin of a Brownian particle in $D$
917: dimensions.) The point $x_t$ is repelled from the origin but it is
918: also subject to a random force. Its ultimate fate can be inferred
919: from the following crude argument: if we ignore the random force,
920: $x_t^2\sim 4t$, while, in the absence of the repulsive term,
921: $\langle x_t^2\rangle\sim\kappa t$. Thus for $\kappa<4$ the
922: repulsive force wins and the particle escapes to infinity, while
923: for $\kappa>4$ the noise dominates and the particle collides with
924: the origin in finite time (at which point the equation breaks
925: down.) A more careful analysis confirms this.
926: %
927: \begin{figure}
928: \centering
929: \includegraphics[width=5cm]{enclose.eps}
930: \caption{\label{fig:enclose}\small The trace is about to hit the
931: axis at $x_0$ and enclose a region. At the time this happens, the
932: whole region including the point $x_0$ is mapped by $g_t$ to the
933: same point $a_t$.}
934: \end{figure}
935: %
936: %
937: What does this collision signify in  terms of the behaviour of the
938: trace? In Fig.~\ref{fig:enclose} we show a trace which is about to
939: hit the real axis at the point $x_0$, thus engulfing a whole
940: region. This is visible from infinity only through a very small
941: opening, which means that, under $g_t$, it gets mapped to a very
942: small region. In fact, as the tip $\tau_t$ approaches $x_0$, the
943: size of the image of this region shrinks to zero. When the gap
944: closes, the whole region enclosed by the trace, as well as
945: $\tau_t$ and $x_0$, are mapped in to the single point $a_t$, which
946: means, in particular, that $x_t\to0$. The above argument shows
947: that for $\kappa<4$ this never happens: the trace never hits the
948: real axis (with probability 1.) For the same reason, it neither
949: hits itself. Thus for $\kappa<4$ the trace $\gamma$ is a \em
950: simple curve\em.
951: 
952: The opposite is true for $\kappa>4$: points on the real axis are
953: continually colliding with the image $a_t$ of the tip. This means
954: that the trace is continually touching both itself and the real
955: axis, at the same time engulfing whole regions. Moreover, since it
956: is self-similar object, it does this on all scales, an infinite
957: number of times within any finite neighbourhood! Eventually the
958: trace swallows the whole half plane: every point is ultimately
959: mapped into $a_t$. For $\kappa<4$ only the points on the trace
960: itself suffer this fate. The case $\kappa=4$ is more tricky: in
961: fact the trace is then also a simple curve.
962: 
963: When $\kappa$ is just above 4, the images of points on the real
964: axis under $g_t$ collide with $a_t$ only when there happen to be
965: rare events when the random force is strong enough to overcome the
966: repulsion. When this happens, whole segments of the real axis are
967: swallowed at one time, corresponding to the picture described
968: above. Conversely, for large $\kappa$, the repulsive force is
969: negligible except for very small $x_t$. In that case, two
970: different starting points move with synchronised Brownian motions
971: until the one which started off closer to the origin is swallowed.
972: Thus the real line is eaten up in a continuous fashion rather than
973: piecemeal. There are no finite regions swallowed by the trace
974: which are not on the trace itself. This means that the trace is
975: \em space-filling\em: $\gamma$ intersects every neighbourhood of
976: every point in the upper half plane. We shall argue later
977: (Sec.~\ref{sec:fractaldim}) that the fractal dimension of the
978: trace is $d_f=1+\kappa/8$ for $\kappa\leq8$ and 2 for
979: $\kappa\geq8$. Thus it becomes space-filling for all
980: $\kappa\geq8$.
981: 
982: \subsubsection{SLE duality}\label{sec:duality}
983: For $\kappa>4$ the curve continually touches itself and therefore
984: its hull $K_t$ contains earlier portions of the trace (see
985: Fig.~\ref{fig:hull}). However, the \em frontier \em of $K_t$ (i.e.
986: the boundary of ${\bf H}\setminus K_t$, minus any portions of the
987: real axis), is by definition a simple curve. A beautiful result,
988: first suggested by Duplantier\cite{Dupduality}, and proved by
989: Beffara\cite{Beffaraduality} for the case $\kappa=6$, is that
990: locally this curve is an SLE$_{\tilde\kappa}$, with
991: \begin{equation}
992: \tilde\kappa=16/\kappa\,.
993: \end{equation}
994: For example, the exterior of a percolation cluster contains many
995: `fjords' which, on the lattice, are connected to the main ocean by
996: a neck of water which is only a finite number of lattice spacings
997: wide. These are sufficiently frequent and the fjords
998: macroscopically large that they survive in the continuum limit.
999: SLE$_{6}$ describes the boundaries of the clusters, including the
1000: coastline of all the fjords. However, the coastline as seen from
1001: the ocean is a simple curve, which is locally SLE$_{8/3}$, the
1002: same as that conjectured for a self-avoiding walk. This suggests,
1003: for example, that locally the frontier of a percolation cluster
1004: and a self-avoiding walk are the same in the scaling limit. In
1005: Sec.~\ref{sec:cft} we show that SLE$_\kappa$ and
1006: SLE$_{\tilde\kappa}$ correspond to CFTs with the same value of the
1007: central charge $c$.
1008: 
1009: 
1010: \subsection{Special values of $\kappa$}
1011: \subsubsection{Locality}\label{sec:locality}
1012: [This subsection and the next are more technical and may be
1013: omitted at a first reading.] We have defined SLE in terms of
1014: curves which connect the origin and infinity in the upper half
1015: plane. Property 2 then allows us to define it for any pair of
1016: boundary points in any simply connected domain, by a conformal
1017: mapping. It is interesting to study how the variation of the
1018: domain affects the SLE equation. Let $A$ be a simply connected
1019: region connected to the real axis which is at some finite distance
1020: from the origin (see Fig.~\ref{fig:bump}).
1021: %
1022: \begin{figure}
1023: \centering
1024: \includegraphics[width=10cm]{bump.eps}
1025: \caption{\label{fig:bump}\small An SLE hull in ${\bf H}\setminus
1026: A$ and two different ways of removing it: either by first removing
1027: $A$ through $h_0$ and then using a Loewner map $\tilde g_t$ in the
1028: image of ${\bf H}\setminus A$; or by removing $K_t$ first with
1029: $g_t$ and then removing the image of $A$ with $h_t$. Since all
1030: maps are normalised, this diagram commutes.}
1031: \end{figure}
1032: %
1033: Consider a trace $\gamma_t$, with hull $K_t$, which grows from the
1034: origin according to SLE in the domain ${\bf H}\setminus A$.
1035: According to Property 2, we can do this by first making a
1036: conformal mapping $h_0$ which removes $A$, and then a map $\tilde
1037: g_t$ which removes the image $\tilde K_t=h_0(K_t)$. This would be
1038: described by SLE in $h_0({\bf H}\setminus A)$, except that the
1039: Loewner `time' would not in general be the same as $t$. However,
1040: another way to think about this is to first use a SLE map $g_t$ in
1041: $\bf H$ to remove $K_t$, then another map, call it $h_t$, to
1042: remove $g_t(A)$. Since both these procedures end up removing
1043: $K_t\cup A$, and all the maps are assumed to be normalised at
1044: infinity in the standard way (\ref{eqn:asymp}), they must be
1045: identical, that is $h_t\circ g_t=\tilde g_t\circ h_0$ (see
1046: Fig.~\ref{fig:bump}). If $g_t$ maps the growing tip $\tau_t$ to
1047: $a_t$, then after both mappings it goes to $\tilde a_t=h_t(a_t)$.
1048: We would like to understand the law of $\tilde a_t$.
1049: 
1050: Rather than working this out in full generality (see for example
1051: \cite{Wreview}), let us suppose that $A$ is a short vertical
1052: segment $(x,x+i\epsilon)$ with $\epsilon\ll x$, and that $t=dt$ is
1053: infinitesimal. Then, under $g_{dt}$, $x\to x+2dt/x$ and
1054: $\epsilon\to\epsilon(1-2dt/x^2)$. The map that removes this is
1055: (see (\ref{eqn:example}))
1056: \begin{equation}\label{eqn:remove}
1057: h_{dt}(z)=\left((z-x-2dt/x)^2+\epsilon^2(1-2dt/x^2)^2\right)^{1/2}
1058: +x+2dt/x\,.
1059: \end{equation}
1060: In order to find $\tilde a_{dt}$, we need to set
1061: $z=a_{dt}=\sqrt\kappa dB_t$ in this expression. Carefully
1062: expanding this to first order in $dt$, remembering that
1063: $(dB_t)^2=dt$, and also taking the first non-zero contribution in
1064: $\epsilon/x$, gives after a few lines of algebra
1065: \begin{equation}\label{eqn:kappa6}
1066: \tilde a_{dt}=(1-\epsilon^2/x^2)\sqrt\kappa
1067: dB_t+\ffrac12(\kappa-6)(\epsilon^2/x^3)dt\,.
1068: \end{equation}
1069: The factor in front of the stochastic term may be removed by
1070: rescaling $dt$: this restores the correct Loewner time. But there
1071: is also a drift term, corresponding to the effect of $A$. For
1072: $\kappa<6$ we see that the SLE is initially repelled from $A$.
1073: From the point of view of the exploration process for the Ising
1074: model discussed in Sec.~\ref{sec:exploration}, this makes sense:
1075: if the spins along the positive real axis and on $A$ are fixed to
1076: be up, then the spin just above the origin is more likely to be up
1077: than down, and so $\gamma$ is more likely to turn to the left.
1078: 
1079: For $\kappa=6$, however, this is no longer the case: the presence
1080: of $A$ does not affect the initial behaviour of the curve. This is
1081: a particular case of the property of \em locality \em when
1082: $\kappa=6$, which states that, for any $A$ as defined above, the
1083: law of $K_t$ in ${\bf H}\setminus A$ is, up to a time
1084: reparametrisation, the same as the law of $K_t$ in $\bf H$, as
1085: long as $K_t\cap A=\emptyset$. That is, up to the time that the
1086: curve hits $A$, it doesn't know it's there. Such a property would
1087: be expected for the cluster boundaries of uncorrelated Ising spins
1088: on the lattice, i.e.~percolation. This is then consistent with the
1089: identification of percolation cluster boundaries with SLE$_6$.
1090: %
1091: \subsubsection{Restriction}\label{sec:restriction}
1092: %
1093: It is also interesting to work out how the local scale transforms
1094: in going from $a_t$ to $\tilde a_t$. A measure of this is
1095: $h'_t(a_t)$. A similar calculation starting from
1096: (\ref{eqn:remove}) gives, in the same limit as above,
1097: \begin{equation}
1098: d\big(h'(a_t)\big)=
1099: h'_{dt}(a_{dt})-h_0'(0)=(\epsilon^2/x^3)\sqrt\kappa
1100: dB_t+\ffrac12\left(
1101: (\epsilon^4/x^6)+(\kappa-\ffrac83)(3\epsilon^2/x^4)\right)\,.
1102: \end{equation}
1103: Now something special happens when $\kappa=\frac83$. The drift
1104: term in $d\big(h'(a_t)\big)$ does not then vanish, but if we take
1105: the appropriate power $d\big(h'_t(a_t)^{5/8}\big)$ it does. This
1106: implies that the \em mean \em of $h'_t(a_t)^{5/8}$ is conserved.
1107: Now at $t=0$ it takes the value $\Phi_A'(0)^{5/8}$, where
1108: $\Phi_A=h_0$ is the map that removes $A$. If $K_t$ hits $A$ at
1109: time $T$ it can be seen from (\ref{eqn:remove}) that $\lim_{t\to
1110: T}h'_t(a_t)^{5/8}=0$. On the other hand, if it never hits $A$ then
1111: $\lim_{t\to\infty}h'_t(a_t)^{5/8}=1$. Therefore $\Phi_A'(0)^{5/8}$
1112: gives the \em probability that the curve $\gamma$ does not
1113: intersect $A$\em.
1114: 
1115: This is a remarkable result in that it depends only on the value
1116: of $\Phi'_A$ at the starting point of the SLE (assuming of course
1117: that $\Phi_A$ is correctly normalised at infinity.) However it has
1118: the following even more interesting consequence. Let
1119: $\hat\Phi_A(z)=\Phi_A(z)-\Phi(0)$. Consider the ensemble of all
1120: SLE$_{8/3}$ in $\bf H$, and the sub-ensemble consisting of all
1121: those curves $\gamma$ which do not hit $A$. Then the measure on
1122: the image $\hat\Phi_A(\gamma)$ in $\bf H$ is again given by
1123: SLE$_{8/3}$. The way to show this is to realise that the measure
1124: on $\gamma$ is characterised by the probability $P(\gamma\cap
1125: A'=\emptyset)$ that $\gamma$ does not hit $A'$ for all possible
1126: $A'$. The probability that $\hat\Phi_A(\gamma)$ does not hit $A'$,
1127: given that $\gamma$ does not hit $A$, is the ratio of the
1128: probabilities $P(\gamma\cap{\hat\Phi_A}^{-1}(A')=\emptyset)$ and
1129: $P(\gamma\cap A=\emptyset)$. By the above result, the first factor
1130: is the derivative at the origin of the map
1131: $\hat\Phi_{A'}\circ\hat\Phi_A$ which removes $A$ then $A'$, while
1132: the second is the derivative of the map which removes $A$. Thus
1133: \begin{equation}
1134: P(\hat\Phi_A(\gamma)\cap A'=\emptyset|\gamma\cap A=\emptyset)=
1135: \left({(\hat\Phi_{A'}\circ\hat\Phi_A)'(0)\over
1136: {\hat\Phi_A}'(0)}\right)^{5/8}={\hat\Phi_{A'}}'(0)^{5/8}
1137: =P(\gamma\cap A'=\emptyset)\,.
1138: \end{equation}
1139: Since this is true for all $A'$, it follows that the law of
1140: $\hat\Phi_A(\gamma)$ given that $\gamma$ does not intersect $A$ is
1141: the same as that of $\gamma$. This is called the \em restriction
1142: property\em. Note that while, according to Property 2, the law of
1143: an SLE in any simply connected subset of $\bf H$ is determined by
1144: the conformal mapping of this subset to $\bf H$, the restriction
1145: property is stronger than this, and it holds only when
1146: $\kappa=\frac83$.
1147: 
1148: We expect such a property to hold for the continuum limit of
1149: self-avoiding walks, assuming it exists. On the lattice, every
1150: walk of the same length is counted with the same weight. That is,
1151: the measure is uniform. If we consider the sub-ensemble of such
1152: walks which avoid a region $A$, the measure on the remainder
1153: should still be uniform. This will be true if the restriction
1154: property holds. This supports the identification of self-avoiding
1155: walks with SLE$_{8/3}$.
1156: 
1157: 
1158: \subsection{Radial SLE and the winding angle}\label{sec:radialsle}
1159: So far we have discussed a version of SLE that gives a conformally
1160: invariant measure on curves which connect two distinct boundary
1161: points of a simply connected domain $\cal D$. For this reason it
1162: is called \em chordal \em SLE. There are variants which describe
1163: other situations. For example, one could consider curves $\gamma$
1164: which connect a given point $r_1$ on the boundary to an interior
1165: point $r_2$. The Riemann mapping theorem allows us to map
1166: conformally onto another simple connected domain, with $r_2$ being
1167: mapped to any preassigned interior point. It is simplest to choose
1168: for the standard domain the unit disc $\bf U$, with $r_2$ being
1169: mapped to the origin. So we are considering curves $\gamma$ which
1170: connect a given point $e^{i\theta_0}$ on the boundary with the
1171: origin. As before, we may consider growing the curve dynamically.
1172: Let $K_t$ be the hull of that portion which exists up to time $t$.
1173: Then there exists a conformal mapping $g_t$ which takes ${\bf
1174: U}\setminus K_t$ to $\bf U$, such that $g_t(0)=0$. There is one
1175: more free parameter, which corresponds to a global rotation: we
1176: use this to impose the condition that $g_t'(0)$ is real and
1177: positive. One can then show that, as the curve grows, this
1178: quantity is monotonically increasing, and we can use this to
1179: reparametrise time so that $g_t'(0)=e^t$. This normalised mapping
1180: then takes the growing tip $\tau_t$ to a point $e^{i\theta_t}$ on
1181: the boundary.
1182: 
1183: Loewner's theorem tells us that $\dot g_t(z)/g_t(z)$, when
1184: expressed as a function of $g_t(z)$, should be holomorphic in
1185: $\overline{{\bf U}}$ apart from a simple pole at $e^{i\theta_t}$.
1186: Since $g_t$ preserves the unit circle outside $K_t$, $\dot
1187: g_t(z)/g_t(z)$ should be pure imaginary when $|g_t(z)|=1$, and in
1188: order that $g_t'(0)=e^t$, it should approach 1 as $g_t(z)\to0$.
1189: The only possibility is
1190: \begin{equation}\label{eqn:radialsle}
1191: {dg_t(z)\over dt}=-g_t(z)\,{g_t(z)+e^{i\theta_t} \over
1192: g_t(z)-e^{i\theta_t}}\,.
1193: \end{equation}
1194: This is the radial Loewner equation. In fact this is the version
1195: considered by Loewner\cite{Loewner}.
1196: 
1197: It can now be argued, as before, that given Properties 1 \& 2
1198: (suitably reworded to cover the case when $r_2$ is an interior
1199: point) together with reflection, $\theta_t$ must be proportional
1200: to a standard Brownian motion. This defines radial SLE. It is not
1201: immediately obvious how the radial and chordal versions are
1202: related. However, it can be shown that, if the trace of radial SLE
1203: hits the boundary of the unit disc at $e^{i\theta_{t_1}}$ at time
1204: $t_1$, then the law of $K_t$ in radial SLE, for $t<t_1$, is the
1205: same chordal SLE conditioned to begin at $e^{i\theta(0)}$ and end
1206: at $e^{i\theta_{t_1}}$, up to a reparametrisation of time. This
1207: assures us that, in using the chordal and radial versions with the
1208: same $\kappa$, we are describing the same physical problem.
1209: 
1210: However, one feature that the trace of radial SLE possesses which
1211: chordal SLE does not is the property that it can wind around the
1212: origin. The winding angle at time $t$ is simply
1213: $\theta_t-\theta_0$. Therefore it is normally distributed with
1214: variance $\kappa t$. At this point we can make a connection to the
1215: Coulomb gas analysis of the O$(n)$ model in
1216: Sec.~\ref{sec:winding1}, where it was shown that the variance in
1217: the winding angle on a cylinder of length $L$ is asymptotically
1218: $(4/g)L$. A semi-infinite cylinder, parametrised by $w$, is
1219: conformally equivalent to the unit disc by the mapping $z=e^{-w}$.
1220: Asymptotically, ${\rm Re}\,w\to{\rm Re}\,w-t$ under Loewner
1221: evolution. Thus we can identify $L\sim t$ and hence
1222: \begin{equation}\label{eqn:kappag}
1223: \kappa=4/g\,.
1224: \end{equation}
1225: \subsubsection{Identification with lattice
1226: models}\label{sec:identification} This result allows use to make a
1227: tentative identification with the various realisations of the
1228: O$(n)$ model described in Sec.~\ref{sec:Onmodel}. We have, using
1229: (\ref{eqn:kappag}), $n=-2\cos(4\pi/\kappa)$ with
1230: $2\leq\kappa\leq4$ describing the critical point at $x_c$, and
1231: $4<\kappa\leq8$ corresponding to the dense phase. Some important
1232: special cases are therefore:
1233: \begin{itemize}
1234: \item $\kappa=-2$: loop-erased random walks (proven in
1235: \cite{LSWUST}); \item $\kappa=\frac83$: self-avoiding walks, as
1236: already suggested by the restriction property,
1237: Sec.~\ref{sec:restriction}; unproven, but see \cite{LSWSAW} for
1238: many consequences; \item $\kappa=3$: cluster boundaries in the
1239: Ising model, however as yet unproven; \item $\kappa=4$: BCSOS
1240: model of roughening transition (equivalent to the 4-state Potts
1241: model and the double dimer model), as yet unproven; also certain
1242: level lines of a gaussian random field and the `harmonic explorer'
1243: (proven in \cite{SSHE}); also believed to be dual to the
1244: Kosterlitz-Thouless transition in the XY model; \item $\kappa=6$:
1245: cluster boundaries in percolation (proven in \cite{Smirnov});
1246: \item $\kappa=8$: dense phase of self-avoiding walks; boundaries
1247: of uniform spanning trees (proven in \cite{LSWUST}).
1248: \end{itemize}
1249: 
1250: It should be noted that no lattice candidates for $\kappa>8$, or
1251: for the dual values $\kappa<2$, have been proposed. This possibly
1252: has to do with the fact that, for $\kappa>8$, the SLE trace is not
1253: reversible: the law on curves from $r_1$ to $r_2$ is not the same
1254: as the law obtained by interchanging the points. Evidently, curves
1255: in equilibrium lattice models should satisfy reversibility.
1256: %
1257: \newpage
1258: %
1259: \section{Calculating with SLE}\label{sec:calc}
1260: SLE shows that the measure on the continuum limit of single curves
1261: in various lattice models is given in terms of 1d Brownian motion.
1262: However, it is not at all clear how thereby to deduce interesting
1263: physical consequences.  We first describe two relatively simple
1264: computations in two-dimensional percolation which can be done
1265: using SLE.
1266: \subsection{Schramm's formula}\label{sec:schrammformula}
1267: Given a curve $\gamma$ connecting two points $r_1$ and $r_2$ on
1268: the boundary of a domain $\cal D$, what is the probability that it
1269: passes to the left of a given interior point? This is not a
1270: question which is natural in conventional approaches to critical
1271: behaviour, but which is very simply answered within
1272: SLE\cite{Schrammformula}.
1273: 
1274: As usual, we can consider $\cal D$ to be the upper half plane, and
1275: take $r_1=a_0$ and $r_2$ to be at infinity. The curve is then
1276: described by chordal SLE starting at $a_0$. Label the interior
1277: point by the complex number $\zeta$.
1278: 
1279: Denote the probability that $\gamma$ passes to the left of $\zeta$
1280: by $P(\zeta,\bar\zeta;a_0)$ (we include the dependence on
1281: $\bar\zeta$ to emphasise the fact that this is a not a holomorphic
1282: function.) Consider evolving the SLE for an infinitesimal time
1283: $dt$. The function $g_{dt}$ will map the remainder of $\gamma$
1284: into its image $\gamma'$, which, however, by Properties 1 \& 2,
1285: will have the same measure as SLE started from
1286: $a_{dt}=a_0+\sqrt\kappa dB_t$. At the same time, $\zeta\to
1287: g_{dt}(\zeta)=\zeta+2dt/(\zeta-a_0)$. Moreover, $\gamma'$ lies to
1288: the left of $\zeta'$ iff $\gamma$ lies to the left of $\zeta$.
1289: Therefore
1290: \begin{equation}
1291: P\big(\zeta,\bar\zeta;a_0\big)= \langle
1292: P\big(\zeta+2dt/(\zeta-a_0),\bar\zeta+2dt/(\bar\zeta-a_0),a_0
1293: +\sqrt\kappa dB_t\big)\rangle\,,
1294: \end{equation}
1295: where the average $\langle\ldots\rangle$\ is over all realisations
1296: of Brownian motion $dB_t$ up to time $dt$. Taylor expanding, using
1297: $\langle dB_t\rangle=0$ and $\langle(dB_t)^2\rangle=dt$, and
1298: equating the coefficient of $dt$ to zero gives
1299: \begin{equation}\label{eqn:pde1}
1300: \left({2\over\zeta-a_0}{\partial\over\partial\zeta}
1301: +{2\over\bar\zeta-a_0}{\partial\over\partial\bar\zeta}
1302: +\frac\kappa 2{\partial^2\over\partial a_0^2}\right)
1303: P\big(\zeta,\bar\zeta;a_0\big)=0\,.
1304: \end{equation}
1305: Thus $P$ satisfies a linear second-order partial differential equation,
1306: typical of conditional probabilities in stochastic differential equations.
1307: 
1308: By scale invariance $P$ in fact depends only on the angle $\theta$
1309: subtended between $\zeta-a_0$ and the real axis. Thus
1310: (\ref{eqn:pde1}) reduces to an ordinary second-order linear
1311: differential equation, which is in fact hypergeometric. The
1312: boundary conditions are that $P=0$ and $1$ when $\theta=\pi$ and
1313: $0$ respectively, which gives (specialising to $\kappa=6$)
1314: \begin{equation}\label{eqn:schrammformula}
1315: P=\frac12+{\Gamma(\ffrac23)\over\sqrt\pi\Gamma(\ffrac16)}
1316: (\cot\theta) {}_2F_1(\ffrac12,\ffrac23,\ffrac32;-\cot^2\theta)\,.
1317: \end{equation}
1318: Note that this may also be written in terms of a single quadrature
1319: since one solution of (\ref{eqn:pde1}) is $P=$ const.
1320: %
1321: \subsection{Crossing probability}
1322: Given a critical percolation problem inside a simply connected
1323: domain $\cal D$, what is the probability that a cluster connects
1324: two disjoint segments $AB$ and $CD$ of the boundary? This problem
1325: was conjectured to be conformally invariant and (probably) first
1326: studied numerically in \cite{Langlandsetal}. A formula based on
1327: CFT as well as a certain amount of guesswork was conjectured in
1328: \cite{JCcrossing}. It was proved, for the continuum limit of site
1329: percolation on the triangular lattice, by Smirnov\cite{Smirnov}.
1330: 
1331: Within SLE, it takes a certain amount of ingenuity\cite{Schramm}
1332: to relate this problem to a question about a single curve. As
1333: usual, let $\cal D$ be the upper half plane. It is always possible
1334: to make a fractional linear conformal mapping which takes $AB$
1335: into $(-\infty,x_1)$ and $CD$ into $(0,x_2)$, where $x_1<0$ and
1336: $x_2>0$. Now go back to the lattice picture and consider critical
1337: site percolation on the triangular lattice in the upper half
1338: plane, so that each site is independently coloured black or white
1339: with equal probabilities $\frac12$. Choose all the boundary sites
1340: on the positive real axis to be white, all those on the negative
1341: real axis to be black (see Fig.~\ref{fig:crossing}). There is a
1342: cluster boundary starting at the origin, which, in the continuum
1343: limit, will be described by SLE$_6$. Since $\kappa>4$, it
1344: repeatedly hits the real axis, both to the L and R of the origin.
1345: Eventually every point on the real axis is swallowed. Either $x_1$
1346: is swallowed before $x_2$, or vice versa.
1347: %
1348: \begin{figure}
1349: \centering
1350: \includegraphics[width=8cm]{crossing.eps}
1351: \caption{\label{fig:crossing}\small Is there a crossing on the
1352: white discs from $(0,x_2)$ to $(-\infty,x_1)$? This happens if and
1353: only if $x_1$ gets swallowed by the SLE before $x_2$.}
1354: \end{figure}
1355: %
1356: 
1357: Note that every site on the L of the curve is black, and every
1358: site on its R is white. Suppose that $x_1$ is swallowed before
1359: $x_2$. Then, at the moment it is swallowed, there exists a
1360: continuous path on the white sites, just to the R of the curve,
1361: which connects $(0,x_2)$ to the row just above $(-\infty,x_1)$. On
1362: the other hand, if $x_2$ is swallowed before $x_1$, there exists a
1363: continuous path on the black sites, just to the L of the curve,
1364: connecting $0-$ to a point on the real axis to the R of $x_2$.
1365: This path forms a barrier (as in the game of Hex) to the
1366: possibility of a white crossing from $(0,x_2)$ to $(-\infty,x_1)$.
1367: Hence there is such a crossing if and only if $x_1$ is swallowed
1368: before $x_2$ by the SLE curve.
1369: 
1370: Recall that in Sec.~\ref{sec:phases} we related the swallowing of
1371: a point $x_0$ on the real axis to the vanishing of
1372: $x_t=g_t(x_t)-a_t$, which undergoes a Bessel process on the real
1373: line. Therefore
1374: \begin{equation}
1375: {\rm Pr}\big({\rm crossing\ from\ }(0,x_2)\ {\rm to\
1376: }(-\infty,x_1)\big) ={\rm Pr}\big(x_{1t}\ {\rm  vanishes\ before\
1377: } x_{2t}\big)\,.
1378: \end{equation}
1379: 
1380: Denote this by $P\big(x_1,x_2\big)$. By generalising the SLE to
1381: start at $a_0$ rather than $0$, we can write a differential
1382: equation for this in similar manner to (\ref{eqn:pde1}):
1383: \begin{equation}\label{eqn:pde2}
1384: \left({2\over x_1-a_0}{\partial\over\partial x_1} +{2\over
1385: x_2-a_0}{\partial\over\partial x_2} +\frac\kappa
1386: 2{\partial^2\over\partial a_0^2}\right)\
1387: P\big(x_1,x_2;a_0\big)=0\,.
1388: \end{equation}
1389: Translational invariance implies that we can replace
1390: $\partial_{a_0}$ by $-(\partial_{x_1}+\partial_{x_2})$. Finally,
1391: $P$ can in fact depend only on the ratio
1392: $\eta=(x_2-a_0)/(a_0-x_1)$, which again reduces the equation to
1393: hypergeometric form. The solution is (specialising to $\kappa=6$
1394: for percolation)
1395: \begin{equation}\label{eqn:crossing}
1396: P={\Gamma(\ffrac23)\over\Gamma(\ffrac43)\Gamma(\ffrac13)}\eta^{1/3}
1397: {}_2F_1(\ffrac13,\ffrac23,\ffrac43;\eta)\,.
1398: \end{equation}
1399: It should be mentioned that this is but one of a number of
1400: percolation crossing formulae. Another, conjectured by
1401: Watts\cite{Watts}, for the probability that there is cluster which
1402: simultaneously connects $AB$ to $CD$ and $BC$ to $DA$, has since
1403: been proved by Dub\'edat\cite{Dubedat}. However, other formulae,
1404: for example for the mean number of distinct clusters connecting
1405: $AB$ and $CD$\cite{JCmean}, and for the probability that exactly
1406: $N$ distinct clusters cross an annulus\cite{JCannulus}, are as yet
1407: unproven using SLE methods.
1408: %
1409: \subsection{Critical exponents from SLE}\label{sec:sleexponents}
1410: Many of the critical exponents which have previously been
1411: conjectured by Coulomb gas or CFT methods may be derived
1412: rigorously using SLE, once the underlying postulates are assumed
1413: or proved. However SLE describes the measure on just a single
1414: curve, and in the papers of LSW a great deal of ingenuity has gone
1415: into showing how to relate this to all the other exponents. There
1416: is not space in this article to do these justice. Instead we
1417: describe three examples which give the flavour of the arguments,
1418: which initially may appear quite unconventional compared with the
1419: more traditional approaches.
1420: 
1421: \subsubsection{The fractal dimension of
1422: SLE}\label{sec:fractaldim} The fractal dimension of any
1423: geometrical object embedded in the plane can be defined roughly as
1424: follows: let $N(\epsilon)$ be the minimum number of small discs of
1425: radius $\epsilon$ required to cover the object. Then if
1426: $N(\epsilon)\sim\epsilon^{-d_f}$ as $\epsilon\to0$, $d_f$ is the
1427: fractal dimension.
1428: 
1429: One way of computing $d_f$ for a random curve $\gamma$ in the
1430: plane is to ask for the probability $P(x,y,\epsilon)$ that a given
1431: point $\zeta=x+iy$ lies within a distance $\epsilon$ of $\gamma$.
1432: A simple scaling argument shows that if $P$ behaves like
1433: $\epsilon^\delta f(x,y)$ as $\epsilon\to0$, then $\delta=2-d_f$.
1434: We can derive a differential equation for $P$ along the lines of
1435: the previous calculation. The only difference is that under the
1436: conformal mapping $g_{dt}$, $\epsilon\to
1437: |g_{dt}'(\zeta)|\epsilon\sim\big(1-2dt\,{\rm
1438: Re}(1/\zeta^2)\big)\epsilon$. The differential equation (written
1439: for convenience in cartesian coordinates) is
1440: \begin{equation}\label{eqn:pde3}
1441: \left({2x\over x^2+y^2}{\partial\over\partial x}-{2y\over
1442: x^2+y^2}{\partial\over\partial y} +\frac\kappa
1443: 2{\partial^2\over\partial x^2}-{2(x^2-y^2)\over(x^2+y^2)^2}
1444: \epsilon{\partial\over\partial\epsilon}\right)P=0\,.
1445: \end{equation}
1446: Now $P$ is dimensionless and therefore should have the form
1447: $(\epsilon/r)^{2-d_f}$ times a function of the polar angle
1448: $\theta$. In fact, the simple ansatz
1449: $P=\epsilon^{2-d_f}y^\alpha(x^2+y^2)^\beta$, with $\alpha+2\beta=
1450: d_f-2$ satisfies the equation. [The reason this works is connected
1451: with the simple form for the correlator
1452: $\langle\Phi_2\phi_{2,1}\phi_{2,1}\rangle$ discussed in
1453: Sec.~\ref{sec:CS}.] This gives $\alpha=(\kappa-8)^2/8\kappa$,
1454: $\beta=(\kappa-8)/2\kappa$ and
1455: \begin{equation}
1456: d_f=1+\kappa/8\,.
1457: \end{equation}
1458: This is correct for $\kappa\leq8$: otherwise there is another
1459: solution with $\alpha=\beta=0$ and $d_f=2$. A more careful
1460: statement and proof of this result can be found in \cite{Beffara}.
1461: 
1462: We see that the fractal dimension increases steadily from the
1463: value 1 when $\kappa=0$ (corresponding to a straight line) to a
1464: maximum value of 2 when $\kappa=8$. Beyond this value $\gamma$
1465: becomes space-filling: every point in the upper half plane lies on
1466: the curve.
1467: %
1468: \subsubsection{Crossing exponent}\label{sec:crossingexponent}
1469: Consider a critical percolation problem in the upper half plane.
1470: What is the asymptotic behaviour as $r\to\infty$ of the
1471: probability that the interval $(0,1)$ on the real axis is
1472: connected to the interval $(r,\infty)$? We expect this to decay as
1473: some power of $r$. The value of this exponent may be found by
1474: taking the appropriate limit of the crossing formula
1475: (\ref{eqn:crossing}), but instead we shall compute it directly. In
1476: order for there to be a crossing cluster, there must be two
1477: cluster boundaries which also cross between the two intervals, and
1478: which bound this cluster above and below. Denote the upper
1479: boundary by $\gamma$. Then we need to know the probability $P(r)$
1480: of there being another spanning curve lying between $\gamma$ and
1481: $(1,r)$, averaged over all realisations of $\gamma$. Because of
1482: the locality property, the measure on $\gamma$ is independent of
1483: the existence of the lower boundary, and is given by SLE$_6$
1484: conditioned not to hit the real axis along $(1,r)$. Note that
1485: because $\kappa>4$ it will eventually hit the real axis at some
1486: point to the right of $r$. For this reason we can do the
1487: computation for general $\kappa>4$, although it gives the actual
1488: crossing exponent only if $\kappa=6$.
1489: 
1490: Consider the behaviour of $P(r)$ under the conformal mapping $\hat
1491: g_{dt}(z)\sim z+(2dt/z)-\sqrt\kappa dB_t$ (which maps the growing
1492: tip $\tau_t$ into $0$.) The crossing probability should be
1493: conformally invariant and depend only on the ratio of the lengths
1494: of the two intervals, hence, by an argument which by now should be
1495: familiar,
1496: \begin{equation}
1497: P(r)=\langle P\big(\hat g_{dt}(r)/\hat g_{dt}(1)\big)\rangle\,.
1498: \end{equation}
1499: Expanding this out, remembering as usual that $(dB_t)^2=dt$, and
1500: setting to zero the $O(dt)$ term, we find for $r\gg1$
1501: \begin{equation}
1502: (\kappa-2)rP'(r)+\ffrac12\kappa r^2P''(r)=0\,,
1503: \end{equation}
1504: with the solution $P(r)\propto r^{-(\kappa-4)/\kappa}$ for
1505: $\kappa>4$. Setting $\kappa=6$ then gives the result $\frac13$.
1506: 
1507: \subsubsection{The one-arm exponent}\label{sec:onearm}
1508: Consider critical lattice percolation inside some finite region
1509: (for example a disc of radius $R$). What is the probability that a
1510: given site (e.g.~the origin) is connected to a finite segment $S$
1511: of the boundary? This should decay like $R^{-\lambda}$, where
1512: $\lambda$ is sometimes called the one-arm exponent. If we try to
1513: formulate this in the continuum, we immediately run up against the
1514: problem that all clusters are fractal with dimension $<2$, and so
1515: the probability of any given point being in any given cluster is
1516: zero. Instead, one may ask about the probability $P(r)$ that the
1517: cluster connected to $S$ gets within a distance $r$ of the origin.
1518: This should behave like $(r/R)^{\lambda}$. We can now set $R=1$
1519: and treat the problem using radial SLE$_6$.
1520: 
1521: Consider now a radial SLE$_\kappa$ which starts at
1522: $e^{i\theta_0}$. If $\kappa>4$ it will continually hit the
1523: boundary. Let $P(\theta-\theta_0,t)$ be the probability that the
1524: segment $(\theta_0,\theta)$ of the boundary has not been swallowed
1525: by time $t$. Then, by considering the evolution as usual under
1526: $g_{dt}$,
1527: \begin{equation}
1528: P(\theta,\theta_0,t)=\langle
1529: P(\theta+d\theta,\theta_0+d\theta_0,t-dt)\rangle\,,
1530: \end{equation}
1531: where $d\theta=\cot\big((\theta-\theta_0)/2\big)dt$ and
1532: $d\theta_0=\sqrt\kappa dB_t$. Setting $\theta_0=0$ and equating to
1533: zero the $O(dt)$ term, we find the time-dependent differential
1534: equation
1535: \begin{equation}\label{eqn:pde4}
1536: \partial_tP=\cot(\theta/2)\partial_\theta P+
1537: \ffrac12\kappa\partial^2_\theta P\,.
1538: \end{equation}
1539: This has the form of a backwards Fokker-Plank equation.
1540: 
1541: Now, since $g_t'(0)=e^t$, it is reasonable that, after time $t$,
1542: the SLE gets within a distance $O(e^{-t})$ of the origin.
1543: Therefore we can interpret $P$ as roughly the probability that the
1544: cluster connected to $(0,\theta)$ gets within a distance $r\sim
1545: e^{-t}$ of the origin. A more careful argument \cite{LSW1arm}
1546: confirms this. The boundary conditions are $P(0,t)=0$ as
1547: $\theta\to0$, and (with more difficulty) $\partial_\theta
1548: P(\theta,t)=0$ at $\theta=2\pi$. The solution may then be found by
1549: inspection to be
1550: \begin{equation}
1551: P\propto e^{-\lambda t}\big(\sin(\theta/4)\big)^{1-4/\kappa}\,,
1552: \end{equation}
1553: where $\lambda=(\kappa^2-16)/32\kappa$. For percolation this gives
1554: $\frac5{48}$, in agreement with Coulomb gas arguments
1555: \cite{Nienhuis}.
1556: 
1557: The appearance of differential operators such as that in
1558: (\ref{eqn:pde4}) will become clear from the CFT perspective in
1559: Sec.~\ref{sec:CS}. If instead of choosing Neuman boundary
1560: conditions at $\theta=2\pi$ we impose $P=0$, the same equation
1561: gives the bulk 2-leg exponent $x_2$, which is also related to the
1562: fractal dimension by $d_f=2-x_2$.
1563: %
1564: \newpage
1565: %
1566: \section{Relation to conformal field theory}\label{sec:cft}
1567: \subsection{Basics of CFT.}\label{sec:basics}
1568: The reader who already knows a little about CFT will have
1569: recognised the differential equations in Sec.~\ref{sec:calc} as
1570: being very similar in form to the BPZ equations\cite{BPZ}
1571: satisfied by the correlation functions of a $\phi_{2,1}$ operator,
1572: corresponding to a highest weight representation of the Virasoro
1573: algebra with a level 2 null state.
1574: 
1575: For those readers for whom the above paragraph makes no sense, and
1576: in any case to make the argument self-contained, we first
1577: introduce the concepts of (boundary) conformal field theory
1578: (BCFT.) We stress that these are heuristic in nature -- they serve
1579: only as a guide to formulating the basic principles of CFT which
1580: can then be developed into a mathematically consistent theory. For
1581: a longer introduction to BCFT see \cite{JCBCFT} and, for a
1582: complete account of CFT, \cite{YellowBook}.
1583: 
1584: We have at the back of our minds a euclidean field theory defined
1585: as a path integral over some set of fundamental fields
1586: $\{\psi(r)\}$. The partition function is $Z=\int
1587: e^{-S(\{\psi\})}[d\psi]$ where the action $S(\{\psi\})=\int_{\cal
1588: D}{\cal L}(\{\psi\})d^2\!r$ is an integral over a local lagrangian
1589: density. These fields may be thought of as smeared-out continuum
1590: versions of the lattice degrees of freedom. As in any field
1591: theory, this continuum limit involves renormalisation. There are
1592: so-called local scaling operators $\phi_j^{(0)}(r)$ which are
1593: particular functionals of the fundamental degrees of freedom,
1594: which have the property that we can define renormalised scaling
1595: operators $\phi_j(r)=a^{-x_j}\phi_j^{(0)}(r)$ whose correlators
1596: are finite in the continuum limit $a\to0$, that is
1597: \begin{equation}
1598: \lim_{a\to0}a^{-\sum_jx_j}\langle\phi_1^{(0)}(r_1)\ldots
1599: \phi_N^{(0)}(r_N)\rangle=\langle\phi_1(r_1)\ldots
1600: \phi_N(r_N)\rangle
1601: \end{equation}
1602: exists. The numbers $x_j$ are called the scaling dimensions, and
1603: are related to the various critical exponents. They are related to
1604: the conformal weights $(h_j,\bar h_j)$ by $x_j=h_j+\bar h_j$; the
1605: difference $h_j-\bar h_j=s_j$ is called the spin of $\phi_j$, and
1606: describes its behaviour under rotations. There are also \em
1607: boundary operators\em, localised on the boundary, which have only
1608: a single conformal weight equal to their scaling dimension.
1609: 
1610: The theory is developed independently of any particular set of
1611: fundamental fields or lagrangian. An important role in this is
1612: played by the stress tensor $T^{\mu\nu}(r)$, defined as the local
1613: response of the action to a change in the metric:
1614: \begin{equation}
1615: \delta S=(1/4\pi)\int_{\cal D}T^{\mu\nu}\delta g_{\mu\nu}d^2r\,.
1616: \end{equation}
1617: Invariance under local rotations and scale transformations usually
1618: implies that $T^{\mu\nu}$ is symmetric and traceless:
1619: $T^\mu_\mu=0$. This also implies invariance under conformal
1620: transformations, corresponding to $\delta g_{\mu\nu}\propto
1621: f(r)g_{\mu\nu}$.
1622: 
1623: In two dimensional flat space, infinitesimal coordinate
1624: transformations $r^\mu\to {r'}^\mu=r^\mu+\alpha^\mu(r)$ correspond
1625: to infinitesimal transformations of the metric with $\delta
1626: g^{\mu\nu}=-(\partial^\mu\alpha^\mu+
1627: \partial^\nu\alpha^\nu)$. It is important to
1628: realise that under these transformations the underlying lattice,
1629: or UV cut-off, is not transformed. Otherwise they would amount to
1630: a trivial reparametrisation. For a conformal transformation,
1631: $\alpha^\mu(r)$ is given by an analytic function: in complex
1632: coordinates $(z,\bar z)$, $\partial_{\bar z}\alpha^z =0$, so
1633: $\alpha^z\equiv\alpha(z)$ is holomorphic. However, such a function
1634: cannot be small everywhere (unless it is constant), so it is
1635: necessary to consider coordinate transformations which are not
1636: everywhere conformal.
1637: 
1638: Consider therefore two concentric semicircles $\Gamma_1$ and $\Gamma_2$
1639: in the upper half plane, centred on the origin, and of radii $R_1<R_2$.
1640: For $|r|<R_1$ let $\alpha^\mu$ be conformal, with $\alpha^z=\alpha(z)$,
1641: while for $|r|>R_2$ take $\alpha^\mu=0$. In between,
1642: $\alpha^\mu$ is not conformal, but is differentiable,
1643: so that
1644: $\delta S=(-1/2\pi)\int_{R_1<|r|<R_2}T^{\mu\nu}\partial_\mu\alpha_\nu d^2\!r$.
1645: This can be integrated by parts to give a term
1646: $(1/2\pi)\int_{R_1<|r|<R_2}\partial_\mu T^{\mu\nu}\alpha_\nu d^2\!r$ (which
1647: must vanish because $\alpha_\nu$ is arbitrary in this region, implying
1648: that $\partial_\mu T^{\mu\nu}=0$) and two surface terms. That on $\Gamma_2$
1649: vanishes because $\alpha^\mu=0$ there. We are left with
1650: \begin{equation}\label{eqn:deltaS}
1651: \delta S=(1/2\pi)\int_{\Gamma_1}T^{\mu\nu}\alpha_\mu
1652: \epsilon^{\nu\lambda}d\ell_\lambda\,,
1653: \end{equation}
1654: where $d\ell^\lambda$ is the line element along $\Gamma_1$.
1655: 
1656: The fact that $T^{\mu\nu}$ is conserved means, in complex
1657: coordinates, that $\partial_{\bar z}T_{zz}=\partial_zT_{\bar z\bar
1658: z}=0$, so that the correlations functions of $T(z)\equiv T_{zz}$
1659: are holomorphic functions of $z$, while those of $\overline
1660: T\equiv T_{\bar z\bar z}$ are antiholomorphic. (\ref{eqn:deltaS})
1661: may then be written
1662: \begin{equation}\label{eqn:deltaS2}
1663: \delta S=(1/2\pi i)\int_{\Gamma_1}T(z)\alpha(z)dz+{\rm c.c.}\,.
1664: \end{equation}
1665: 
1666: In any field theory with a boundary, it is necessary to impose
1667: some boundary condition. It can be argued that any translationally
1668: invariant boundary condition flows under the RG to conditions
1669: satisfying $T_{xy}=0$, which in complex coordinates means that
1670: $T=\overline T$ on the real axis. This means that the correlators
1671: of $\overline T$ are those of $T$ analytically continued into the
1672: lower half plane. The second term in (\ref{eqn:deltaS2}) may then
1673: be dropped if the contour in the first term is around a complete
1674: circle.
1675: 
1676: The conclusion of all this is that the effect of an infinitesimal
1677: conformal transformation on any correlator of observables inside
1678: $\Gamma_1$ is the same as inserting a contour integral $\int
1679: T(z)\alpha(z)dz/2\pi i$ into the correlator.
1680: 
1681: Another important element of CFT is the operator product expansion
1682: (OPE) of the stress tensor with other local operators. Since $T$
1683: is holomorphic, this has the form
1684: \begin{equation}\label{eqn:Tphi}
1685: T(z)\cdot\phi(0)=\sum_n z^{-n-2}\phi^{(n)}(0)\,,
1686: \end{equation}
1687: where the $\phi^{(n)}$ are (possibly new) local operators. By
1688: taking $\alpha(z)\propto z$ (corresponding to a scale
1689: transformation) we see that $\phi^{(0)}=h\phi$, where $h$ is its
1690: scaling dimension. Similarly, by taking $\alpha=$ const.,
1691: $\phi^{(-1)}=\partial_x\phi$. Local operators for which
1692: $\phi^{(n)}$ vanishes for $n\geq1$ are called primary. $T$ itself
1693: is not primary: its OPE with itself takes the form
1694: \begin{equation}\label{eqn:TT}
1695: T(z)\cdot T(0)=c/2z^4+(2/z^2)T(0)+(1/z)\partial_zT(0)+\cdots\,,
1696: \end{equation}
1697: where $c$ is the conformal anomaly number, ubiquitous in CFT. For
1698: example, the partition function on a long cylinder of length $L$
1699: and circumference $\ell$ behaves as $\exp(\pi cL/\ell)$,
1700: cf.~(\ref{eqn:c6}).
1701: %
1702: \subsection{Radial quantisation}
1703: This is the most important concept in understanding the link
1704: between SLE and CFT. We introduce it in the context of boundary
1705: CFT. As before, suppose there is some set of fundamental fields
1706: $\{\psi(r)\}$, with a Gibbs measure $e^{-S[ \psi]}[d\psi]$. Let
1707: $\Gamma$ be a semicircle in the upper half plane, centered on the
1708: origin. The Hilbert space of the BCFT is the function space (with
1709: a suitable norm) of field configurations $\{\psi_\Gamma\}$ on
1710: $\Gamma$.
1711: 
1712: The vacuum state is given by weighting each state
1713: $|\psi'_\Gamma\rangle$ by the (normalised) path integral
1714: restricted to the interior of $\Gamma$ and conditioned to take the
1715: specified values $\psi'_\Gamma$ on the boundary:
1716: \begin{equation}
1717: |0\rangle=\int[d\psi'_\Gamma]\int_{\psi_\Gamma=\psi'_\Gamma}
1718: [d\psi]\,e^{-S[\psi]}\,|\psi'_\Gamma\rangle\,.
1719: \end{equation}
1720: Note that because of scale invariance different choices of the
1721: radius of $\Gamma$ are equivalent, up to a normalisation factor.
1722: 
1723: Similarly, inserting a local operator $\phi(0)$ at the origin into
1724: the path integral defines a state $|\phi\rangle$. This is called
1725: the operator-state correspondence of CFT. If we also insert
1726: $(1/2\pi i)\int_Cz^{n+1}T(z)dz$, where $C$ lies inside $\Gamma$,
1727: we get a state $L_n|\phi\rangle$. The $L_n$ act linearly on the
1728: Hilbert space. From the OPE (\ref{eqn:Tphi}) we see that
1729: $L_n|\phi\rangle\propto|\phi^{(n)}\rangle$, and that, in
1730: particular, $L_0|\phi\rangle=h_\phi|\phi\rangle$. If $\phi$ is
1731: primary, $L_n|\phi\rangle=0$ for $n\geq1$. From the OPE
1732: (\ref{eqn:TT}) of $T$ with itself follow the commutation relations
1733: for the $L_n$
1734: \begin{equation}\label{eqn:vir}
1735: [L_n,L_m]=(n-m)L_{n+m}+\ffrac1{12}cn(n^2-1)\delta_{n,-m}\,,
1736: \end{equation}
1737: which are known as the Virasoro algebra. The state $|\phi\rangle$
1738: together with all its descendants, formed by acting on
1739: $|\phi\rangle$ an arbitrary number of times with the $L_n$ with
1740: $n\leq-1$, give a highest weight representation (where the weight
1741: is defined as the eigenvalue of $-L_0$.)
1742: 
1743: There is another way of generating such a highest weight
1744: representation. Suppose the boundary conditions on the negative
1745: and positive real axes are both conformal, that is they satisfy
1746: $T=\overline T$, but they are different. The vacuum with these
1747: boundary conditions gives a highest weight state which it is
1748: sometimes useful to think of as corresponding to the insertion of
1749: a `boundary condition changing' (bcc) operator at the origin. An
1750: example is the continuum limit of an Ising model in which the
1751: spins on the negative real axis are $-1$, and those on the
1752: positive axis are $+1$.
1753: 
1754: \subsection{Curves and states.}\label{sec:curves} In this section
1755: we describe a way of associating states in the Hilbert space of
1756: the BCFT with the growing curves of the Loewner process. This was
1757: first understood by M.~Bauer and D.~Bernard\cite{BB}, but we shall
1758: present the arguments slightly differently.
1759: 
1760: The boundary conditions associated with a bcc operator guarantee
1761: the existence, on the lattice, of a domain wall connecting the
1762: origin to infinity. Given a particular realisation $\gamma$, we
1763: can condition the Ising spins on its existence. We would like to
1764: be able to assume that this property continues to hold in the
1765: continuum limit: that is, we can condition the fields $\{\psi\}$
1766: on the existence of a such a curve. However, this involves
1767: conditioning on an event with probability zero: it turns out that
1768: in general the probability that, with respect to the measure in
1769: the path integral, the probability that a domain wall hits the
1770: real axis somewhere in an interval of length $\epsilon$ vanishes
1771: like $\epsilon^h$. In what follows we shall regard $\epsilon$ as
1772: small but fixed, and assume that the usual properties of SLE are
1773: applicable to this more general case.
1774: 
1775: Any such curve may be generated by a Loewner process: denote as
1776: before the part of the curve up to time $t$ by $\gamma_t$. The
1777: existence of this curve depends on only the field configurations
1778: $\psi$ in the interior of $\Gamma$, as long as $\gamma_t$ lies
1779: wholly inside this region. Then we can condition the fields
1780: contributing to the path integral on the existence of $\gamma_t$,
1781: thus defining a state
1782: \begin{equation}
1783: |\gamma_t\rangle =
1784: \int[d\psi'_\Gamma]\int_{\psi_\Gamma=\psi'_\Gamma;\gamma_t}
1785: [d\psi]\,e^{-S[\psi]}\,|\psi'_\Gamma\rangle\,.
1786: \end{equation}
1787: The path integral (over the whole of the upper half plane, not
1788: just the interior of $\Gamma$), when conditioned on $\gamma_t$,
1789: gives a measure $d\mu(\gamma_t)$ on these curves. The state
1790: \begin{equation}\label{eqn:ht}
1791: |h\rangle=|h_t\rangle\equiv\int d\mu(\gamma_t)|\gamma_t\rangle
1792: \end{equation}
1793: is in fact independent of $t$, since it is just given by the path
1794: integral conditioned on there being a curve connecting the origin
1795: to infinity, as guaranteed by the boundary conditions. In fact we
1796: see that $|h\rangle$ is just the state corresponding to a boundary
1797: condition changing operator at the origin.
1798: 
1799: However, $d\mu(\gamma_t)$ is also given by the measure on $a_t$ in
1800: Loewner evolution, through the iterated sequence of conformal
1801: mappings satisfying $d\hat g_t=2dt/\hat g_t-da_t$. This
1802: corresponds to an infinitesimal conformal mapping of the upper
1803: half plane minus $K_t$. As explained in the previous section,
1804: $d\hat g_t$ corresponds to inserting $(1/2\pi
1805: i)\int_C(2dt/z-da_t)T(z)dz$. In operator language, this
1806: corresponds to acting on $|\gamma_t\rangle$ with
1807: $2L_{-2}dt-L_{-1}da_t$ where $L_n=(1/2\pi i)\int_Cz^{n+1}T(z)dz$.
1808: Thus, for any $t_1<t$,
1809: \begin{equation}
1810: |g_{t_1}(\gamma_t)\rangle ={\bf
1811: T}\exp\left(\int_0^{t_1}\big(2L_{-2}dt'-L_{-1}da_{t'}\big)\right)
1812: |\gamma_t\rangle\,,
1813: \end{equation}
1814: where $\bf T$ denotes a time-ordered exponential.
1815: 
1816: The measure on $\gamma_t$ is the product of the measure of
1817: $\gamma_t\setminus\gamma_{t_1}$, conditioned on $\gamma_{t_1}$,
1818: with the unconditioned measure on $\gamma_{t_1}$. The first is the
1819: same as the unconditioned measure on $g_{t_1}(\gamma_t)$, and the
1820: second is given by the measure on $a_{t'}$ for $t'\in[0,t_1]$.
1821: Thus we can rewrite both the measure and the state in
1822: (\ref{eqn:ht}) as
1823: \begin{equation}
1824: |h_t\rangle=\int d\mu(g_{t_1}(\gamma_t))\int
1825: d\mu(a_{t';t'\in[0,t_1]}) {\bf
1826: T}e^{\int_{t_1}^0\big(2L_{-2}dt'-L_{-1}da_{t'}\big)}
1827: |g_{t_1}(\gamma_t)\rangle\,.
1828: \end{equation}
1829: For SLE, $a_t$ is proportional to a Brownian process. The
1830: integration over realisations of this for $t'\in[0,t_1]$ may be
1831: performed by breaking up the time interval into small segments of
1832: size $\delta t$, expanding out the exponential to $O(\delta t)$,
1833: using $(B_{\delta t})^2\approx \delta t$, and re-exponentiating.
1834: The result is
1835: \begin{equation}
1836: |h_t\rangle=\exp\left(-\big(2L_{-2}-(\kappa/2)L_{-1}^2\big)t_1\right)
1837: |h_{t-t_1}\rangle\,.
1838: \end{equation}
1839: But, as we argued earlier, $|h_t\rangle$ is independent of $t$,
1840: and therefore
1841: \begin{equation}
1842: \label{l2l1} \big(2L_{-2}-(\kappa/2)L_{-1}^2\big)|h\rangle=0\,.
1843: \end{equation}
1844: 
1845: This means that the descendant states $L_{-2}|h\rangle$ and
1846: $L^2_{-1}|h\rangle$ are linearly dependent. We say that the
1847: Virasoro representation corresponding to $|h\rangle$ has a null
1848: state at level 2. From this follow an number of important
1849: consequences. Acting on (\ref{l2l1}) with $L_1$ and $L_2$, and
1850: using the Virasoro algebra (\ref{eqn:vir}) and the fact that
1851: $L_1|h\rangle=L_2|h\rangle=0$ while $L_0|h\rangle=h|h\rangle$,
1852: leads to
1853: \begin{eqnarray}
1854: h&=&h_{2,1}={6-\kappa\over2\kappa}\,;\\
1855: c&=&{(3\kappa-8)(6-\kappa)\over2\kappa}\,.
1856: \end{eqnarray}
1857: These are the fundamental relations between the parameter $\kappa$
1858: of SLE and the data of CFT. The conventional notation $h_{2,1}$
1859: comes from the Kac formula in CFT which we do not need here. In
1860: fact this is appropriate to the case $\kappa<4$: for $\kappa>4$ it
1861: corresponds to $h_{1,2}$. (To further confuse the matter, some
1862: authors reverse the labels.) Note that the boundary exponent $h$
1863: parametrises the failure of locality in (\ref{eqn:kappa6}). From
1864: CFT we may also deduce that, with respect to the path integral
1865: measure, the probability that a curve connects small intervals of
1866: size $\epsilon$ about points $r_1$, $r_2$ on the real axis behaves
1867: like
1868: \begin{equation}
1869: \epsilon^{2h_{2,1}}\langle\phi_{2,1}(r_1)\phi_{2,1}(r_2)\rangle
1870: \propto\left({\epsilon\over|r_1-r_2|}\right)^{2h_{2,1}}\,.
1871: \end{equation}
1872: Such a result, elementary in CFT, is difficult to obtain directly
1873: from SLE in which the curves are conditioned to begin and end at
1874: \em given \em points.
1875: 
1876: Note that the central charge $c$ vanishes when either locality
1877: ($\kappa=6$) or restriction ($\kappa=\frac83$) hold. These cases
1878: correspond to the continuum limit of percolation and self-avoiding
1879: walks respectively, corresponding to formal limits $Q\to1$ in the
1880: Potts model and $n\to0$ in the $O(n)$ model for which the
1881: unconditioned partition function is trivial.
1882: 
1883: \subsection{Differential Equations}
1884: In this section we discuss how the linear second order
1885: differential equations for various observables which arise from
1886: the stochastic aspect of SLE follow equivalently from the null
1887: state condition in CFT. In this context they are known as the BPZ
1888: equations\cite{BPZ}. As an example consider Schramm's formula
1889: (\ref{eqn:schrammformula}) for the probability $P$ that a point
1890: $\zeta$ lies to the right of $\gamma$, or equivalently the
1891: expectation value of the indicator function ${\cal O}(\zeta)$
1892: which is 1 if this is satisfied and zero otherwise. In SLE, this
1893: expectation value is with respect to the measure on curves which
1894: connect the point $a_0$  to infinity. In CFT, as explained above,
1895: we can only consider curves which intersect some
1896: $\epsilon$-neighbourhood on the real axis. Therefore $P$ should be
1897: written as a ratio of expectation values with respect to the CFT
1898: measure:
1899: \begin{equation}
1900: P(\zeta;a_0)=\lim_{r_2\to\infty}{\langle\phi_{2,1}(a_0){\cal
1901: O}(\zeta)\phi_{2,1}(r_2)\rangle\over
1902: \langle\phi_{2,1}(a_0)\phi_{2,1}(r_2)\rangle}\,.
1903: \end{equation}
1904: We can derive differential equations for the correlators in the
1905: numerator and denominator by inserting into each of them a factor
1906: $(1/(2\pi i)\int_\Gamma\alpha(z)T(z)dz+$ c.c., where
1907: $\alpha(z)=2/(z-a_0)$, and $\Gamma$ is a small semicircle
1908: surrounding $a_0$. This is equivalent to making the infinitesimal
1909: transformation $z\to z+2\epsilon/(z-a_0)$. As before, the
1910: c.c.~term is equivalent to extending the contour in the first term
1911: to a full circle. The effect of this insertion may be evaluated in
1912: two ways: by shrinking the contour onto $a_0$ and using the OPE
1913: between $T$ and $\phi_{2,1}$ we get
1914: \begin{equation}\label{var1}
1915: 2L_{-2}\phi_{2,1}(a_0)=(\kappa/2)L_{-1}^2\phi_{2,1}(a_0)=
1916: \partial^2_{a_0}\phi_{2,1}(a_0)\,,
1917: \end{equation}
1918: while wrapping it about $\zeta$ (in a clockwise sense) we get
1919: \begin{equation}\label{var2}
1920: -\big(2/(\zeta-a_0)\big)\partial_\zeta{\cal
1921: O}-\big(2/(\bar\zeta-a_0)\big)\partial_{\bar\zeta}{\cal O}\,.
1922: \end{equation}
1923: The effect on $\phi_{2,1}(r_2)$ vanishes in the limit
1924: $r_2\to\infty$. As a result we can ignore the variation of the
1925: denominator in this case. Equating (\ref{var1}) and (\ref{var2})
1926: inside the correlation function in the numerator then leads to the
1927: differential equation (\ref{eqn:pde1}) for $P$ found in
1928: Sec.~\ref{sec:schrammformula}.
1929: 
1930: \subsubsection{Calagero-Sutherland model}\label{sec:CS}
1931: While many of the results of SLE may be re-derived in CFT with less
1932: rigour but perhaps greater simplicity, the latter contains much
1933: more information which is not immediately apparent from the SLE
1934: perspective. For example, one may consider correlation functions
1935: $\langle\phi_{1,2}(r_1)\phi_{1,2}(r_2)
1936: \ldots\phi_{2,1}(r_N)\ldots\rangle$ of multiple boundary condition
1937: changing operators with other operators either in the bulk or on
1938: the boundary. Evaluating the effect of an insertion $(1/2\pi
1939: i)\int_\Gamma T(z)dz/(z-r_j)$ where $\Gamma$ surrounds $r_j$ leads
1940: to a second order differential equation satisfied by the
1941: correlation function for \em each \em $j$.
1942: 
1943: This property is very powerful in the radial version. Consider the
1944: correlation function
1945: \begin{equation}
1946: C_\Phi(\theta_1,\ldots,\theta_N)=\langle\phi_{2,1}(\theta_1)
1947: \ldots\phi_{2,1}(\theta_N)\Phi(0)\rangle
1948: \end{equation}
1949: of $N$ $\phi_{2,1}$ operators on the boundary of the unit disc
1950: with a single bulk operator $\Phi$ at the origin. Consider the
1951: effect of inserting $(1/2 \pi i)\int_\Gamma\alpha_j(z)T(z)dz$
1952: where (cf. (\ref{eqn:radialsle}))
1953: \begin{equation}
1954: \alpha_j(z)=-z{z+e^{i\theta_j}\over z-e^{i\theta_j}}
1955: \end{equation}
1956: and $\Gamma$ surrounds the origin. Once again, this may be
1957: evaluated in two ways: by taking $\Gamma$ up to the boundary, with
1958: exception of small semicircles around the points $e^{i\theta_k}$,
1959: we get $G_jC_\Phi$, where $G_j$ is the second order differential
1960: operator
1961: \begin{equation}
1962: G_j=-\frac\kappa
1963: 2{\partial^2\over\partial\theta_j^2}+\frac{h_{2,1}}6+\frac c{12} -
1964: \sum_{k\not=j}\left(\cot{\theta_k-\theta_j\over2}
1965: {\partial\over\partial\theta_k}
1966: -{1\over2\sin^2(\theta_k-\theta_j)/2}h_{2,1}\right)\,.
1967: \end{equation}
1968: The first three terms come from evaluating the contour integral
1969: near $e^{i\theta_j}$, where $\alpha_j$ acts like
1970: $2L_{-2}-\frac16L_0-\frac{c}{12}$, (the term $\frac{c}{12}$ comes
1971: from the curvature of the boundary,) and the term with $k\not=j$
1972: from the contour near $e^{i\theta_k}$, where it acts like
1973: $\alpha_j(e^{i\theta_k})L_{-1}+{\rm
1974: Re}\,\alpha_j'(e^{i\theta_k})L_0$.
1975: 
1976: On the other hand, shrinking the contour down on the origin we see
1977: that $\alpha_j(z)=z+O(z^2)$, so that on $\Phi(0)$ it has the
1978: effect of $L_0+\overline L_0+\ldots$, where the omitted terms
1979: involve the $L_n$ and $\overline L_n$ with $n>0$. Assuming that
1980: $\Phi$ is primary, these other terms vanish, leaving simply
1981: $(L_0+\overline L_0)\Phi=x_\Phi\Phi$. Equating the two evaluations
1982: we find the differential equation
1983: \begin{equation}
1984: G_j\,C_\Phi=x_\Phi\,C_\Phi\,.
1985: \end{equation}
1986: 
1987: In general there is an $(N-1)$-dimensional space of independent
1988: differential operators $G_j$ with a common eigenfunction $C_\Phi$.
1989: (There is one fewer dimension because they all commute with the
1990: total angular momentum $\sum_j(\partial/\partial\theta_j)$.) For
1991: the case $N=2$, setting $\theta=\theta_2-\theta_1$, we recognise
1992: the differential operator in Sec.~\ref{sec:onearm}.
1993: 
1994: In general these operators are not self-adjoint and their spectrum
1995: is difficult to analyse. However, if we form the equally-weighted
1996: linear combination $G\equiv \sum_{j=1}^NG_j$, the terms with a
1997: single derivative may be written in the form $\sum_k (\partial
1998: V/\partial\theta_k)(\partial/\partial\theta_k)$ where $V$ is a
1999: potential function. In this case it is well known from the theory
2000: of the Fokker-Plank equation that $G$ is related by a similarity
2001: transformation to a self-adjoint operator. In fact\cite{JCCS} if
2002: we form $|\Psi_N|^{2/\kappa}G|\Psi_N|^{-2/\kappa}$ where $\Psi_N=
2003: \prod_{j<k}\big(e^{i\theta_j}-e^{\theta_k}\big)$ is the
2004: `free-fermion' wave function on the circle, the result is, up to
2005: calculable constants the well-known $N$-particle
2006: Calogero-Sutherland hamiltonian
2007: \begin{equation}
2008: H_N(\beta)=-\frac12\sum_{j=1}^N{\partial^2\over\partial\theta_j^2}
2009: +{\beta(\beta-2)\over16}\sum_{j<k}{1\over\sin^2(\theta_j-\theta_k)/2}\,,
2010: \end{equation}
2011: with $\beta=8/\kappa$. It follows that the scaling dimensions of
2012: bulk operators like $\Phi$ are simply related to eigenvalues
2013: $\Lambda_N$ of $H_N$ by
2014: \begin{equation}
2015: x_\Phi=(\kappa/N)\Lambda_N(8/\kappa)-(4/\kappa N)E_N^{\rm
2016: ff}+\ffrac16h_{2,1}+\ffrac1{12}c\,,
2017: \end{equation}
2018: where $E_N^{\rm ff}=\frac1{24}N(N^2-1)$. Similarly $C_\Phi$ is
2019: proportional to the corresponding eigenfunction. In fact the
2020: ground state (with conventional boundary conditions) turns out to
2021: correspond to the bulk $N$-leg operator discussed in
2022: Sec.~\ref{Nleg1}. The corresponding correlator is
2023: $|\Psi_N|^{2/\kappa}$.
2024: 
2025: 
2026: %
2027: \newpage
2028: %
2029: \section{Related ideas}\label{sec:relatedideas}
2030: \subsubsection{Multiple SLEs}\label{sec:multiplesles}
2031: We pointed out earlier that the boundary operators $\phi_{2,1}$
2032: correspond to the continuum limits of lattice curves which hit
2033: the boundary at a given point. For a single curve, these are
2034: described by SLE, and we have shown in that case how the resulting
2035: differential equations also appear in CFT. Using the $N$-particle
2036: generalisation of the CFT results of the previous section, we may
2037: now `reverse engineer' the problem and conjecture the
2038: generalisation of SLE to $N$ curves.
2039: 
2040: The expectation value of some observable $\cal O$ given that $N$
2041: curves, starting at the origin, hit the boundary at
2042: $(\theta_1,\ldots,\theta_N)$ is
2043: \begin{equation}
2044: P_{\cal O}(\theta_1,\ldots,\theta_N) = {F_{\cal
2045: O}(\theta_1,\ldots,\theta_N)\over F_{\bf
2046: 1}(\theta_1,\ldots,\theta_N)}\,,
2047: \end{equation}
2048: where $F_{\cal O}=\langle{\cal O}\,\phi_{2,1}(e^{i\theta_1})
2049: \ldots \phi_{2,1}(e^{i\theta_N})\Phi_N(0)\rangle$. This satisfies
2050: the BPZ equation
2051: \begin{equation}
2052: G_j\,F_{\cal O}=\langle(\delta_j{\cal O}\phi_{2,1}(e^{i\theta_1})
2053: \ldots \phi_{2,1}(e^{i\theta_N})\Phi_N(0)\rangle\,,
2054: \end{equation}
2055: where $\delta_j{\cal O}$ is the variation in $\cal O$ under
2056: $\alpha_j$. If we now write $F_{\cal O}=F_{\bf 1}\cdot P_{\cal O}$
2057: and use the fact that $G_jF_{\bf 1}=x_\Phi F_{\bf 1}$, we find a
2058: relatively simple differential equation for $P_{\cal O}$, since
2059: the non-derivative terms in $G_j$ cancel. There is also a
2060: complication since the second derivative gives a cross term
2061: proportional to $\big(\partial_{\theta_j}F_{\bf 1}\big)
2062: \big(\partial_{\theta_j}P_{\cal O}\big)$. However, this may be
2063: evaluated from the explicit form $F_{\bf 1}=|\Psi_N|^{2/\kappa}$.
2064: The result is
2065: \begin{equation}
2066: \left(\frac\kappa2{\partial^2\over\partial\theta_j^2} +
2067: \sum_{k\not=j}\cot{\theta_k-\theta_j\over2}
2068: \left({\partial\over\partial\theta_k}-
2069: {\partial\over\partial\theta_j}\right)\right)P_{\cal O}=
2070: \delta_jP_{\cal O}\,,
2071: \end{equation}
2072: where the right hand side comes from the variation in $\cal O$.
2073: 
2074: The left hand side may be recognised as the generator (the adjoint
2075: of the Fokker-Planck operator) for the stochastic process
2076: \begin{eqnarray}
2077: d\theta_j&=&\sqrt\kappa
2078: dB_t+\sum_{k\not=j}(\rho_k/2)\cot\big((\theta_j-\theta_k)/2\big)\,
2079: dt\,;\label{mult1}\\
2080: d\theta_k&=&\cot\big((\theta_k-\theta_j)/2\big)\,dt\,,\label{mult2}
2081: \end{eqnarray}
2082: where $\rho_k=2$. [For general values of the parameters $\rho_k$
2083: this process is known as (radial) SLE$(\kappa,\vec\rho\,)$,
2084: although this is more usually considered in the chordal version.
2085: It has been argued\cite{JCSLEkr} that this applies to the level
2086: lines of a free gaussian field with piecewise constant Dirichlet
2087: boundary conditions: the parameters $\rho_k$ are related to the
2088: size of the discontinuities at the points $e^{i\theta_k}$.
2089: SLE$(\kappa,\vec\rho\,)$ has also been used to give examples of
2090: restriction measures on curves which are not reflection
2091: symmetric\cite{SLEkrrefs}.]
2092: 
2093: We see that $e^{i\theta_j}$ undergoes Brownian motion but is also
2094: repelled by the other particles at $e^{i\theta_k}$ ($k\not=j$):
2095: these particles are themselves repelled deterministically from
2096: $e^{i\theta_j}$. The infinitesimal transformation $\alpha_j$
2097: corresponds to the radial Loewner equation
2098: \begin{equation}\label{mult3}
2099: {dg_{j,t}\over dt}=-g_{j,t}\,{g_{j,t}+e^{i\theta_{j,t}} \over
2100: g_{j,t}-e^{i\theta_{j,t}}}\,.
2101: \end{equation}
2102: 
2103: The conjectured interpretation of this is as follows: we have $N$
2104: non-intersecting curves connecting the boundary points
2105: $e^{i\theta_{k,0}}$ to the origin. The evolution of the $j$th
2106: curve in the presence of the others is given by the radial Loewner
2107: equation with, however, the driving term not being simple Brownian
2108: motion but instead the more complicated process
2109: (\ref{mult1},\ref{mult2}).
2110: 
2111: However, from the CFT point of view we may equally well consider
2112: the linear combination $\sum_jG_j$. The Loewner equation is now
2113: \begin{equation}\label{mult4}
2114: \dot g_t=-g_t\sum_{j=1}^N{g_t+e^{i\theta_{j,t}} \over
2115: g_t-e^{i\theta_{j,t}}}\,,
2116: \end{equation}
2117: where
2118: \begin{equation}\label{mult5}
2119: d\theta_j=\sqrt\kappa
2120: dB_t^j+2\sum_{k\not=j}\cot\big((\theta_j-\theta_k)/2\big)\,dt\,.
2121: \end{equation}
2122: This is known in the theory of random matrices as Dyson's Brownian
2123: motion. It describes the statistics of the eigenvalues of unitary
2124: matrices. The conjectured interpretation is now in terms of $N$
2125: random curves which are all growing in each other's mutual
2126: presence at the \em same \em mean rate (measured in Loewner time).
2127: From the point of view of SLE, it is by no means obvious that the
2128: measure on $N$ curves generated by process
2129: (\ref{mult1},\ref{mult2},\ref{mult3}) is the same as that given by
2130: (\ref{mult4},\ref{mult5}). However CFT suggests that, for curves
2131: which are the continuum limit of suitable lattice models, this is
2132: indeed the case.
2133: \subsection{Other variants of SLE}\label{sec:variants}
2134: So far we have discussed only chordal SLE, which describes curves
2135: connecting distinct points on the boundary of a simple connected
2136: domain, and radial SLE, in which the curve connects a boundary
2137: point to an interior point. Another simple variant is dipolar
2138: SLE\cite{dipolar}, in which the curve is constrained to start at
2139: boundary point and to end on some finite segment of the boundary
2140: not containing the point. The canonical domain is an infinitely
2141: long strip, with the curve starting a point on one edge and ending
2142: on the other edge. This set-up allows the computation of several
2143: interesting physical quantities.
2144: 
2145: The study of SLE in multiply-connected domains is very
2146: interesting. Their conformal classes are characterised by a set of
2147: moduli, which change as the curve grows. R.~Friedrich and
2148: co-workers\cite{BFriedrich} have argued that SLE in such a domain
2149: is characterised by diffusion in moduli space as well as diffusion
2150: on the boundary.
2151: 
2152: It is possible to rewrite the differential equations which arise
2153: from null state conditions in extended CFTs (for example
2154: super-conformal CFTs\cite{Ras} and WZWN models\cite{GLW}) in terms
2155: of the generators of stochastic conformal mappings which
2156: generalise that of Loewner. However, a physical interpretation in
2157: terms of the continuum limit of lattice curves appears so far to
2158: be missing.
2159: 
2160: \subsection{Other growth models}
2161: SLE is in fact just one very special, solvable, example of an
2162: approach to growth processes in two dimensions using conformal
2163: mappings which has been around for a number of years. For a recent
2164: review see \cite{BBrev}. The prototypical problem of this type is
2165: diffusion-limited aggregation (DLA). In this model of cluster
2166: formation, particles of finite radius diffuse in, one by one, from
2167: infinity until they hit the existing cluster, where they stick.
2168: The probability of sticking at a given point is proportional to
2169: the local electric field, if we imagine the cluster as being
2170: charged. The resultant highly branched structures are very similar
2171: to those observed in smoke particles, and in viscous fingering
2172: experiments where one fluid is forced into another in which it is
2173: immiscible. Hastings and Levitov\cite{HL} proposed an approach to
2174: this problem using conformal mappings. At each time $t$, the
2175: boundary of the cluster is described by the conformal mapping
2176: $f_t(z)$ which takes it to the unit disc. The cluster is grown by
2177: adding a small semicircular piece to the boundary. The way this
2178: changes $f_t$ is well known according to a theorem of Hadamard.
2179: The difficulty is that the probability of adding this piece at a
2180: given point depends on the local electric field which itself
2181: depends on $f_t'$. The equation of motion for $f_t$ is therefore
2182: more complicated than in SLE. Moreover it may be shown that almost
2183: all initially smooth boundary curves evolve towards a finite-time
2184: singularity: this is thought to be responsible for branching, but
2185: just at this point the equations must be regularised to reflect
2186: the finite size of the particles (or, in viscous fingering, the
2187: effects of finite surface tension.)
2188: 
2189: It is also possible to generate branching structures by making the
2190: driving term $a_t$ in Loewner's equation discontinuous, for
2191: example taking it to be a Levy process. Unfortunately this does
2192: not appear to describe a physically interesting model.
2193: 
2194: Finally, Hastings\cite{Hastings} has proposed two related growth
2195: models which each lead, in the continuum limit, to SLE. These are
2196: very similar to DLA, except that growth is only allowed at the
2197: tip. The first, called the arbitrary Laplacian random walk, takes
2198: place on the lattice. The tip moves to one of the neighbouring
2199: unoccupied sites $r$ with relative probability $E(r)^\eta$, where
2200: $E(r)$ is the lattice electric field, that is the potential
2201: difference between the tip and $r$, and $\eta$ is a parameter. The
2202: second growth model takes place in the continuum via iterated
2203: conformal mappings, in which pieces of length $\ell_1$ are added
2204: to the tip, but shifted to the left or right relative to the
2205: previous growth direction by a random amount $\pm\ell_2$. This
2206: model depends on the ratio $\ell_2/\ell_1$, and leads, in the
2207: continuum limit, to SLE$_\kappa$ with $\kappa=\ell_2/4\ell_1$. For
2208: the lattice model there is no universal relation between $\kappa$
2209: and $\eta$, except for $\eta=1$, which is the same as the
2210: loop-erased random walk (Sec.~\ref{sec:Onmodel}) and converges to
2211: SLE$_2$.
2212: %
2213: \begin{thebibliography}{99}
2214: %
2215: \bibitem{JCcrossing} J.~Cardy, {\sl Critical percolation in finite
2216: geometries}, J. Phys. A {\bf 25}, L201, 1992 (hep-th/9111026).
2217: %
2218: \bibitem{Lang} R.~Langlands, P.~Pouliot and Y.~Saint-Aubin,
2219: {\sl Conformal invariance and two-dimensional percolation,}
2220: Bulletin of the AMS {\bf 30}, 1, 1994 (math.MP/9401222).
2221: %
2222: \bibitem{Nienhuis} B.~Nienhuis, {\sl Coulomb gas formulation of
2223: two-dimensional phase transitions,} in {\sl Phase transitions and
2224: critical phenomena}, vol. 11, C.~Domb and J.L.~Lebowitz, eds.
2225: (Academic, 1987.)
2226: %
2227: \bibitem{Loewner} K.~L\"owner, {\sl Untersuchungen \"uber schlichte
2228: konforme Abbildungen des Einheitskreises,} I, Math. Ann. {\bf 89},
2229: 103, 1923.
2230: %
2231: \bibitem{Schramm} O.~Schramm, {\sl Scaling limits of loop-erased random
2232: walks and uniform spanning trees,} Israel J. Math. {\bf 118}, 221,
2233: 2000 (math.PR/9904022).
2234: %
2235: \bibitem{LSW} G.F.~Lawler, O.~Schramm and W.~Werner, {\sl Values of
2236: Brownian intersection exponents I: Half-plane exponents,} Acta
2237: Math. {\bf 187(2)}, 237, 2001 (math.PR/9911084); {\sl II: Plane
2238: exponents,} \em ibid.\em  {\bf 187(2)}, 275, 2001
2239: (math.PR/0003156); {\sl III: Two-sided exponents,} Ann. Inst. H.
2240: Poincar\'e Statist. {\bf 38(1)}, 109, 2002 (math.PR/0005294).
2241: %
2242: \bibitem{Smirnov} S.~Smirnov, {\sl Critical percolation in the plane:
2243: conformal invariance, Cardy's formula, scaling limits,} C. R.
2244: Acad. Sci. Paris S\'er. I Math. {\bf 333(3)}, 239, 2001; a longer
2245: version is available at http://www.math.kth.se/$\sim$stas/papers.
2246: %
2247: \bibitem{SmirnovWerner} S.~Smirnov and W.~Werner, {\sl Critical
2248: exponents for two-dimensional percolation,} Math. Res. Lett. {\bf
2249: 8}, 729, 2001; (math.PR/0109120).
2250: %
2251: \bibitem{JCsurf} J.~Cardy, {\sl Conformal Invariance and Surface Critical
2252: Behavior,} Nucl. Phys. B {\bf 240}, 514, 1984.
2253: %
2254: \bibitem{BB} M.~Bauer and D.~Bernard, {\sl Conformal field theories
2255: of stochastic Loewner evolutions}, Comm. Math. Phys. {\bf 239},
2256: 493, 2003 (hep-th/0210015); {\sl SLE$_\kappa$ growth processes and
2257: conformal field theory,} Phys. Lett. B {\bf 543}, 135, 2002
2258: (math-ph/0206028); {\sl SLE martingales and the Virasoro algebra,}
2259: Phys. Lett. B {\bf 557}, 309, 2003 (hep-th/0301064); {\sl CFTs of
2260: SLES: the radial case,} Phys. Lett. B {\bf 583}, 324, 2004
2261: (math-ph/0310032); {\sl Conformal transformations and the SLE
2262: partition function martingale,} Ann. Henri Poincar\'e {\bf 5},
2263: 289, 2004 (math-ph/0305061).
2264: %
2265: \bibitem{FriedrichWerner} R.~Friedrich and W.~Werner, {\sl Conformal
2266: fields, restriction properties, degenerate representations and
2267: SLE,} C.R. Acad. Sci. Paris, Ser. I Math. {\bf 335}, 947
2268: (math.PR/0209382); {\sl Conformal restriction, highest-weight
2269: representations and SLE,} Comm. Math. Phys. {\bf 243}, 105, 2003
2270: (math-ph/0301018).
2271: %
2272: \bibitem{Wreview} W.~Werner, {\sl Random planar curves and
2273: Schramm-Loewner evolutions,} in {\sl Ecole d'Et\'e de
2274: Probabilit\'es de Saint-Flour XXXII (2002)}, Springer Lecture
2275: Notes in Mathematics {\bf 1180}, 113, 2004 (math.PR/0303354).
2276: %
2277: \bibitem{Lreview} G.F.~Lawler, {\sl Conformally invariant processes in the
2278: plane,} ICTP Lecture Notes {\bf 17}, 305, 2004
2279: (http://www.ictp.trieste.it/$\sim$pub\_off/lectures/vol17.html);
2280: {\sl Stochastic Loewner Evolution,} to appear in {\sl Encyclopedia
2281: of Mathematical Physics}, J.-P.~Fran\c coise, G.~Naber and
2282: T.S.~Tsun, eds. (Elsevier, 2005.)
2283: (http://www.math.cornell.edu/~lawler/encyclopedia.ps); {\sl
2284: Conformally Invariant Processes in the Plane}, book, (American
2285: Math. Soc., 2005).
2286: %
2287: \bibitem{KNreview} W.~Kager and B.~Nienhuis, {\sl A guide to stochastic
2288: Loewner evolution and its applications,} J. Stat. Phys. {\bf 115},
2289: 1149, 2004 (math-ph/0312251).
2290: %
2291: \bibitem{KadGruz} I.A.~Gruzberg and L.P.~Kadanoff, {\sl The Loewner
2292: equation: maps and shapes,} J. Stat. Phys. {\bf 114}, 1183, 2004
2293: (cond-mat/0309292).
2294: %
2295: \bibitem{Dupreview} B.~Duplantier, {\sl Conformal fractal geometry and
2296: boundary quantum gravity,} math-ph/0303034.
2297: %
2298: \bibitem{Kondev} J.~Kondev, {\sl Liouville theory of fluctuating
2299: loops,}
2300: Phys. Rev. Lett. {\bf 78}, 4320, 1997 (cond-mat/9703113).
2301: %
2302: \bibitem{KadKager} W.~Kager, B.~Nienhuis and L.P.~Kadanoff, {\sl Exact
2303: solutions for Loewner evolutions,} J. Stat. Phys. {\bf 115}, 805,
2304: 2004 (math-ph/0309006).
2305: %
2306: \bibitem{RS} S.~Rohde and O.~Schramm, {\sl Basic properties of
2307: SLE,}
2308: Ann. Math., to appear (math.PR/0106036).
2309: %
2310: \bibitem{Dupduality} B.~Duplantier, {\sl Conformally invariant fractals
2311: and potential theory,} Phys. Rev. Lett. {\bf 84}, 1363, 2000
2312: (cond-mat/9908314).
2313: %
2314: \bibitem{Beffaraduality} V.~Beffara, {\sl Hausdorff dimensions for
2315: SLE$_6$,} Ann. Probab. {\bf 32}, 2606, 2004 (math.PR/0204208).
2316: %
2317: \bibitem{LSWSAW} G.F.~Lawler, O.~Schramm and W.~Werner, {\sl On the
2318: scaling limit of the planar self-avoiding walk,} to appear in {\sl
2319: Fractal geometry and applications, a jubilee of Benoit
2320: Mandelbrot}, AMS Proc. Symp. Pure Math. (math.PR/02044277).
2321: %
2322: \bibitem{SSHE} O.~Schramm and S.~Sheffield, {\sl The harmonic explorer
2323: and its convergence to SLE$_4$} (math.PR/0310210).
2324: %
2325: \bibitem{LSWUST} G.F.~Lawler, O.~Schramm and W.~Werner, {\sl Conformal
2326: invariance of planar loop-erased random walks and uniform spanning
2327: trees,} Ann. Probab. {\bf 32}, 939, 2004 (math.PR/0112234).
2328: %
2329: \bibitem{Schrammformula} O.~Schramm, {\sl A percolation formula,}
2330: Electronic Comm. Probab. {\bf 8}, paper no. 12, 2001
2331: (math.PR/0107096).
2332: %
2333: \bibitem{Langlandsetal} R.P.~Langlands, C.~Pichet, Ph.~Pouliot and
2334: Y.~Saint-Aubin, {\sl On the universality of crossing probabilities
2335: in two-dimensional percolation,} J. Stat. Phys. {\bf 67}, 553,
2336: 1992.
2337: %
2338: \bibitem{Watts} G.~Watts, {\sl A crossing probability for percolation in two dimensions},
2339: J. Phys. A {\bf 29}, L363, 1996 (cond-mat/9603167).
2340: %
2341: \bibitem{Dubedat} J.~Dub\'edat, {\sl Excursion decompositions for SLE$_6$ and
2342: Watts' crossing formula} (math.PR/0405074).
2343: %
2344: \bibitem{JCmean} J.~Cardy, {\sl Linking numbers for self-avoiding loops and percolation:
2345: application to the spin quantum Hall transition,} Phys. Rev. Lett.
2346: {\bf 84}, 3507, 2000 (cond-mat/9911457).
2347: %
2348: \bibitem{JCannulus} J.~Cardy, {\sl Crossing Formulae for Critical Percolation in an
2349: Annulus,} J. Phys. A {\bf 35}, L565, 2002 (math-ph/0208019).
2350: %
2351: \bibitem{Beffara} V.~Beffara, {\sl The dimension of the SLE
2352: curves} (math.PR/0211322).
2353: %
2354: \bibitem{LSW1arm} G.F.~Lawler, O.~Schramm and W.~Werner, {\sl One-arm
2355: exponent for critical 2D percolation,} Electronic J. Probab. {\bf
2356: 7}, paper no. 2, 2002 (math.PR/0108211).
2357: %
2358: \bibitem{BPZ} A.A.~Belavin, A.M.~Polyakov and A. B. Zamolodchikov,
2359: J. Stat. Phys. {\bf 34}, 763, 1984.
2360: %
2361: \bibitem{JCBCFT} J.~Cardy, {\sl Boundary Conformal Field Theory,} to appear in
2362: {\sl Encyclopedia of Mathematical Physics}, J.-P.~Fran\c coise,
2363: G.~Naber and T.S.~Tsun, eds. (Elsevier, 2005) (hep-th/0411189).
2364: %
2365: \bibitem{YellowBook} P.~Di Francesco, P.~Mathieu and D.~Senechal,
2366: {\sl Conformal field theory} (Springer, 1997).
2367: %
2368: \bibitem{JCCS} J.~Cardy, {\sl Calogero-Sutherland Model and Bulk-Boundary Correlations in
2369: Conformal Field Theory,} Phys. Lett. B {\bf 582}, 121, 2004
2370: (hep-th/0310291).
2371: %
2372: \bibitem{JCSLEkr} J.~Cardy, {\sl SLE$(\kappa,\rho)$ and Conformal Field
2373: Theory} (math-ph/0412033).
2374: %
2375: \bibitem{SLEkrrefs} J.~Dub\'edat, {\sl SLE$(\kappa,\rho)$ martingales and
2376: duality}, Ann. Probab. {\bf 33}, 223, 2005 (math.PR/0303128);
2377: W.~Werner, {\sl Girsanov's transformation for SLE$(\kappa,\rho)$
2378: processes, intersection exponents and hiding exponents,} Ann. Fac.
2379: Sci. Toulouse (6) {\bf 13}, 121, 2004 (math.PR/0302115).
2380: %
2381: \bibitem{dipolar} M.~Bauer, D.~Bernard and J.~Houdayer {\sl Dipolar SLEs}
2382: (math-ph/0411038).
2383: %
2384: \bibitem{BFriedrich} R.~Friedrich and J.~Kalkkinen, {\sl On conformal
2385: field theory and stochastic Loewner evolution,} Nucl. Phys. B {\bf
2386: 687}, 279, 2004 (hep-th/03087023); R.~Friedrich, {\sl On
2387: connections of conformal field theory and stochastic Loewner
2388: evolution} (math-ph/0410029); R.O.~Bauer and R.~Friedrich, {\sl On
2389: chordal and bilateral SLE in multiply connected domains}
2390: (math.PR/0503178).
2391: %
2392: \bibitem{Ras} J.~Nagi and J.~Rasmussen, {\sl On stochastic evolutions
2393: and super-conformal field theory,} Nucl. Phys. B {\bf 704}, 475,
2394: 2005 (math-ph/0407049).
2395: %
2396: \bibitem{GLW} J.~Rasmussen, {\sl On SU$(2)$
2397: Wess-Zumino-Witten models and stochastic evolutions}
2398: (hep-th/0409026); E.~Bettelheim, I.~Gruzberg, A.W.W.~Ludwig and
2399: P.~Wiegmann, {\sl Stochastic Loewner evolution for conformal field
2400: theories with Lie group symmetries} (hep-th/0503013).
2401: %
2402: \bibitem{BBrev} M.~Bauer and D.~Bernard, {\sl Loewner Chains}
2403: (cond-mat/0412372).
2404: %
2405: \bibitem{HL} M. Hastings and L.S.~Levitov, {\sl Laplacian growth as one-dimensional
2406: turbulence,} Physica D {\bf 116}, 244, 1998 (cond-mat/9607021).
2407: %
2408: \bibitem{Hastings} M.~Hastings, {\sl Exact multi-fractal spectra for
2409: arbitrary laplacian random walks,} Phys. Rev. Lett. {\bf 88},
2410: 055506, 2002 (cond-mat/0109304).
2411: %
2412: \end{thebibliography}
2413: \end{document}
2414: