1:
2: \documentclass{elsart}
3: \journal{Nuclear Physics B}
4: \usepackage{graphicx}
5: \usepackage{amsmath}
6:
7: \newcommand{\la}{\label}
8: \newcommand{\be}{\begin{equation}}
9: \newcommand{\ee}{\end{equation}}
10: \newcommand{\bea}{\begin{eqnarray}}
11: \newcommand{\eea}{\end{eqnarray}}
12:
13: \def\la{\label}
14: \def\be{\begin{equation}}
15: \def\beq{\begin{equation}}
16: \def\eeq{\end{equation}}
17: \def\ee{\end{equation}}
18: \def\bea{\begin{eqnarray}}
19: \def\eea{\end{eqnarray}}
20: \def\p{\partial}
21:
22: \begin{document}
23:
24: \begin{frontmatter}
25:
26: \title{Normal random matrix ensemble as a growth problem}
27:
28: \author[la]{R.~Teodorescu \thanksref{now}},
29: \ead{rteodore@uchicago.edu}
30: \author[lb]{E.~Bettelheim \thanksref{now1}},
31: \ead{eldadb@phys.huji.ac.il}
32: \author[lb]{O.~Agam},
33: \author[lc]{A.~Zabrodin},
34: \ead{zabrodin@itep.ru} and
35: \author[ld]{P.~Wiegmann}
36: \ead{wiegmann@uchicago.edu}
37: \thanks[now]{Present address: Physics Department, Columbia University, New York.}
38: \thanks[now1]{Present address: James Frank Institute, University of Chicago, 5640 S. Ellis Ave. Chicago, IL 60637, USA.}
39: \address[la]{James Frank Institute, University of Chicago, 5640 S. Ellis Ave. Chicago, IL
40: 60637, USA.}
41: \address[lb]{Racah Institute of Physics, Hebrew University, Givat Ram, Jerusalem, Israel
42: 91904}
43: \address[lc]{Institute of Biochemical Physics, Kosygina str. 4, 117334 Moscow, Russia also at ITEP, Bol. Cheremushkinskaya str. 25,
44: 117259 Moscow, Russia}
45: \address[ld]{James Frank Institute, Enrico Fermi Institute, University of Chicago, 5640 S. Ellis Ave. Chicago, IL and the Landau
46: Institute, Moscow, Russia}
47:
48:
49: \begin{abstract}
50: In general or normal random matrix ensembles,
51: the support of eigenvalues of large size matrices is a planar domain
52: (or several domains) with a sharp boundary.
53: This domain evolves under a change
54: of parameters of the potential and of the size of matrices.
55: The boundary of the support of eigenvalues is a real section
56: of a complex curve. Algebro-geometrical properties
57: of this curve encode physical properties of random matrix ensembles.
58: This curve can be treated as a limit of a spectral curve
59: which is canonically defined for models of finite matrices.
60: We interpret the evolution of the eigenvalue distribution
61: as a growth problem, and describe the growth in terms of evolution
62: of the spectral curve. We discuss algebro-geometrical
63: properties of the spectral curve and describe the wave functions
64: (normalized characteristic polynomials) in terms
65: of differentials on the curve. General formulae and emergence
66: of the spectral curve are illustrated by three meaningful examples.
67: \end{abstract}
68:
69: \begin{keyword}
70: Integrable Systems \sep Random Matrix Theory \sep Stochastic Growth
71:
72: \PACS 02.50.Ey \sep 02.30.-f
73: \end{keyword}
74:
75: \end{frontmatter}
76:
77:
78: \section{Introduction and Preliminaries}
79: \label{sec:intro}
80: Recently, random matrix theory has found
81: new applications in growth problems, where the evolution
82: of the interface separating domains of different nature is a subject
83: of interest. In some realizations, the growing domain is an aggregate of randomly
84: deposited subunits.
85:
86: In general random matrix ensembles, complex eigenvalues
87: usually occupy planar
88: domains in the complex plane. Their boundaries
89: become sharp as the size of matrices, $N$, goes to infinity.
90: It appears that, for an important class of growth models, the aggregates evolve
91: similarly to the support of eigenvalues of general matrix ensembles.
92: In Refrs. \cite{mwz,ABWZ02,mkwz}, one of the most interesting classes
93: of growth problems -- Laplacian growth -- has been linked to the
94: evolution of normal random matrices.
95:
96: The interpretation of random matrix
97: theory in terms of an aggregation process seems to be a
98: productive approach
99: in a number of different applications. Among them are the growth problems
100: mentioned above, the
101: semiclassical behavior of electronic droplets in the Quantum Hall
102: regime \cite{ABWZ02}, and ${\mathcal N}=1$
103: supersymmetric Yang-Mills theory \cite{DV03}.
104: They reveal a relatively new, geometrical aspect of
105: random matrix theory. From the mathematical point of view,
106: the connection of random matrix ensembles
107: with isomonodromic deformations of differential
108: equations \cite{BEH,Kapaev,gravity,Its} seems to be especially
109: interesting for these applications.
110:
111: It has been emphasized in
112: recent papers
113: \cite{DV03} and \cite{KM03} that the boundary of the domains of
114: eigenvalues is to be thought of as a real section
115: of a complex algebraic curve (a Riemann surface). This curve
116: appears to be a fundamental object
117: of random matrix theory.
118: It encodes the most interesting physical
119: properties of matrix ensembles (see, e.g., \cite{DV03,KM03,Chekhov,David}
120: for the complex curves in one- and two-matrix models).
121:
122: In the Hermitian or unitary ensembles, the support
123: of eigenvalues is a set of line intervals.
124: Their boundaries are just points.
125: In general matrix ensembles, eigenvalues are complex numbers, and their
126: support
127: is a planar domain. The boundaries are planar curves, evolving
128: in a complicated and unstable manner.
129: The notion of complex curve
130: naturally links the planar geometry of the boundaries to
131: the algebro-geometric properties of the matrix ensemble.
132:
133:
134: In this paper we present
135: general aspects of the normal matrix ensemble from this standpoint.
136: One may canonically associate a complex curve
137: to the matrix ensemble, not only in the large
138: $N$ limit, but for any finite $N$ as well.
139: It is the spectral curve of
140: the operator which generates recursion relations
141: (the Lax operator),
142: projected to a certain finite-dimensional subspace.
143: In the large $N$ limit (a semiclassical limit),
144: this curve becomes the complex curve which describes
145: the support of eigenvalues.
146:
147: We introduce the spectral curve for normal matrix ensembles, and
148: describe the evolution of the curve
149: with respect to parameters of the statistical weight of the
150: ensemble {\it (deformation parameters)}
151: and the size of matrices (a parameter of
152: {\it growth}).
153: In a forthcoming paper, we will prove that the semiclassical limit of the evolution
154: (or deformation) equations
155: is the universal Whitham hierarchy associated with the complex curve.
156: The semiclassical wave function can be expressed
157: through differentials on the curve. We also notice that the
158: Whitham hierarchy is identical to the set of equations
159: which describe Laplacian growth processes --
160: unstable dynamics of an interface
161: between two immiscible phases.
162:
163: Three meaningful examples illustrate the general formulae.
164:
165:
166: \subsection{Normal matrix ensemble}
167:
168:
169: A matrix $M$ is called normal if it commutes with its
170: Hermitian conjugate: $[M, M^{\dag}]=0$, so that both $M$ and $M^{\dag}$
171: can be diagonalized simultaneously. Eigenvalues of normal matrices
172: are complex.
173: The statistical weight of the normal matrix ensemble is given through a
174: general potential $W(M,M^{\dag})$ \cite{Zaboronsky}:
175: \begin{equation}
176: \label{ZN}
177: e^{\frac{1}{\hbar}{\rm tr} \, W(M, M^{\dag})} d\mu (M).
178: \end{equation}
179: Here $\hbar$ is a parameter, and the measure of integration
180: over normal matrices
181: is induced by the flat metric
182: on the space
183: of all complex matrices $d_C M$, where
184: $d_C M = \prod_{ij}d\, {\mathcal R}e \, M_{ij} d\, {\mathcal I}m \, M_{ij}$.
185: Using the standard procedure,
186: (see, e.g., \cite{Mehta91}) one passes to the joint
187: probability distribution
188: of eigenvalues of normal matrices $z_1,\dots,z_N$, where $N$ is
189: size of the matrix:
190: \begin{equation}
191: \label{mean}
192: \frac{1}{N! \tau_N}|\Delta_N (z)|^2 \,\prod_{j=1}^N
193: e^{\frac{1}{\hbar}W(z_j,\bar z_j)} d^2 z_j
194: \end{equation}
195: Here
196: $d^2 z_j \equiv dx_j \, dy_j$ for $z_j =x_j +iy_j$,
197: $\Delta_N(z)=\det (z_{j}^{i-1})_{1\leq i,j\leq N}=
198: \prod_{i>j}^{N}(z_i -z_j)$
199: is the Vandermonde determinant, and
200: \begin{equation} \label{tau}
201: \tau_N = \frac{1}{N!}\int
202: |\Delta_N (z)|^2 \,\prod_{j=1}^{N}e^{\frac{1}{\hbar}
203: W(z_j,\bar z_j)} d^2z_j
204: \end{equation}
205: is the normalization factor. This is the partition function
206: of the matrix model (a $\tau$-function).
207:
208:
209: A particularly important special case arises
210: if the potential $W$ has the form
211: \begin{equation}
212: \label{potential}
213: W=-|z|^2+V(z)+\overline{V(z)},
214: \end{equation}
215: where $V(z)$ is a holomorphic function in a domain which
216: includes the support of eigenvalues (see also a comment in the end of
217: Sec.~\ref{F1}
218: about a proper definition of the ensemble with this potential).
219: In this case, a normal matrix ensemble gives the same distribution as
220: a general complex matrix ensemble.
221: A general complex matrix can be decomposed as $M=U(Z+R)U^\dagger$,
222: where $U$ and $Z$
223: are unitary and diagonal matrices, respectively, and $R$
224: is an upper triangular matrix. The distribution (\ref{mean})
225: holds for the elements of the diagonal matrix $Z$
226: which are eigenvalues of $M$ (Ref. \cite{Mehta91}).
227: Here we mostly focus on
228: the special potential (\ref{potential}), and also assume
229: that the field
230: \be\la{a}
231: A(z)=\p_z V(z)
232: \ee
233: (a ``vector potential'', see Sec. \ref{QH})
234: is a globally defined meromorphic function.
235:
236:
237:
238:
239:
240:
241: \subsection{Droplets of eigenvalues}
242: In a proper large $N$ limit ($\hbar\to 0$,
243: $N\hbar$ fixed), the eigenvalues of matrices from the
244: ensemble densely occupy a connected domain $D$
245: in the complex plane, or, in general, several
246: disconnected domains.
247: This set (called the support of eigenvalues)
248: has sharp edges
249: (Fig.~\ref{droplets}). We refer to the connected components
250: $D_\alpha$ of the domain $D$ as {\it droplets}.
251:
252: \begin{figure} \begin{center}
253: \includegraphics*[width=5cm]{dropletsa.eps}
254: \caption{A support of eigenvalues consisting of four disconnected components.}
255: \label{droplets}
256: \includegraphics*[width=5cm]{ellipse.eps}
257: \caption{The distribution of eigenvalues for the Gaussian potential. The droplet is an ellipse with quadrupole moment $2|t_2|$ and area $\pi\hbar N$.}
258: \label{ellipse}
259: \end{center} \end{figure}
260:
261: In the case of algebraic domains (the definition follows)
262: the eigenvalues are distributed with the density
263: $\rho=-\frac{1}{4 \pi}\Delta W$,
264: where $\Delta =4\p_z \p_{\bar z}$ is the 2-D Laplace operator
265: \cite{WZ03}.
266: For the potential (\ref{potential}) the density is uniform.
267: The shape of the support
268: of eigenvalues is a more involved subject. We discuss it below.
269: For example, if the
270: potential is Gaussian \cite{Ginibre65},
271: \be\la{e}
272: A(z)=2t_2 z,
273: \ee
274: the domain is an ellipse (see Fig.~\ref{ellipse}).
275: If $A$ has one simple pole,
276: \be\la{J1}
277: A(z)=-\frac{\alpha}{z-\beta}-\gamma
278: \ee
279: the droplet (under certain conditions discussed below) has the profile
280: of an aircraft wing given by the Joukowsky map
281: (Fig.~\ref{Joukowsky}).
282: \begin{figure} \begin{center}
283: \includegraphics*[width=5cm]{splitjuk.eps}
284: \caption{The distribution of eigenvalues for the potential $V(z) = - \alpha \log (1- z/\beta) - \gamma z $.
285: If the area of the droplet is $\pi\hbar N$, the complement of the droplet is an algebraic domain which can be conformally
286: mapped to the exterior of the unit disk by the Joukowsky map (\ref{222}). In this case, the cut of the Schwarz function
287: shown in the figure (corresponding to a virtual droplet located on the unphysical sheet) shrinks to a double point. }
288: \label{Joukowsky}
289: \end{center} \end{figure}
290: If $A$ has one double pole
291: (say, at infinity),
292: \be\la{hyp}
293: A(z)=3t_3 z^3,
294: \ee
295: the droplet is a hypotrochoid
296: (Fig.~\ref{hypotrochoid}).
297: These are the three examples that we will discuss below in detail.
298: \begin{figure} \begin{center}
299: \includegraphics*[width=5cm]{splithyp.eps}
300: \caption{The distribution of eigenvalues for the potential
301: $V(z) = t_3z^3$. Cuts of the Schwarz function are shown. The cuts
302: correspond to virtual droplets located on unphysical sheets.
303: If the area of the droplet is $\pi\hbar N$, the cuts shrink to double points.
304: The boundary contour is a hypotrochoid.}
305: \label{hypotrochoid}
306: \end{center} \end{figure}
307: If $A$ has two or more simple poles, there may be
308: more than one droplet.
309:
310:
311: If $A(z)$ is a more complicated function,
312: the domain of eigenvalues develops
313: an unstable fingering pattern,
314: similar to the one in Fig.~\ref{sw}.
315: \begin{figure} \begin{center}
316: \includegraphics*[width=5cm]{sw.eps}
317: \caption{\label{sw}
318: A grown fingering pattern observed
319: in a radial Hele-Shaw cell.
320: Air is inserted under pressure to the cell filled by
321: silicon oil \cite{Sw}.}
322: \end{center} \end{figure}
323:
324: Boundary components of the droplets form a real section of
325: a complex curve.
326:
327: \subsection{Semiclassical complex curve}
328: Jumping ahead, we describe the construction of the semiclassical
329: complex curve which emerges in the
330: normal matrix ensemble. Let us
331: represent the boundary of the domain
332: as a real curve
333: $F(x,\,y)=0$. If the vector potential $A(z)$ is a meromorphic
334: function (we always assume that this is the case), the function
335: $F$ can be chosen to be an irreducible polynomial.
336: Then we rewrite it
337: in holomorphic coordinates as
338: \be\label{F}
339: F\left (\frac{z+\bar z}{2},\frac{z-\bar z}{2i}\right )=
340: f(z,\bar z)
341: \ee
342: and treat $z$ and $\bar z$ as independent complex coordinates
343: $z$, $\tilde z$.
344: The equation
345: $f(z,\tilde z)=0$ defines a complex curve.
346: This curve is a finite-sheet covering of the $z$-plane.
347: The single-valued function $\tilde z(z)$
348: on the curve is a multivalued function on
349: the $z$-plane. Making cuts, one can fix single-valued
350: branches of this function.
351:
352: The boundary of the domain is a section of the curve by the plane where
353: $\tilde z$ is complex conjugate of $z$.
354: It belongs to a particular
355: sheet (we call it the {\it physical sheet}) of the covering.
356: It appears that physical properties of the ensemble are
357: determined not only by the physical sheet,
358: but rather by the entire algebraic
359: covering, including all the sheets other than physical. The Riemann
360: surface for
361: Example (\ref{J1}) is presented in Fig.~\ref{torus}.
362:
363:
364:
365:
366: \subsection{Growth and deformation parameters. Evolution of the curve}
367: Similarly to the Hermitian matrix ensemble
368: \cite{DV03,KM03,Chekhov,David},
369: the complex curve of the normal matrix ensemble
370: is characterized by the potential $W$, or by
371: the ``vector potential"
372: $A(z)$, and by a set of $g+1$
373: integers $\nu_\alpha$ (not necessarily positive), where $g$ is
374: genus of the curve.
375: The integers are subject to the constraint
376: $\sum_{\alpha=0}^g\nu_\alpha=N$.
377: We will discuss the meaning of these numbers later on.
378: For now, we only mention that if they are all positive, then they
379: are proportional to the areas of the droplets
380: of uniformly distributed eigenvalues. Every droplet contains
381: $\nu_\alpha$ eigenvalues, so a quantum of area is
382: $\pi\hbar$ per particle.
383:
384: As one varies the potential
385: and the filling factors $\nu_\alpha$,
386: the curve and the interface bounding the droplets evolve.
387: Parameters of the potential
388: (for example, poles and residues of the meromorphic function (\ref{a}))
389: and filling numbers are
390: {\it deformation parameters} and {\it parameters of growth}.
391: They are coordinates
392: in the moduli space of the complex curves.
393:
394:
395: An infinitesimal variation of the potential
396: generates correlation functions of the ensemble
397: (Ref. \cite{WZ03}). For example,
398: irreducible correlation functions
399: of resolvents (holomorphic currents)
400: $J (z)=\mbox{tr}\, \frac{\hbar}{z-M}=\sum_i\frac{\hbar}{z-z_i}$
401: are generated by variations of $A(z)$.
402: In Ref. \cite{WZ03} we showed that the correlation functions
403: are expressed through algebro-geometrical properties of the boundary of
404: the domain.
405:
406: During the evolution, the genus of the
407: complex curve may change. This
408: corresponds to a coalescence of droplets or a droplet breakup, or
409: even more complicated
410: degenerations, when a droplet shrinks to a point
411: and simultaneously merges with another droplet.
412: Close to degeneracy points (also called critical points), matrix
413: ensembles have universal scaling behavior.
414: We will present a study of this problem in a future paper.
415:
416: A particularly interesting process is {\it growth}. In this case, one
417: changes the
418: total number of eigenvalues $N$
419: and keeps the potential fixed. Below we refer to the normalized
420: total number of particles
421: $$t=\hbar N$$
422: as {\it time of growth}. In the algebraic case, the time $t$ is
423: the normalized (modulo $\pi$) area of the droplets. While $N$
424: increases, the new eigenvalues aggregate at the boundary of the
425: existing droplets,
426: so that the domain of eigenvalues (an aggregate) grows
427: (Fig.~\ref{semi}).
428: \begin{figure} \begin{center}
429: \includegraphics*[width=5cm]{semicl.eps}
430: \caption{\label{semi}The area of a droplet grows with the rate $\pi \hbar$
431: per eigenvalue. A new eigenvalue aggregates at the boundary of the
432: droplet. The shaded area
433: depicts the support of the semiclassical wave function.}
434: \end{center} \end{figure}
435: Growth of a hypotrochoid is depicted in Fig.~\ref{evolutionhyp}.
436:
437:
438: \begin{figure} \begin{center}
439: \includegraphics*[width=5cm]{evolutionhyp.eps}
440: \caption{\label{evolutionhyp}Hypotrochoid grows
441: until it reaches a critical point.}
442: \end{center} \end{figure}
443:
444: There are several important physical problems where this kind of evolution
445: occurs. We list them in historical order.
446:
447: The first one is the celebrated Hele-Shaw problem -- the most studied
448: example of Laplacian growth.
449: The Hele-Shaw problem describes the non-equilibrium
450: dynamics of a planar interface between two immiscible fluids
451: confined to a thin 2-D cell.
452: The second one is the matrix
453: model description of 2-D quantum gravity.
454: The third one is the evolution of energy levels in mesoscopic systems.
455: A similar problem also emerges in isomonodromic deformation of
456: differential equations.
457:
458: The last three problems and their connection with the matrix ensemble
459: are well reflected
460: in the literature (see e.g., \cite{gravity,Its,mesoscopic} for reviews).
461: The connection of matrix ensembles to the Hele-Shaw
462: problem is relatively new.
463: It nicely
464: illustrates the growth and gives
465: a transparent hydrodynamic interpretation to abstract
466: objects of algebraic geometry of curves and complex analysis on
467: Riemann surfaces.
468: We find that a brief description of the Hele-Shaw problem in this
469: paper is in order (see Sec. \ref{Appendix_A}).
470:
471:
472:
473: \subsection{Normal matrix ensemble
474: and quantum Hall effect}\la{QH}
475:
476: A useful interpretation of the Coulomb gas distribution (\ref{mean}) is
477: a coherent state of relativistic electrons in the Quantum Hall regime
478: \cite{ABWZ02}. In this case, the electrons are
479: situated in the plane in a strong, not necessarily
480: uniform, magnetic field
481: $B(z)=-\frac{1}{2} \Delta W$, and fully occupy the lowest
482: energy level. The exact $N$-particle wave function, defined up to a phase, is
483: \begin{equation}\label{psi}
484: \Psi_N(z_1,\dots,z_N)=
485: \frac{1}{\sqrt{N! \tau_N}}\Delta_N (z) \,e^{\sum_{j=1}^{N}
486: \frac{1}{2\hbar}W(z_j, \bar z_j)}.
487: \end{equation}
488: The joint probability distribution (\ref{mean})
489: is then equal to $|\Psi(z_1,\dots,z_N)|^2$.
490: It is the probability to find electrons
491: at the points $z_i$. In the semiclassical
492: regime, the wave function (\ref{psi})
493: describes an incompressible electronic droplet with sharp edges.
494: The function $A(z)$ (\ref{a}) can be thought of as
495: a holomorphic component of the vector potential
496: generated by magnetic impurities located far away from the
497: electronic droplet.
498:
499:
500: In this language, the growth problem translates into the evolution of
501: a semiclassical
502: electronic droplet under a change of magnetic
503: field or chemical potential.
504:
505:
506: \subsection{Orthogonal polynomials as a measure of growth}\la{F1}
507: Let the number of eigenvalues (particles)
508: increase while the potential stays fixed. If the
509: support of eigenvalues is
510: simply-connected, its area
511: grows as $\hbar N$. One can describe the
512: evolution of the domain through the density of particles
513: \begin{equation}\la{51}
514: \rho_N(z)=N \int|\Psi_{N}(z,z_1,z_2,\dots,z_{N-1})|^2
515: d^2z_1\dots d^2 z_{N-1},
516: \end{equation}
517: where $\Psi_N$ is given by (\ref{psi}).
518:
519: We introduce a set of orthonormal one-particle functions
520: on the complex plane as matrix elements of transitions
521: between $N$ and $(N+1)$-particle states:
522: \begin{equation}\la{5}
523: \frac{\psi_N(z)}{\sqrt{N+1}}= \int\Psi_{N+1} (z,z_1,z_2,\dots,z_N)
524: \overline{\Psi_{N}(z_1,z_2,\dots,z_N)} d^2z_1\dots d^2 z_N
525: \end{equation}
526: Then the rate of the density change is
527: \begin{equation}\la{1}
528: \rho_{N+1}(z)-\rho_{N}(z)=|\psi_N(z)|^2.
529: \end{equation}
530: The proof of this formula
531: is based on the representation of the $\psi_n$ through
532: holomorphic biorthogonal polynomials $P_n(z)$. Up to a phase
533: \begin{equation}\label{O3}
534: \psi_n (z)=
535: e^{\frac{1}{2\hbar}W(z, \bar z )}P_n (z),\quad
536: \quad P_n (z)=\sqrt{\frac{\tau_n}{\tau_{n+1}}} z^n +\ldots
537: \end{equation}
538: The polynomials $P_n(z)$ are
539: biorthogonal on the complex plane
540: with the weight $e^{W/\hbar}$:
541: \begin{equation}\la{29}
542: \int e^{W/\hbar}P_n(z)\overline{P_m(z)}d^2 z=
543: \delta_{mn}
544: .
545: \end{equation}
546: The proof of these formulae
547: is standard in the theory of orthogonal
548: polynomials. Extension to the biorthogonal case adds no
549: difficulties.
550:
551:
552:
553: A physical interpretation of the
554: wave function $\psi_n$ is clear in the QHE
555: setup. The eigenvalues $z_i$
556: are the positions of electrons in a strong magnetic field.
557: Then $|\psi_N(z)|^2$ is
558: a probability of adding an additional electron to
559: an aggregate of $N$ electrons at the point $z$. Since the droplet is
560: incompressible, it
561: is only possible to add the particle to the boundary of the droplet.
562: We will see that the wave function
563: in a semiclassical limit is indeed localized on the boundary.
564:
565: We note that, with
566: the choice of potential (\ref{potential}), the integral
567: representation (\ref{29}) has only a formal meaning,
568: since the integral diverges
569: unless the potential is Gaussian.
570: A proper definition of the wave functions goes through
571: recursive relations (\ref{L1}, \ref{M}) which follow from the integral
572: representation.
573: The same comment applies to the $\tau$-function (\ref{tau}).
574: The wave function is not normalized everywhere
575: in the complex plane. It may diverge at the poles of the
576: vector potential field.
577:
578: Below we describe the evolution and
579: semiclassical behavior
580: of the wave function in the region of the complex plane close to the
581: boundary of physical droplets.
582:
583:
584:
585:
586: \section{Equations for the wave functions and the
587: spectral curve}\la{1A}
588: In this section we specify the potential to be
589: of the form (\ref{potential}).
590: It is convenient to modify the
591: exponential factor of the wave function.
592: Namely, we define
593: \be \la{chin}
594: %\nonumber
595: \psi_n (z)=
596: e^{-\frac{|z|^2}{2\hbar}+\frac{1}{\hbar}V(z)}P_n (z),\quad\mbox{and}\quad\chi_n (z)=
597: e^{\frac{1}{\hbar}V(z)}P_n (z),
598: \ee
599: where the holomorphic
600: functions $\chi_n(z)$
601: are orthonormal in the complex plane
602: with the weight
603: $e^{-|z|^2/\hbar}$.
604: Like traditional orthogonal polynomials, the biorthogonal polynomials
605: $P_n$ (and the corresponding wave functions)
606: obey a set of differential equations with
607: respect to the argument $z$, and recurrence relations with respect to
608: the degree $n$. Similar equations
609: for two-matrix models are discussed in numerous papers
610: (see, e.g., \cite{Aratyn}).
611:
612: We introduce the $L$-operator (the Lax operator)
613: as multiplication by $z$ in the
614: basis $\chi_n$:
615: \begin{equation}\la{L1}
616: L_{nm}\chi_m(z)=z\chi_n(z)
617: \end{equation}
618: (summation over repeated indices is implied).
619: Obviously, $L$ is a lower triangular matrix with one
620: adjacent upper diagonal, $L_{nm}=0$ as $m>n+1$.
621: Similarly, the differentiation
622: $\p_z$ is represented by an
623: upper triangular matrix with one adjacent lower diagonal.
624: Integrating by parts the matrix elements of the $\p_z$, one finds:
625: \begin{equation}\la{M}
626: (L^{\dag})_{nm}\chi_m =
627: \hbar\p_z\chi_n,
628: \end{equation}
629: where $L^{\dag}$
630: is the Hermitian conjugate operator.
631:
632: The matrix elements of $L^{\dag}$ are
633: $(L^{\dag})_{nm}=\bar L_{mn}=A(L_{nm})+
634: \int e^{\frac{1}{\hbar}W}\bar P_m(\bar z)\p_z P_n(z) d^2z$, where
635: the last term is a lower triangular matrix. The latter can be written
636: through negative powers of the Lax operator.
637: Writing $\p_z\log P_n(z)=\frac{n}{z}+\sum_{k>1}v_k(n)z^{-k}$,
638: one represents $L^{\dag}$ in the form
639: \begin{equation}\la{M1}
640: L^{\dag}=A(L)+(\hbar n) L^{-1}+\sum_{k>1}v^{(k)}L^{-k},
641: \end{equation}
642: where $v^{(k)}$ and $(\hbar n)$ are
643: diagonal matrices with elements $v_n^{(k)}$
644: and $(\hbar n)$.
645: The coefficients $v_{n}^{(k)}$ are determined by the condition that
646: lower triangular matrix elements of $A(L_{nm})$ are cancelled.
647:
648: In order to emphasize the structure of the operator $L$, we
649: write it in the basis of the
650: shift operator \footnote{The shift operator
651: $\hat w$ has no
652: inverse. Below $\hat w^{-1}$ is understood as a shift to the left
653: defined as $\hat w^{-1}\hat w=1$.
654: Same is applied to the operator $L^{-1}$. To avoid a possible
655: confusion, we emphasize that although $\chi_n$
656: is a right-hand eigenvector of $L$, it is not a right-hand
657: eigenvector of $L^{-1}$.}
658: $\hat w$ such that $\hat w f_n =f_{n+1}\hat w$ for any
659: sequence $f_n$. Acting on the wave function, we have:
660: $$\hat w
661: \chi_n=\chi_{n+1}.$$
662: In the $n$-representation, the operators $L$, $L^{\dag}$
663: acquire the form
664: \begin{equation}\la{M11}
665: L=r_n \hat w+\sum_{k\geq 0} u_{n}^{(k)} \hat w^{-k},\quad
666: L^{\dag} = \hat w^{-1} r_n+
667: \sum_{k\geq 0} \hat w^{k} \bar u_{n}^{(k)}.
668: \end{equation}
669: Clearly, acting on $\chi_n$, we have the commutation
670: relation (``the string equation")
671: \begin{equation}\la{string}
672: [L, \, L^{\dag}]=\hbar.
673: \end{equation}
674: This is the compatibility condition of Eqs. (\ref{L1}) and (\ref{M}).
675:
676: Equations (\ref{M11}) and (\ref{string})
677: completely determine the coefficients
678: $v_n^{(k)}$, $r_n$ and $u_n^{(k)}$. The first one connects the coefficients
679: to the parameters of the potential.
680: The second equation is used to determine how the coefficients
681: $v_n^{(k)}$, $r_n$ and $u_n^{(k)}$ evolve with $n$. In particular,
682: the diagonal part of it reads
683: \be\la{Area}
684: n\hbar=r_n^2-\sum_{k\geq 1}\sum_{p=1}^k|u_{n+p}^{(k)}|^2.
685: \ee
686: Moreover, we note that all the coefficients can be expressed
687: through the $\tau$-function (\ref{tau}) and its derivatives
688: with respect to parameters of the potential.
689: This representation is particularly simple for $r_n$:
690: $
691: r_n^2 =\tau_n \tau_{n+1}^{-2}\tau_{n+2}
692: $.
693:
694:
695: \subsection{Finite dimensional reductions}
696: If the vector potential $A(z)$ is a rational function,
697: the coefficients $u_{n}^{(k)}$ are not all independent.
698: The number of independent coefficients
699: equals the number of independent parameters of the potential.
700: For example, if the holomorphic part
701: of the potential, $V(z)$, is a polynomial of degree $d$,
702: the series
703: (\ref{M11}) are truncated at $k= d-1$.
704:
705:
706: In this case the semi-infinite system of linear equations (\ref{M})
707: and the recurrence relations (\ref{L1})
708: can be cast in the form of a set of finite dimensional equations whose
709: coefficients are rational functions of $z$, one
710: system for every $n>0$.
711: The system of differential equations generalizes the
712: Cristoffel-Daurboux second order
713: differential equation
714: valid for orthogonal polynomials. This fact has been observed in
715: recent papers \cite{BEHdual,Eynard03}
716: for biorthogonal polynomials emerging in the
717: Hermitian two-matrix model
718: with a polynomial potential. It is
719: applicable to our case (holomorphic biorthogonal
720: polynomials) as well.
721:
722: In a more general case,
723: when $A(z)$ is a general rational function with $d-1$ poles
724: (counting multiplicities), the series (\ref{M11})
725: is not truncated. However, $L$ can be represented
726: as a ``ratio",
727: \beq\label{LKK}
728: L=K_{1}^{-1}K_{2}=M_2 M_{1}^{-1},
729: \eeq
730: where the operators $K_{1,2}$, $M_{1,2}$ are
731: polynomials in $\hat w$:
732: \beq\label{KK}
733: K_1 =\hat w^{d-1}+\sum_{j=0}^{d-2}A_{n}^{(j)} \hat w^j\,,
734: \quad
735: K_2 =r_{n\! +\! d\! -\! 1}\hat w^{d}+\sum_{j=0}^{d-1}B_{n}^{(j)} \hat w^j
736: \eeq
737: \beq\label{MM}
738: M_1 =\hat w^{d-1}+\sum_{j=0}^{d-2}C_{n}^{(j)} \hat w^j\,,
739: \quad
740: M_2 =r_{n}\hat w^{d}+\sum_{j=0}^{d-1}D_{n}^{(j)} \hat w^j
741: \eeq
742: These operators obey the
743: relation
744: \beq\label{KMKM}
745: K_1 M_2 =K_2 M_1.
746: \eeq
747: It can be proven that
748: the pair of operators $M_{1,2}$ is uniquely determined
749: by $K_{1,2}$ and vice versa. We note that the
750: reduction (\ref{LKK}) is a difference analog of the
751: ``rational" reductions of the Kadomtsev-Petviashvili
752: integrable hierarchy considered in \cite{Krichev-red}.
753:
754: The linear problems (\ref{L1}), (\ref{M}) acquire the form
755: \beq\label{L1a}
756: (K_2\chi )_n =z \, (K_1 \chi )_n\,, \quad
757: (M_{2}^{\dag} \chi )_n =\hbar \p_z (M_{1}^{\dag}\chi ) _n.
758: \eeq
759: These equations are of {\it finite order}
760: (namely, of order $d$), i.e., they connect values of $\chi_n$
761: on $d+1$ subsequent sites of the lattice.
762:
763:
764: The semi-infinite set $\{\chi_0,\chi_1,\dots\}$ is
765: then a ``bundle" of
766: $d$-dimensional vectors
767: $${\underline\chi}(n)= (\chi_n,\chi_{n+1},\dots,
768: \chi_{n+d-1})^{{\rm t}}$$
769: (the index ${\rm t}$
770: means transposition, so
771: ${\underline\chi}$ is a column vector).
772: The dimension of the vector is the number of poles
773: of $A(z)$ plus one.
774: Each vector obeys a closed $d$-dimensional linear
775: differential equation
776: \begin{equation}
777: \la{M2}
778: \hbar\p_z{\underline\chi}(n)={\mathcal L}_n (z){\underline\chi}(n),
779: \end{equation}
780: where the $d\times d$ matrix
781: ${\mathcal L}_n$ is a ``projection" of the operator $L^{\dag}$
782: onto the $n$-th $d$-dimensional space. Matrix elements of the
783: ${\mathcal L}_n$ are rational functions of
784: $z$ having the same poles as $A(z)$ and also a pole at
785: the point $\overline{A(\infty )}$. (If $A(z)$ is a polynomial,
786: all these poles accumulate to a multiple pole at infinity).
787:
788: We briefly describe the procedure of
789: constructing the finite dimensional matrix
790: differential equation. We use the first
791: linear problem in (\ref{L1a})
792: to represent the shift operator
793: as a $d\times d$ matrix ${\mathcal W}_n (z)$ with
794: $z$-dependent coefficients:
795: \be\la{W}
796: {\mathcal W}_n(z){\underline\chi}(n)={\underline\chi}(n \! +\! 1).
797: \ee
798: This is nothing else than rewriting the scalar linear problem
799: in the matrix form.
800: Then the matrix ${\mathcal W}_n (z)$ is to be substituted into the
801: second equation of (\ref{L1a}) to
802: determine ${\mathcal L}_n(z)$ (examples follow).
803: The entries of ${\mathcal W}_n(z)$ and ${\mathcal L}_n (z)$
804: obey the Schlesinger
805: equation, which follows from compatibility of
806: (\ref{M2}) and (\ref{W}):
807: \be\la{WW}
808: \hbar\p_z {\mathcal W}_n=
809: {\mathcal L}_{n+1} {\mathcal W}_n
810: -{\mathcal W}_n {\mathcal L}_n.
811: \ee
812:
813: This procedure has been realized explicitly for
814: polynomial potentials in
815: recent papers \cite{BEHdual,Eynard03}.
816: We will work it out in detail for our three examples:
817: ${\underline\chi}(n)=(\chi_n,\,\chi_{n+1})^{{\rm t}}$ for the ellipse
818: (\ref{e}) and
819: the aircraft wing (\ref{J1}) and
820: ${\underline\chi}(n)=(\chi_n,\,\chi_{n+1},\chi_{n+2})^{{\rm t}}$
821: for the hypotrochoid (\ref{hyp}).
822:
823:
824:
825: \subsection{Spectral curve}\la{Curve1}
826:
827:
828: According to the general theory of
829: linear differential equations, the
830: semiclassical (WKB) asymptotics of solutions to
831: Eq. (\ref{M2}), as $\hbar \to 0$,
832: is found by solving the eigenvalue problem for
833: the matrix ${\mathcal L}_n (z)$ (see, e.g., \cite{Wasow}) .
834: More precisely, the basic object of the WKB approach is the
835: spectral curve \cite{curve} of the matrix ${\mathcal L}_n$, which is
836: defined, for every integer $n>0$, by the secular equation
837: $\det ({\mathcal L}_n (z)-\tilde z) = 0$
838: (here $\tilde z$ means $\tilde z \cdot {\bf 1}$, where
839: ${\bf 1}$ is the unit $d\times d$ matrix).
840: It is clear that the left hand side of the secular
841: equation is a polynomial in $\tilde z$ of degree $d$.
842: We define the spectral curve by an equivalent equation
843: \begin{equation}\la{qc}
844: f_n (z,\tilde z)=a(z)\det ({\mathcal L}_n (z)-\tilde z) = 0,
845: \end{equation}
846: where the factor $a(z)$ is added
847: to make $f_n(z,\tilde z)$ a polynomial in $z$ as well.
848: The factor $a(z)$ then has zeros at the points where
849: poles of the matrix function ${\mathcal L}(z)$ are located.
850: It does not depend on $n$.
851: We will soon see that the degree of
852: the polynomial $a(z)$ is equal to $d$.
853: Assume that all poles of $A(z)$ are simple, then zeros of
854: the $a(z)$ are
855: just the $d-1$ poles of $A(z)$ and another simple zero
856: at the point $\overline{A(\infty)}$. Therefore, we conclude that
857: the matrix ${\mathcal L}_n (z)$ is rather special.
858: For a general $d\times d$ matrix function with the same
859: $d$ poles, the factor $a(z)$ would be of degree $d^2$.
860:
861: Note that the matrix ${\mathcal L}_n (z) -\bar z$ enters
862: the differential equation
863: \begin{equation}\la{M2'}
864: \hbar\p_{z}|{\underline\psi}(n)|^2=\bar {\underline\psi}(n)
865: ({\mathcal L}_n (z)- \bar z)
866: {\underline\psi}(n)
867: \end{equation}
868: for the squared amplitude
869: $|{\underline \psi}(n) |^2={\underline \psi}^{\dag}(n)
870: {\underline \psi}(n) =
871: e^{-\frac{|z|^2}{\hbar}} |{\underline \chi}(n)|^2$
872: of the vectors
873: ${\underline \psi}(n)$ built from the orthonormal
874: wave functions (\ref{O3}).
875:
876: The equation of the curve can be interpreted as a
877: ``resultant" of the non-commutative polynomials
878: $K_2 -zK_1$ and $M_{2}^{\dag}-\tilde z M_{1}^{\dag}$
879: (cf. \cite{BEHdual}). Indeed, the point $(z, \tilde z)$
880: belongs to the curve if and only if the linear system
881: \beq\label{linsys}
882: \left \{ \begin{array}{l}
883: (K_2 c)_k =z (K_1 c)_k \quad \quad n-d\leq k \leq n-1
884: \\ \\
885: (M_{2}^{\dag} c)_k =\tilde
886: z (M_{1}^{\dag} c)_k \quad \quad n\leq k \leq n+d-1
887: \end{array}\right.
888: \eeq
889: has non-trivial solutions. The system contains $2d$ equations
890: for $2d$ variables $c_{n-d}\, , \ldots , c_{n+d -1}$.
891: Vanishing of the
892: $2d \, \times \, 2d$ determinant yields the equation of
893: the spectral curve.
894: Below we use this method to find the equation of the curve
895: in the examples.
896: It appears to be much easier than the determination of the
897: matrix ${\mathcal L}_n(z)$.
898:
899:
900: The spectral curve (\ref{qc}) possesses an important property:
901: it admits an antiholomorphic involution.
902: In the coordinates $z, \tilde z$ the involution reads
903: $(z, \tilde z)\mapsto (\overline{\tilde z}, \bar z)$.
904: This simply means that the secular equation
905: $\det (\bar {\mathcal L}_n (\tilde z)-z) = 0$
906: for the matrix $\bar {\mathcal L}_n (\tilde z)\equiv
907: \overline{{\mathcal L}_n (\overline{\tilde z})}$
908: defines the same curve.
909: Therefore, the polynomial $f_n$
910: takes real values for $\tilde z =\bar z$:
911: \be\la{anti1}
912: f_n(z,\bar z)=\overline{f_n ( z, \bar z)}.
913: \ee
914: Points of the real section of the curve
915: ($\tilde z =\bar z$) are
916: fixed points of the involution.
917:
918: The curve (\ref{qc})
919: was discussed in recent papers \cite{BEHdual,Eynard03}
920: in the context of Hermitian two-matrix models with
921: polynomial potentials. The dual realizations of the curve
922: pointed out in \cite{BEHdual} correspond to the
923: antiholomorphic involution in our case.
924: The involution can be proven along the lines of these
925: works. The proof is rather technical and we omit
926: it, restricting ourselves to the
927: examples below. We simply note that the involution relies on the fact
928: that the squared modulus of the wave function is real.
929:
930: \subsection{Schwarz function}
931:
932: The polynomial $f_n(z, \bar z)$ can be
933: factorized in two ways:
934: \be\la{h1}f_n(z,\bar z)=a(z)(\bar z-S_n^{(1)}(z))
935: \dots (\bar z-S_n^{(d)}(z)),
936: \ee
937: where $S_n^{(i)}(z)$ are eigenvalues of the matrix
938: ${\mathcal L}_n (z)$, or
939: \be\la{ah1}
940: f_n(z,\bar z)=\overline{a(z)}( z-\bar S_n^{(1)}(\bar z))\dots ( z-\bar
941: S_n^{(d)}(\bar z)),
942: \ee
943: where $\bar S_n^{(i)}(\bar z)$ are eigenvalues of the matrix
944: $\bar {\mathcal L}_n (\bar z)$.
945: One may understand them as different branches of a
946: multivalued function $S(z)$ (respectively, $\bar S(z)$)
947: on the plane (here we do not indicate the dependence on $n$,
948: for simplicity of the notation).
949: It then follows that
950: $S(z)$ and $\bar S(z)$
951: are mutually inverse functions:
952: \begin{equation}\la{anti11}
953: \bar S(S(z))=z.
954: \end{equation}
955:
956:
957: An algebraic function with this property
958: is called {\it the Schwarz function}.
959: By the equation $f(z, S(z))=0$, it defines a complex curve
960: with an antiholomorphic involution. An upper bound
961: for genus of this curve is $g=(d-1)^2$, where $d$ is the
962: number of branches of the Schwarz function.
963: The real section of this curve is a set of all fixed
964: points of the involution. It consists of a number
965: of contours on the plane (and possibly a number of
966: isolated points, if the curve is not smooth).
967: The structure of this set is known to be
968: complicated. Depending on
969: coefficients of the polynomial, the number of disconnected
970: contours in the real section may vary from $0$ to $g+1$.
971: If the contours divide the complex curve into two
972: disconnected ``halves", or sides (related by the involution), then
973: the curve can be realized as
974: the {\it Schottky double} \cite{Alhfors50,C,SS} of one of
975: these sides. Each side is a Riemann surface with a boundary.
976:
977: We will discuss general properties of the Schwarz function and the
978: Schottky double in Sec. \ref{S}.
979:
980: Let us come back to
981: equation (\ref{M2}). It has $d$ independent solutions.
982: They are functions on the spectral curve.
983: One of them is a physical solution
984: corresponding to biorthogonal polynomials.
985: The physical solution defines the ``physical sheet"
986: of the curve.
987:
988:
989: The Schwarz function on the physical sheet
990: is a particular root, say $S^{(1)}_n(z)$, of the
991: polynomial $f_n(z, \tilde z)$ (see (\ref{h1})).
992: It follows from (\ref{M1}) that
993: this root is
994: selected by the requirement that
995: it has the same poles and residues as the potential $A$.
996:
997:
998: A formal \footnote{This formal expression ignores
999: the Stokes phenomenon. It
1000: is valid only around boundaries of physical droplets.}
1001: semiclassical asymptote of the
1002: solution, in the leading order in $\hbar$, is
1003: \be\la{semi1}
1004: \chi_n\sim e^{\frac{1}{\hbar}\int^z d\Omega^{(1)}_n}.
1005: \ee
1006: Here
1007: $$d\Omega_n^{(1)}=S_n^{(1)}dz.$$
1008: The differential $d\Omega^{(1)}$ is a physical branch of the generating
1009: differential on the curve (see below).
1010:
1011:
1012:
1013: The semiclassical asymptotics is
1014: discussed in more details in Sec. \ref{Semi2}.
1015: Here we notice that the amplitude of the wave function
1016: $\psi_n (z)$ ( not $\chi_n (z)$!) peaks at the solution
1017: of the equation
1018: \be\la{S1}
1019: S^{(1)}(z)=\bar z,
1020: \ee
1021: where now $z=x+iy$ and $\bar z=x-iy$ are complex conjugated
1022: coordinates in the plane.
1023: Solutions to this equation describe
1024: those contours of a real section of the complex curve
1025: which belong to the physical sheet.
1026: It is a set of closed
1027: planar curves, as shown in Fig.~\ref{droplets}.
1028: These curves
1029: are boundaries of semiclassical droplets. Evolution and growth of
1030: the droplets
1031: translate into evolution of the complex curve.
1032:
1033:
1034:
1035:
1036:
1037:
1038:
1039: \subsection{Deformation parameters}
1040: The essential information about the spectral curve
1041: (which, in what follows, we will call {\it quantum curve},
1042: unless the semiclassical limit is considered)
1043: is contained in two conditions:
1044: \begin{itemize}
1045: \item
1046: the antiholomorphic involution, and
1047: \item poles and residues
1048: of the Schwarz function on one selected sheet (the physical
1049: sheet) are given by the vector potential field $A(z)$.
1050: \end{itemize}
1051: These requirements determine all but $g$ coefficients
1052: of the polynomial $f_n(z,\bar z)$, where $g$ is genus of the curve.
1053: The remaining $g$ coefficients do not depend on the potential
1054: (deformation parameters).
1055:
1056:
1057:
1058: The Bohr-Sommerfield quantization condition for the semiclassical
1059: wave function
1060: (\ref{semi1}) requires that
1061: the integral over all the cuts (or boundaries of the droplets) are integer
1062: multiples of $2\pi i$. Also, the integral around the point
1063: at infinity, counting
1064: the degree of the polynomial, must be $2\pi iN$. The integrals over
1065: the cuts transform
1066: into the integrals over $\bf a$-cycles of the curve. Later we
1067: will interpret the $\bf a$-cycles
1068: as boundaries of the droplets. Therefore, we have the
1069: set of integers
1070: \begin{equation}\la{18}
1071: \nu_a=
1072: \frac{1}{2\pi i \hbar}\oint_{{\bf a}_\alpha} S(z) dz,
1073: \quad\sum_\alpha\nu_\alpha=N,\quad\alpha=0,\,1,\dots, g.
1074: \end{equation}
1075:
1076: The $d$-dimensional linear differential equations (\ref{M2})
1077: give an interpretation of the potential data as deformation parameters.
1078: Varying the poles and residues of
1079: $A(z)$, one does not change the monodromy of solutions, given by $N$
1080: and $\nu_\alpha$.
1081: They are similar to deformation parameters of the theory of
1082: isomonodromic deformations of the system of linear differential
1083: equations \cite{Miwa} (see also Ref. \cite{Its} for applications of the
1084: theory of isomonodromic deformations to Hermitian matrix models
1085: and orthogonal polynomials in one variable, and references therein).
1086: A connection between isomonodromic deformations and
1087: Whitham equations, which seems to be relevant
1088: to our discussion, was pointed out in Ref. \cite{Takasaki}.
1089:
1090: In the spirit of the theory of isomonodromic deformations of
1091: linear differential equations, we conjecture that a specific value of
1092: $N$, a potential $V(z)$, and a set of
1093: $g$ numbers (\ref{18}) form a complete data set which uniquely
1094: determines the coefficients
1095: $r_n,\,u^{(k)}_n$ of the recurrence equations, and the solution of the
1096: differential equation (\ref{M2}). We do not know, however,
1097: whether the requirement that the physical solutions are polynomials gives a
1098: restriction on
1099: possible values of the integers $\nu_\alpha$. It is likely that a general
1100: set of $\nu_\alpha$ (positive and negative)
1101: corresponds to more general solutions of (\ref{M2}) rather than polynomials.
1102:
1103: We also conjecture that, for physical values
1104: of the parameters of the potential in general position,
1105: the quantum curve is smooth (i.e., non-degenerate), and
1106: has the maximal possible number of the disconnected
1107: closed contours (``real ovals") in the real section.
1108: In the latter case, the involution is known to divide the
1109: complex curve into two ``halves", thus making the
1110: Schottky double construction applicable.
1111:
1112: Before discussing general properties of the curve, we illustrate
1113: its appearance in the semiclassical limit.
1114:
1115: \section{Semiclassical curve}\la{SCurve}
1116:
1117: By semiclassical case we mean the limit $N \to \infty$, $\hbar \to 0$, while
1118: the time $t=\hbar N$,
1119: and the potential remain fixed. In this case, all eigenvalues are distributed
1120: continuously within
1121: droplets with sharp boundaries.
1122: The boundary of the droplets is a real section of the {\it
1123: classical curve}.
1124:
1125: \subsection{Saddle point equation and Riemann-Hilbert
1126: problem}\la{Saddle}In the semiclassical limit,
1127: the distribution
1128: (\ref{mean}) peaks at the minimum of the function
1129: $W(z)-\varphi(z)$,
1130: where
1131: $$
1132: \varphi(z)= -\hbar\sum_{i=1}^N \log|z-z_i|^2
1133: $$
1134: is the Newton potential,
1135: logarithmic in 2-D, of the domain (the support of eigenvalues).
1136: In the semiclassical limit, the potential $\varphi$ is
1137: continuous. One
1138: writes the extremum condition in the form
1139: \begin{equation}\label{s}
1140: \p_z\left(\varphi - W\right)=
1141: \p_{\bar z}\left(\varphi - W\right) =0\,,
1142: \quad \quad \mbox{at}\quad z=z_i.
1143: \end{equation}
1144: The points where this
1145: equations holds are the most probable positions of
1146: eigenvalues $z_i$.
1147:
1148: In addition, if one assumes that $z_i$ have a continuous distribution
1149: (a strong assumption), i.e.,
1150: that
1151: the density
1152: \begin{equation}\label{rho}
1153: \rho_N(z)=\hbar\sum_{i=1}^N\delta(z-z_i)=- \,
1154: \frac{1}{4 \pi} \Delta\varphi
1155: \end{equation}
1156: is a smooth positive function in some domain $D$, and zero outside
1157: this domain, then
1158: it follows from (\ref{s}) that the density is $-\frac{1}{4 \pi}
1159: \Delta W$ inside the domain, if $-\Delta W$ is not singular and
1160: positive inside the support of eigenvalues.
1161:
1162: The shape of the domain is found in two steps.
1163: First, one finds the potential $\varphi$. Then, having the potential,
1164: one has to restore the shape of the
1165: domain. The second step
1166: is the {\it inverse potential problem}.
1167:
1168: Under this assumption, $\varphi$ is a harmonic function
1169: in the exterior of the support of eigenvalues. Then
1170: equation (\ref{s}) reads as a Riemann-Hilbert problem:
1171: \begin{description}
1172: \item[(i)]The boundary value of the analytic function $\p_z\varphi$ in
1173: the exterior of the domain $D$
1174: is $\p_z W$;
1175: \item[(ii)]
1176: The function $\p\varphi$ behaves at infinity as $-\hbar Nz^{-1}=-t/z$
1177: (we assume that there are no eigenvalues around infinity).
1178: \end{description}
1179: These conditions uniquely specify the domain if it is simply connected.
1180: In this case, $\pi\hbar N$ is the area of the domain.
1181: Multiply-connected domains are also possible solutions. The number of
1182: disconnected droplets
1183: cannot exceed the number
1184: of poles of $A(z)$ plus one. To find them one needs
1185: additional data.
1186: For example, one may
1187: fix the filling factors -- the number of eigenvalues
1188: that each droplet contains.
1189: Obviously, all the filling factors are positive and their sum is $N$.
1190: Domains constructed this way, provided
1191: $A(z)$ is a rational function, belong to a special class of
1192: {\it algebraic domains} \cite{Ahar-Shap}.
1193:
1194: An important property of the Riemann-Hilbert problem is that the function
1195: $\p\varphi(z)$, with $z$ inside the domain, does not
1196: depend on $N$, but rather on the potential. The former does not
1197: change during the growth.
1198:
1199:
1200: At this point, we specify the potential to be of the form
1201: (\ref{potential}). Then the boundary
1202: value of the analytic function $\p\varphi(z)$ is
1203: $\p W=-\bar z + A(z)$. The density of eigenvalues
1204: is uniform, and the filling factors fix
1205: the areas of the droplets
1206: to be $\pi\hbar\nu_\alpha$.
1207:
1208: The potential $\varphi(z)$ can be constructed with the help of
1209: the Schwarz function.
1210:
1211: \subsection{The Schwarz function and the inverse potential problem}
1212: Given a closed (in general a multiply-connected) domain
1213: $D$ in the plane,
1214: the {\it Schwarz function} $S(z)$
1215: is an analytic function
1216: in some neighborhood of the boundary of the domain with the
1217: value $\bar z$
1218: on the boundary \cite{Alhfors50}:
1219: \begin{equation}\label{Schwarz}
1220: S(z)=\bar z \;\;\;\;\;
1221: \mbox{on the contour}.
1222: \end{equation}
1223: The involution property of the Schwarz
1224: function follows from the definition (cf. (\ref{anti11})):
1225: \begin{equation} \nonumber
1226: \bar S(S(z))=z.
1227: \end{equation}
1228: Let us represent the Schwarz function in the form
1229: $S(z)=S_{+}(z)+S_{-}(z)$, where $S_{+}$ (respectively,
1230: $S_{-}$) is analytic inside (respectively, outside) $D$,
1231: with the condition that $S_{-}(\infty )=0$.
1232: Choose $S_+(z)$ to be equal to $A(z)$,
1233: \be\la{21}
1234: A(z)=\frac{1}{2\pi i}
1235: \oint_{\partial D}\frac{\bar\zeta d\zeta}{z-\zeta}, \quad z\in D.
1236: \ee
1237: Then the function $S_-(z)=S(z)-A(z)$, being analytic outside $D$,
1238: has the boundary value
1239: $\bar z-A(z)$. We conclude that
1240: $A(z)-S(z)=\p\varphi(z)$.
1241:
1242: Therefore, the problem of the previous section reads:
1243: given a potential $A(z)$, find the Schwarz function
1244: \be\la{123}
1245: S(z)=A(z)+S_-(z)
1246: \ee
1247: whose poles and residues outside of the
1248: support of eigenvalues are given by
1249: the poles and residues of $A(z)$ and
1250: such that $S_{-}(z) \, \to \, \hbar N/z$ as
1251: $z\to\infty$.
1252: If the Schwarz function
1253: is known, Eq. (\ref{Schwarz}) determines the domain. Inside the droplet
1254: $S_-(z)$, and therefore, $S(z)$ have cuts
1255: (Fig.~\ref{dropletandacut}).
1256: \begin{figure} \begin{center}
1257: \includegraphics*[width=5cm]{dropletandacut.eps}
1258: \caption{\label{dropletandacut} The boundary
1259: of a droplet and a cut of the
1260: Schwarz function
1261: located inside the droplet. }
1262: \end{center} \end{figure}
1263:
1264: A note is in order.
1265: In Sec.~\ref{Saddle} we assumed that the density of
1266: eigenvalues is a continuous function
1267: everywhere except for the boundary of $D$. As a consequence,
1268: $\varphi$ is a harmonic function, $\p\varphi$ and $S_-(z)$ are analytic
1269: functions
1270: outside $D$, and the Schwarz function
1271: outside the droplets has the same poles as $A(z)$.
1272: As a result, the complement to the
1273: support of eigenvalues appears to be an
1274: algebraic domain.
1275:
1276: It seems that this assumption is too restrictive. Moreover,
1277: the definition of the curve through the linear differential equations
1278: for the wave function of Sec. \ref{Curve1} does not forbid the
1279: physical branch of the Schwarz function
1280: to have branching points and cuts outside physical droplets, as shown
1281: in Figs.~\ref{Joukowsky}, \ref{hypotrochoid}.
1282: The semiclassical
1283: wave function still peaks on the real section of the curve (\ref{S1}),
1284: regardless of
1285: whether there are cuts outside the droplets, or if those cuts shrink
1286: to points.
1287: In other words, a solution with cuts outside
1288: physical droplets still has a
1289: local maximum on the boundary of the physical droplets and it is stable.
1290: It becomes unstable, however,
1291: in a vicinity of cuts along a normal direction to a cut.
1292: %where ${\mathcal A}_0$ has a saddle point.
1293:
1294: These arguments suggest
1295: that one can safely allow
1296: $\p\varphi (z)$ to have branching points, i.e., the Schwarz function is
1297: allowed to have
1298: cuts outside physical droplets. Then we seek for solutions of the
1299: Riemann-Hilbert problem
1300: of Sec. \ref{Saddle} in a wider class of functions.
1301: This leads to a wider class of domains (not algebraic). However, in this case,
1302: one is no longer able to
1303: treat eigenvalues as continuously distributed.
1304: We will come back to this issue in
1305: the next section.
1306:
1307:
1308:
1309:
1310:
1311:
1312: \section {The Schwarz function and its Riemann surface. The Schottky
1313: double.}\label{S}
1314:
1315: The Schwarz function describes more than just the boundary of
1316: clusters of eigenvalues.
1317: Together with other sheets
1318: it defines a Riemann surface.
1319: If the potential $A(z)$ is meromorphic,
1320: the Schwarz function is an algebraic function. It
1321: satisfies a polynomial equation $f(z, S(z))=0$.
1322:
1323:
1324: \begin{figure} \begin{center}
1325: \includegraphics*[width=5cm]{double.eps}
1326: \caption{\label{Schottky} The Schottky double. A Riemann surface with
1327: boundaries
1328: along the droplets (a front side) is
1329: glued to its mirror image (a back side).}
1330: \end{center} \end{figure}
1331:
1332:
1333:
1334:
1335:
1336: The function $f(z,\tilde z)$, where $z$
1337: and $\tilde z$ are treated as two independent complex arguments, defines
1338: a Riemann surface with antiholomorphic involution (\ref{anti1}).
1339: If the involution divides the surface into two disconnected
1340: parts, as explained above,
1341: the Riemann surface is
1342: the {\it Schottky double} \cite{C,SS} of
1343: one of these parts.
1344:
1345: There are two complementary ways to describe
1346: this surface. One is through the algebraic covering
1347: (\ref{h1}, \ref{ah1}). Among $d$ sheets we distinguish
1348: a {\it physical} sheet.
1349: The physical sheet is
1350: selected by the condition that the differential $S(z)dz$ has the same
1351: poles and residues as
1352: the differential of the potential $A(z)dz$ (as in (\ref{123})).
1353: It may happen that the condition
1354: $\bar z=S^{(i)}(z)$ defines a planar curve (or
1355: several curves, or a set of isolated points)
1356: for branches
1357: other than the physical one. We refer to the interior of these
1358: planar curves as {\it virtual} (or
1359: unphysical) droplets
1360: situated on sheets other than physical.
1361:
1362:
1363:
1364: Another way emphasizes the antiholomorphic involution. Consider a
1365: meromorphic function $h(z)$ defined
1366: on a Riemann surface with boundaries. We call this surface the front side.
1367: The Schwarz reflection principle extends any meromorphic function on
1368: the front side
1369: to a meromorphic function on the Riemann surface without
1370: boundaries. This is done
1371: by adding another copy of the Riemann surface with
1372: boundaries (a back side),
1373: glued to the front side along the boundaries, Fig~\ref{Schottky}.
1374: The value of the
1375: function $h$ on the mirror point on the back side
1376: is $h(\overline{S(z)})$.
1377: The copies are glued along the boundaries:
1378: $h(z)=h(\overline{S(z)})$
1379: if the point $z$ belongs to the boundary.
1380: The same extension rule applies to differentials.
1381: Having a meromorphic differential $h(z)dz$ on the
1382: front side, one extends it to a
1383: meromorphic differential
1384: $h(\overline{S( z)})d\overline{S(z)}$ on the back side.
1385:
1386: This definition can be applied to the Schwarz function itself.
1387: We say that the Schwarz
1388: function on the double is $S(z)$ if the point is
1389: on the front side, and
1390: $\bar z$ if the point belongs to the back side
1391: (here we understand $S(z)$
1392: as a function defined on the complex curve, not just on the physical sheet).
1393:
1394:
1395: The number of sheets of the curve is the number of poles (counted
1396: with their multiplicity)
1397: of the function $A(z)$ plus one.
1398: Indeed, according to (\ref{123}) poles of $A$ are poles of
1399: the Schwarz function on the front side of the double. On the
1400: back side, there is also a pole at
1401: infinity. Since $S(z=\infty)=A(\infty)$, we have
1402: $\bar S(\bar z=A(\infty))=\infty$.
1403: Therefore, the factor $a(z)$ is a polynomial
1404: with zeros at the poles of $A(z)$
1405: and at $\overline{A(\infty)}$, and
1406: $$d\equiv\mbox{number of sheets} =
1407: \mbox{number of poles of $A$ + 1}.$$
1408: The front and back sides meet at planar curves $\bar z=S(z)$.
1409: These curves are
1410: boundaries of the droplets. We repeat that not all droplets
1411: are physical.
1412: Some of them may belong to unphysical sheets, Fig.~\ref{torus}.
1413:
1414: Boundaries of droplets, physical and virtual, form a subset of the
1415: $\bf a$-cycles on the curve. Their number
1416: cannot exceed the genus of the
1417: curve plus one:
1418: $$\mbox{number of droplets} \le g+1.$$
1419:
1420:
1421: The sheets meet along cuts located inside droplets.
1422: The cuts that belong to physical droplets show up on unphysical sheets.
1423: On the other hand, some
1424: cuts show up on the physical sheet (Fig~\ref{torus}).
1425: They correspond to droplets situated on
1426: unphysical sheets.
1427: \begin{figure} \begin{center}
1428: \includegraphics*[width=5cm]{torus.eps}
1429: \caption{\label{torus}Physical and unphysical droplets on a torus.
1430: The physical sheet (shaded) meets the unphysical sheet
1431: along the cuts. The cut situated inside the
1432: unphysical droplet appears on the physical sheet.
1433: The boundaries of the droplets (physical and virtual)
1434: belong to different sheets.
1435: This torus is the Riemann surface corresponding to the ensemble
1436: with the potential $V(z) = - \alpha \log (1- z/ \beta) - \gamma z $.}
1437: \end{center} \end{figure}
1438:
1439: The Riemann-Hurwitz theorem computes the genus of the curve as
1440: $$ g=\mbox{half
1441: the number of
1442: branching points} - d+1. $$
1443:
1444: With the help of the Stokes formula, the numbers $\{ \nu_{\alpha} \}$
1445: are identified with areas of
1446: the droplets:
1447: $|\nu_a|=\frac{1}{2\pi \hbar}\int_{D_a} d^2z$.
1448: For a nondegenerate curve,
1449: these numbers are not necessarily positive. Negative
1450: numbers correspond to droplets located on unphysical sheets. In this
1451: case, $\{\nu_a\}$
1452: do not correspond to the number of eigenvalues located inside each droplet,
1453: as it is the case for algebraic domains, when all
1454: filling numbers are positive.
1455:
1456: A comment is in order. In the Hermitian or unitary
1457: one-matrix ensembles, the
1458: construction of the Riemann
1459: surface is different (Refrs. \cite{DV03,Chekhov,David,Kostov}). In
1460: the Hermitian random matrix
1461: ensemble, for example,
1462: eigenvalues are real. The support of eigenvalues is a set of
1463: disjoint intervals on the real line.
1464: In this case, branch cuts are identical to the support of
1465: eigenvalues. The Riemann surface
1466: is constructed by gluing $g$ sheets along $g+1$ cuts \cite{Kostov}.
1467: Conversely, in the Schottky double the cuts generally do not touch
1468: boundaries of the droplets (Fig.~\ref{dropletandacut}).
1469: Complex curves for the Hermitian or unitary ensembles are simpler.
1470: They are always hyperelliptic.
1471:
1472: \subsection{Degeneration of the spectral curve}
1473: Degeneration of the complex curve gives the most interesting physical
1474: aspects of growth.
1475: There are several levels of degeneration. We briefly discuss them below.
1476:
1477: \subsubsection{Algebraic domains and double points}\la{alg}
1478: A special case occurs when the Schwarz function on the physical sheet
1479: is meromorphic. It
1480: has no other singularities than poles of $A$.
1481: This is the case of algebraic domains \cite{Ahar-Shap}.
1482: They appear in the
1483: semiclassical case (see Sec.~\ref{SCurve}).
1484: This situation occurs if cuts
1485: on the physical sheet, situated outside physical droplets, shrink
1486: to points, i.e., two or more
1487: branching points merge. Then the physical sheet meets other
1488: sheets along cuts situated inside physical droplets only and also at
1489: some points on their exterior ({\it double points}) (see figure captions for
1490: Figs.~\ref{Joukowsky}, \ref{hypotrochoid}).
1491: In this case the Riemann surface degenerates. The genus
1492: is given by the number of physical
1493: droplets only. The filling factors (\ref{18}) are all positive.
1494:
1495: \begin{figure} \begin{center}
1496: \includegraphics*[width=5cm]{degtorus.eps}
1497: \caption{\label{degtorus}Degenerate torus corresponds to the
1498: algebraic domain for the Joukowsky map.}
1499: \end{center} \end{figure}
1500:
1501:
1502: In the case of algebraic domains, the physical branch of the
1503: Schwarz function
1504: is a well-defined
1505: meromorphic function. Analytic continuations of $\bar z$
1506: from different disconnected parts of the boundary
1507: give the same result. In this case, the
1508: Schwarz function can be written through the Cauchy transform of the
1509: physical droplets:
1510: \begin{equation}\label{28}
1511: S(z)=A(z)+\frac{1}{\pi}\int_{D}\frac{d^2 \zeta}{z-\zeta}.
1512: \end{equation}
1513:
1514:
1515: Although algebraic domains occur in physical problems
1516: such as Laplacian growth,
1517: their semiclassical evolution is limited.
1518: Almost all algebraic domains will be broken in a growth process.
1519: Within a finite
1520: time (the area of the domain) they degenerate further into
1521: critical curves
1522: (Fig.~\ref{evolutionhyp}, \ref{criticalwing}).
1523: The Gaussian potential (the Ginibre-Girko ensemble),
1524: which leads to a single
1525: droplet of the form of an ellipse is a known exception.
1526: \begin{figure} \begin{center}
1527: \includegraphics*[width=5cm]{criticalwing.eps}
1528: \caption{\label{criticalwing}Critically degenerate Joukowsky map.}
1529: \end{center} \end{figure}
1530:
1531: \subsubsection{Critical degenerate curves}
1532:
1533: Algebraic domains appear as
1534: a result of merging of simple branching points on the physical
1535: sheet. The double points are located
1536: outside physical droplets.
1537: Remaining branching points belong to
1538: the interior of
1539: physical droplets. Initially, they survive
1540: in the degeneration process.
1541: However, as known in the theory of Laplacian growth,
1542: the process necessarily leads to a further
1543: degeneration. Sooner or later, at least one of the
1544: interior branching points merges with one of the double points in
1545: the exterior. Curves degenerated in this
1546: manner are called
1547: {\it critical}. For the genus one and three
1548: this degeneration is discussed below.
1549: The critical degenerations
1550: are depicted, respectively, in Fig.~\ref{criticalwing} and Fig.~\ref{evolutionhyp}.
1551:
1552: Since interior branching points
1553: can only merge with exterior branching points
1554: on the boundary of the
1555: droplet, the boundary develops a cusp, characterized by a pair $p, q$
1556: of mutually prime integers. In local coordinates
1557: around such a cusp, the curve looks like $x^p\sim y^q$.
1558:
1559: The fact that the growth of algebraic domains always leads to critical
1560: curves is known
1561: in the theory of Laplacian growth (see e.g., \cite{review}) as
1562: finite time singularities. It is also
1563: known in the Hermitian one- and two-matrix
1564: models as
1565: critical points (intensively
1566: studied in 2-D-gravity \cite{gravity}).
1567:
1568: The degeneration process seems to be a feature of the semiclassical
1569: approximation. Curves treated beyond this approximation never
1570: degenerate.
1571:
1572:
1573: \subsubsection{Simply-connected domains. Conformal maps.}\la{C}
1574: The case when the complement to the support of eigenvalues is a
1575: simply-connected algebraic domain
1576: in the extended (i.e. including $\infty$) complex plane
1577: is particularly important.
1578: All the filling factors are zero except
1579: one, which is equal to $N$, so there is only one droplet $D$.
1580: In this case, the Schottky double of the
1581: exterior of the droplet is a Riemann sphere.
1582:
1583: The exterior domain can be
1584: conformally and univalently mapped onto the exterior of the unit disc.
1585: We call this map $w(z)$, and its inverse
1586: $z(w)$.
1587: For algebraic domains, the inverse map is
1588: a rational function of $w$.
1589: Choosing the normalization such that
1590: $z(\infty )=\infty$, we represent the
1591: inverse map
1592: by a half-infinite Laurent polynomial
1593: \begin{equation}\la{291}
1594: z(w)=r w+\sum_{k>0}u_{k} w^{-k},\quad |w|\geq 1.
1595: \end{equation}
1596: The coefficient $r$, called
1597: the external conformal radius of the
1598: domain $D$, is chosen to be real positive.
1599: The Schwarz reflection of the inverse map is
1600: \begin{equation}\la{33}
1601: \bar z(w^{-1})=r w^{-1}+\sum_{k>0}{\bar u^{(k)}}{w^k},\quad |w|\geq 1.
1602: \end{equation}
1603: Given a point $z$, there are $d$ values of $w$ solving the
1604: equation (\ref{291}), where $d$ is the number of poles
1605: of the rational function $z(w)$. Among them,
1606: only one solution corresponds to
1607: the conformal map. It is the solution
1608: such that $w\to z/r$ as $z\to\infty$.
1609: The global coordinate $w$ on the Riemann sphere
1610: provides a uniformization of
1611: the Schottky double. In this coordinate, the involution
1612: reads $w\to 1/\bar w$.
1613: The Schwarz function is
1614: \be\la{ss}S(z)=\bar z(w^{-1}(z)).
1615: \ee
1616:
1617: Choose a regular point $\xi_0$ of the potential inside the
1618: droplet and expand
1619: \begin{equation}\la{}
1620: V(z)=\sum_{k>0} t_k (z-\xi_0)^{k}.
1621: \end{equation}
1622: The coefficients $t_k$ have a simple geometric
1623: interpretation. It follows from (\ref{21}) that they are
1624: {\it harmonic moments}
1625: of the exterior of the droplet
1626: (with respect to the point $\xi_0$):
1627: \begin{equation}\label{im2}
1628: t_k =-\frac{1}{\pi k}\int_{{\bf C}\setminus D}(z-\xi_0)^{-k}d^2 z.
1629: \end{equation}
1630: We also mention the area formula
1631: \be\la{area}
1632: t=r^2-\sum_{k\geq 1}k|u^{(k)}|^2.
1633: \ee
1634:
1635: If the vector potential $A(z)$ is
1636: a polynomial of degree $d-1$, then
1637: the inverse map $z(w)$ is a (Laurent) polynomial:
1638: $u^{(k)}=0,\quad k>d-1$. If $A(z)$
1639: is a rational function with $d-1$ simple poles
1640: in the finite part of the complex plane,
1641: the map $z(w)$ is a rational function
1642: with $d-1$ simple poles and one extra simple pole at infinity.
1643:
1644: \subsection{The generating differential}\la{Diff}
1645:
1646: The meromorphic differential
1647: \be\la{Omega}
1648: d\Omega=S(z)dz
1649: \ee
1650: plays an important role.
1651: It is called generating
1652: differential \cite{mkwz,kkmwz}.
1653: According to (\ref{123}), on the physical sheet
1654: it has the same poles and residues
1655: as the differential $Adz$:
1656: \be\la{1234}
1657: d\Omega=Adz+S_-(z)dz.
1658: \ee
1659:
1660: Below we use the following properties of the
1661: generating differential.
1662: \begin{itemize}
1663: \item [(i)]
1664: The periods over $\bf a$-cycles
1665: (boundaries of the droplets) are
1666: purely imaginary, and are integer multiples
1667: of $2\pi i$. They compute areas
1668: of the droplets (the filling factors (\ref{18})):
1669: $$
1670: \nu_{\alpha}=\frac{1}{2\pi i \hbar}\oint_{{\bf a}_{\alpha}}
1671: d\Omega.
1672: $$
1673: The filling factors of physical droplets
1674: (belonging to the physical sheet)
1675: are positive.
1676: \item[(ii)]
1677: The real part of the integral of the differential
1678: $(\bar z-S(z))dz$ from some fixed
1679: point $\xi_0$ to a point on the boundary
1680: of a droplet (a point on a $\bf a$-cycle)
1681: has the same value for all points of the boundary:
1682: \be \nonumber
1683: \phi_\alpha=-|z|^2+2 {\mathcal R}e\int_{\xi_0}^z d\Omega=
1684: \mbox{const,\quad\quad for all}\quad z\in {\bf a}_\alpha.
1685: \ee
1686: This quantity does not depend on $z$,
1687: but does depend on $\xi_0$ unless $\xi_0$ is on the boundary.
1688: However, the difference $\phi_{\alpha}-\phi_{\beta}$
1689: depends on the ${\bf a}$-cycles only.
1690: It is equal to a $\bf b$-period of the differential
1691: $d\Omega$:
1692: \be\la{32}
1693: \phi_\alpha-\phi_\beta=\oint_{{\bf b}_{\alpha\beta}}d\Omega,
1694: \ee
1695: where ${\bf b}_{\alpha\beta}$ is a
1696: cycle connecting ${\bf a}_{\alpha}$ and
1697: ${\bf a}_{\beta}$ cycles.
1698: \end{itemize}
1699: For proofs and more details, see Ref. \cite{mkwz}.
1700:
1701: Periods over $\bf b$-cycles $\phi_\alpha-\phi_0$ play a role of
1702: chemical potentials for the filling factors.
1703: Here $0$ denotes a chosen reference droplet.
1704: One can use chemical potentials to
1705: characterize evolution of the curve instead
1706: of filling factors.
1707:
1708: \section{Examples}\label{Example}
1709:
1710: \subsection{Genus zero -- ellipse}\la{classicalellipse}
1711:
1712: The potential is Gaussian, $V(z)= t_2 z^2$,
1713: $2|t_2| <1$, $A(z)=2t_2 z$. It has a simple
1714: pole at infinity.
1715: The number of sheets
1716: of the Schwarz function is two.
1717: At infinity, $S'(z)$ is finite. Therefore,
1718: $S'(z)$ takes every value
1719: twice. It must have two zeros at most, i.e.,
1720: there are two branching points.
1721: The Riemann-Hurwitz theorem says that genus is zero.
1722: The curve with two branching points and two sheets
1723: has the form
1724: $f(z,\bar z)=z\bar z+k_1z^2+\bar k_1\bar z^2+k_2=0$. Conditions that
1725: $S(z)\sim 2t_2z+n\hbar z^{-1}+\dots$ at $z\to\infty$ for the
1726: physical branch and the unitarity
1727: condition (\ref{anti11}) determine the curve in full:
1728: \be\la{curveellipse}
1729: f_n(z,\bar z)=z\bar z - ( t_2z^2+ \bar t_2\bar z^2)
1730: \frac{2}{1+4|t_2|^2}-n\hbar\frac{1-4|t_2|^2}{1+4|t_2|^2}=0.
1731: \ee
1732: The physical branch
1733: $$
1734: S^{(1)}(z)=2t_2 z+\frac{(2\bar
1735: t_2)^{-1}-2t_2}{2}z
1736: \left (1-\sqrt{1-\frac{2}{(2\bar
1737: t_2)^{-1}-2t_2}\frac{2n\hbar}{z^2}}\right )
1738: $$
1739: is the Schwarz function of ellipse.
1740: The second branch $S^{(2)}(z)\to (2\bar t_2)^{-1}z$
1741: does not correspond to any real curve.
1742: The droplet (see Fig.~\ref{ellipse}) is an ellipse with the
1743: quadrupole moment $2|t_2|$ and area $\pi t=\pi n\hbar$.
1744:
1745:
1746:
1747: Recurrence relations
1748: Eqs. (\ref{M11}, \ref{M}) truncate
1749: \be\la{59}
1750: z\psi_n=r_n\psi_{n+1}+u_n\psi_{n-1},\quad\quad
1751: (L^{\dag}\psi)_n=r_{n-1}\psi_{n-1}+\bar u_{n+1}\psi_{n+1}.
1752: \ee
1753: Equations (\ref{M1}, \ref{Area}) imply
1754: \be\la{58}
1755: 2t_2=\frac{\bar u_{n+1}}{r_n},\quad n\hbar=r_n^2-|u_{n+1}|^2,
1756: \ee
1757: where the second equation is an analog of the area formula (\ref{area}).
1758: Together they give a growth
1759: equation for the quantum analog of conformal radius
1760: $n\hbar=(1-|2t_2|^2)r_{n}^2$.
1761:
1762: The operator $L^{\dag}$ can be cast in the form of a $2\times 2$
1763: matrix (the number of sheets of the curve). Writing (\ref{59}) for
1764: $n+1$, we express $\chi_{n-1}$ and
1765: $\chi_{n+2}$ through $\underline\chi_n=(\chi_n,\,\chi_{n+1})$ from the
1766: first pair, and substituting them to the second pair we obtain
1767: \be\la{601}
1768: {\mathcal L}_n \underline\chi_n=
1769: \left (
1770: \begin{array}{cc}
1771: z\frac{ r_{n-1}}{u_n} & \bar u_{n+1} - \frac{r_n r_{n-1}}{u_n} \\
1772: r_n- \frac{\bar u_{n+2}u_{n+1}}{r_{n+1}} &
1773: \frac{\bar u_{n+2}}{r_{n+1}}z
1774: \end{array}
1775: \right)\underline\chi_n.
1776: \ee
1777: Formally, one can say that the recurrence relation (\ref{59})
1778: represents the shift operator as the $2\times 2$ matrix
1779: $$
1780: {\mathcal W}_{n-1}=
1781: \left (
1782: \begin{array}{cc}
1783: 0 & 1 \\
1784: -\frac{u_n}{r_n} &
1785: \frac{z}{r_n}
1786: \end{array}
1787: \right).
1788: $$
1789: The relation ${\mathcal L}_n =\left (
1790: \begin{array}{cc}
1791: r_{n-1} & 0 \\
1792: 0 & r_{n}
1793: \end{array}
1794: \right){\mathcal W}^{-1}_{n-1}+\left (\begin{array}{cc}
1795: \bar u_{n+1} & 0 \\
1796: 0 & \bar u_{n+2}
1797: \end{array}
1798: \right){\mathcal W}_n
1799: $ then
1800: yields (\ref{601}).
1801: With the help of (\ref{58}), ${\mathcal L}_n$ reads
1802: $$
1803: {\mathcal L}_n (z)=
1804: \left (
1805: \begin{array}{cc}
1806: z(2\bar t_2)^{-1} & -r_n(2\bar t_2)^{-1}(1-4|t_2|^2) \\
1807: r_n(1-4|t_2|^2) & 2t_2 z
1808: \end{array}
1809: \right).
1810: $$
1811: Computing the determinant $\det ({\mathcal L}_n (z) - \bar z)$, we obtain the
1812: curve (\ref{curveellipse}), already
1813: determined by the singularities.
1814:
1815: The recurrence relation can be used to generate the
1816: Hermite polynomials.
1817: The model with potential $V(z) = t_2 z^2 + Q\log z$ can
1818: also be solved explicitly (Ref. \cite{Akemann02}). It
1819: generates the Laguerre polynomials. These are the cases where
1820: biorthogonal polynomials are classical orthogonal
1821: polynomials.
1822:
1823: Classical polynomials correspond to a
1824: genus zero curve. Higher genus curves generate more
1825: complicated, but more interesting polynomials.
1826:
1827: The classical limit of the recurrence relations
1828: gives a conformal map of the exterior
1829: of the unit disk to the exterior of the ellipse,
1830: $z(w)=rw+\frac{u}{w}$.
1831: The double-valued function
1832: $
1833: w_{1,2}(z) = \frac{1}{2r} \left [
1834: z \pm \sqrt{(z -z_1)(z-z_2) } \right ],\quad z_{1,2}=\pm 2\sqrt{ur}
1835: $
1836: becomes single-valued on a two-sheet covering
1837: of the $z$-sphere plane. The branch $w_1(z)$ is
1838: such that $w_1\to\infty$ as $z\to\infty$.
1839: It defines the inverse map from the ellipse exterior
1840: onto the exterior of the unit disk in the $w$-plane.
1841: The function
1842: $
1843: \bar z(w^{-1})=rw^{-1} + \bar u w
1844: $
1845: is a meromorphic function of
1846: $w$ with one simple pole at $w=0$. Treated as functions of $z$, the
1847: two sheets $S^{(1,2)}(z)=\bar z(w_{1,2}^{-1}(z))$
1848: have two critical points $z_{1,2}$ and constitute a sphere.
1849: Parameters of the map are related to the parameters of the potential
1850: through the classical limit
1851: of (\ref{58}): $t_2=\frac{\bar
1852: u}{2r},\quad t=r^2-|u|^2$.
1853: The general formula for the semiclassical behavior of the wave function
1854: (\ref{sp1})
1855: gives the known asymptotics of the
1856: Hermite polynomials \cite{BE}.
1857:
1858: The cut inside the droplet may shrink to a point.
1859: In this case only one (physical) sheet remains,
1860: and the Schwarz function becomes a rational
1861: function with one simple pole.
1862: The wave functions are monomials.
1863: The droplet is a disk.
1864:
1865: A more interesting degeneration
1866: occurs at the critical value of $|t_2|=\frac{1}{2}$,
1867: when the two branching points
1868: reach the boundary of the droplet.
1869: The ellipse degenerates into the cut $[z_1,\,z_2]$.
1870:
1871: Beyond the critical value
1872: ($|t_2|> \frac{1}{2}$), the ellipse appears again, but on
1873: the unphysical sheet. In this case, there is no physical droplet,
1874: but there is a virtual droplet. At this value of $t_2$, the matrix
1875: integral diverges,
1876: but a solution of the differential equation (\ref{M2}) exists. It
1877: is not a polynomial,
1878: however.
1879:
1880:
1881: Below we use a more direct method to find the quantum curve
1882: bypassing the matrix
1883: representation of $L^{\dag}$. The method reflects the triangular
1884: structure of $L^{\dag}$,
1885: and closely resembles the procedure used in the classical case.
1886: We will use this method for the following, more complicated, examples.
1887:
1888:
1889:
1890:
1891: Apply (\ref{59}) to an eigenvector $c_n$ of the matrix
1892: ${\mathcal L}_n (z)$ with an eigenvalue $\tilde z$. The equations
1893: \be \nonumber
1894: \left
1895: \{
1896: \begin{array}{lcl}
1897: z c_{n}& = & r
1898: _n c_{n+1} + u_n c_{n-1} \\
1899: \tilde z c_{n} & = & \bar u_{n+1} c_{n+1} + r_{n-1} c_{n-1}
1900: \end{array}
1901: \right.
1902: \ee
1903: are compatible if the point $(z, \tilde z)$
1904: belongs to the curve. To obtain a
1905: compatibility condition,
1906: we express $c_{n+1}$ and $c_{n-1}$ in terms of $c_n$. The
1907: results must differ
1908: by a shift $n\to n+2$
1909: \be \nonumber
1910: \det
1911: \Big |
1912: \begin{array}{cc}
1913: z & u_n \\
1914: \bar z & r_{n-1}
1915: \end{array}
1916: \Big |\det
1917: \Big |
1918: \begin{array}{cc}
1919: r_{n+1} & z \\
1920: \bar u_{n+2} & \bar z
1921: \end{array}
1922: \Big |= d_nd_{n+1},
1923: \quad
1924: d_n = \det
1925: \Big |
1926: \begin{array}{cc}
1927: r_n & u_n \\
1928: \bar u_{n+1} & r_{n-1}
1929: \end{array}
1930: \Big |.
1931: \ee
1932: This gives the curve (\ref{curveellipse}).
1933:
1934:
1935: \subsection{Genus one -- an aircraft wing}\la{J}
1936:
1937: The potential is
1938: $V(z) = - \alpha \log (1- z/\beta) -
1939: \gamma z,\quad A(z)=-\frac{\alpha}{z-\beta}-\gamma$.
1940: There is one pole at $z=\beta$ on the first (physical) sheet.
1941: At $z=\infty$ on the first sheet
1942: $S(z)\to-\gamma+\frac{n\hbar-\alpha}{z}$.
1943: Therefore, the Schwarz function has another pole
1944: at the point $-\bar\gamma$
1945: on another sheet. All the poles are simple.
1946: According to the general arguments of
1947: Sec.~\ref{S}, the number of sheets is 2,
1948: the number of branching points is 4. The genus is 1.
1949: The curve has the form
1950: $$
1951: f(z, \bar z) = z^2 \bar z^2 +k_1
1952: z^2\bar z +\bar k_1 z\bar z^2 + k_2 z^2 + \bar k_2 \bar z^2 +
1953: k_3 z\bar z
1954: +k_4z+
1955: \bar k_4\bar z +h =0.
1956: $$
1957: The points at infinity and $-\bar\gamma$ belong to
1958: the second
1959: sheet of the algebraic covering. Summing up,
1960: \[ S(z)=
1961: \left \{\begin{array}{rll}
1962: -\frac{\alpha}{z-\beta}& \mbox{as} & z\to\beta_1, \\
1963: (-\gamma+\frac{n\hbar-\alpha}{z} ) & \mbox{as} & z\to\infty_1, \\
1964: \frac{n\hbar-\bar\alpha}{z+\bar\gamma}& \mbox{as}
1965: & z\to -\bar\gamma_2, \\
1966: (\bar\beta-\frac{\bar\alpha}{z})& \mbox{as} & z\to\infty_2.
1967: \end{array}
1968: \right. \]
1969: where, by 1 and 2 we indicate the sheets.
1970:
1971:
1972:
1973:
1974: Poles and residues of the Schwarz function
1975: determine all the coefficients of the curve
1976: $f(z,\bar z)=a(z) (\bar z-S^{(1)}(z))(\bar z-S^{(2)}(z))=
1977: \overline{a(z)} (z-\bar S^{(1)}(\bar z))(z-\bar
1978: S^{(2)}(\bar z))$ except one.
1979: The behavior at $\infty$ of $z, \bar z$
1980: gives $k_1=\gamma -\bar \beta, \, k_2=-\gamma \bar \beta$.
1981: Hereafter we choose the origin by
1982: setting $\gamma=0$.
1983: The equation of the curve then reads $f_n(z, \bar z) = 0$, where
1984: $f_n(z, \bar z) $ is given by
1985: \be\la{8}
1986: z^2 \bar z^2 -
1987: z^2\bar z \bar\beta -z\bar z^2\beta +
1988: \left ( |\bar\beta|^2 +\alpha+\bar\alpha-n\hbar
1989: \right ) z\bar z
1990: +z\bar\beta(n\hbar-\alpha)+
1991: \bar z\beta(n\hbar-\bar\alpha)
1992: +h_n
1993: \ee
1994: The free term $h_n$ is to be determined by
1995: filling factors of the two
1996: droplets $\nu_1$ and $\nu_2=n-\nu_1$.
1997: A detailed analysis shows
1998: that the droplets belong to different
1999: sheets (Fig.~\ref{torus}).
2000: Therefore, $\nu_2$ is negative.
2001:
2002: A boundary of a physical droplet is given by the
2003: equation
2004: $\bar z=S^{(1)}(z)$ (Fig.~\ref{Joukowsky}).
2005: The second droplet
2006: belongs to the unphysical sheet. Its
2007: boundary is given by
2008: $\bar z=S^{(2)}(z)$. The explicit form of both branches is
2009: $$S^{(1,2)}=\frac{1}{2}\bar\beta-
2010: \frac{\beta(n\hbar-\bar\alpha)+(\alpha+\bar\alpha-n\hbar)z\mp
2011: \sqrt{(z-z_1)(z-z_2)(z-z_3)(z-z_4)}}{2(z-\beta)z},$$
2012: where the branching points $z_i$ depend on $h_n$.
2013:
2014: If the filling factor of the physical droplet is equal to $n$, the cut
2015: inside the unphysical droplet is of the order of
2016: $\sqrt{\hbar}$. Although it never vanishes, it
2017: shrinks to a double point
2018: $z_3=z_4=z_*$ in a semiclassical limit.
2019: The sheets meet at the double
2020: point $z_*$ rather than along the cut:
2021: $\sqrt{(z-z_1)(z-z_2)(z-z_3)(z-z_4)}\to
2022: (z-z_*)\sqrt{(z-z_1)(z-z_2)}$.
2023: In this case, genus of the curve reduces to zero and
2024: the exterior of the physical droplet becomes an algebraic domain.
2025: This condition determines $h$,
2026: and also the position of the double point (Fig.~\ref{degtorus}).
2027: The double point is a saddle point for the
2028: level curves of $f(z,\bar z)$.
2029: If all the parameters are real, the double point is stable in
2030: $x$-direction and unstable in
2031: $y$-direction. A saddle point is a signature of a virtual droplet
2032: \cite{DV03}.
2033:
2034: If this solution is chosen,
2035: the exterior of the physical droplet can be mapped
2036: to the exterior of the unit disk by the Joukowsky map
2037: \be\la{222}
2038: z(w)=rw+u_0 + \frac{u}{w-a},\quad |w|>1,\quad |a|<1.
2039: \ee
2040: The inverse map
2041: is given by the branch $w_1(z)$ (such that $w_1\to\infty$ as
2042: $z\to\infty$)
2043: of the double valued function
2044: $$
2045: w_{1,2}(z) = \frac{1}{2r} \left [
2046: z-u_0 + ar \pm \sqrt{(z -z_1)(z-z_2) } \right ],\quad
2047: z_{1,2}=u_0+ar\mp 2\sqrt{r(u+au_0)}.
2048: $$
2049:
2050: The function
2051: \be\la{23}
2052: \bar z(w^{-1})=rw^{-1}+\bar u_0 + \frac{\bar u}{w^{-1}-\bar a}
2053: \ee
2054: is a meromorphic function of
2055: $w$ with two simple poles at $w=0$ and $w=\bar a^{-1}$. Treated as a
2056: function of $z$, it covers
2057: the $z$-plane twice. Two branches of the Schwarz function are
2058: $S^{(1,2)}(z)=\bar z(w_{1,2}^{-1}(z))$. On the physical sheet,
2059: $S^{(1)}(z)=\bar z(w_1(z))$ is the analytic continuation
2060: of $\bar z$ away from the boundary. This function is
2061: meromorphic outside the droplet. Apart from
2062: a cut between the branching points
2063: $z_{1,2}$, the sheets also meet at
2064: the double point $z_*=-\bar\gamma+a^{-1}re^{2i\phi}$,
2065: where $S^{(1)}(z_*)=S^{(2)}(z_*), \phi = \arg (ar+\frac{u\bar a}{1-|a|^2})$.
2066:
2067: Analyzing singularities of the Schwarz function,
2068: one connects parameters of
2069: the conformal map with the deformation parameters:
2070: \be\la{27}
2071: \left \{
2072: \begin{array}{rcl}
2073:
2074: \gamma & = & \frac{\bar u}{\bar a} - \bar u_0, \\
2075: n\hbar -\bar\alpha & = & r^2 -\frac{ur}{a^2}, \\
2076:
2077: \beta & = & \frac{r}{\bar a} + u_0 + \frac{u \bar a}{1-|a|^2}
2078: \end{array}\right.
2079: \ee
2080: $$
2081: \mbox{Area of the droplet}\sim n\hbar=r^2-\frac{|u|^2}{(1-|a|^2)^2}.
2082: $$
2083:
2084: A critical degeneration occurs when the double point merges with a
2085: branching point located inside the droplet
2086: ($z_*=z_2$) to form a triple point $z_{**}$.
2087: This may happen on the boundary only.
2088: At this point, the boundary has a $(2,\,3)$
2089: cusp (Fig.~\ref{criticalwing}).
2090: In local coordinates, it is
2091: $x^2\sim y^3$. This is a critical point of the conformal map:
2092: $w'(z_{**})=\infty$.
2093: A critical point
2094: inevitably results from the evolution
2095: at some finite critical area.
2096:
2097:
2098:
2099: A direct way to obtain the
2100: complex curve from the conformal map is the following. First,
2101: rewrite (\ref{222}) and (\ref{23}) as
2102: \be\la{22}\left
2103: \{
2104: \begin{array}{lcl}
2105: z-u_0+ar & = & rw+a(z+\bar\gamma)w^{-1}\\
2106: \bar z-\bar u_0+\bar a r & = & rw^{-1}+\bar a(\bar z+\gamma)w,
2107: \end{array}
2108: \right.
2109: \ee
2110: and treat $w$ and $1/w$ as independent variables.
2111: Then impose the condition $w\cdot w^{-1} = 1$. One obtains
2112: $$
2113: \left |\det
2114: \left [
2115: \begin{array}{cc}
2116: z- u_0 + ar & a(z+\gamma) \\
2117: \bar z -\bar u_0 + \bar a r & r
2118: \end{array}
2119: \right ]\right |^2
2120: =\left (\det
2121: \left [
2122: \begin{array}{cc}
2123: r & a(z+\bar\gamma) \\
2124: \bar a( \bar z+\gamma) & r
2125: \end{array}
2126: \right ]\right ) ^2.
2127: $$
2128: This gives the equation of the curve and in particular $h$,
2129: in terms of $u,\, u_0,\,r,\,a$ and eventually through the deformation
2130: parameters $\alpha,\,\beta,\,\gamma$ and $t$.
2131:
2132:
2133:
2134: The semiclassical analysis gives a guidance for the form of the
2135: recurrence relations.
2136: Let us use an ansatz for the $L$-operator, which resembles
2137: the conformal map (\ref{222}):
2138: $$
2139: L=r_n \hat w +u_{n}^{(0)} + (\hat w -a_n )^{-1} u_n,
2140: $$
2141: so that
2142: \begin{eqnarray}
2143: (\hat w -a_n ) L = (\hat w -a_n ) r_n \hat w +
2144: (\hat w - a_n ) u^{(0)}_n + u_n, \la{701} \\
2145: L^{\dag}(\hat w^{-1} -\bar a_n ) =
2146: \hat w^{-1} r_n (\hat w^{-1} -\bar a_n ) +
2147: \bar u^{(0)}_n (\hat w^{-1} -\bar a_n ) + \bar u_n \, \la{7111},
2148: \end{eqnarray}
2149: where $\hat w$ is the shift operator $n\to n+1$.
2150:
2151: Now we follow the procedure of the previous section. Since the
2152: potential has only
2153: one pole, ${\mathcal L}_n$ can be cast into
2154: $2\times 2$ matrix form. Let us
2155: apply the lines (\ref{701}, \ref{7111})
2156: to an eigenvector $(c_n, c_{n+1})$ of a yet unknown operator
2157: ${\mathcal L}_n$, and set
2158: the eigenvalue to be $\tilde z$:
2159: \be\la{94}\left
2160: \{
2161: \begin{array}{ccc}
2162: (z+r_{n-1} a_{n-1}-u^{(0)}_{n})c_{n} & = & r_{n} c_{n+1} +
2163: a_{n-1}(z+\bar \gamma_{n-1})c_{n-1} \\
2164: (\tilde z + r_{n}\bar a_{n} -
2165: \bar u^{(0)}_{n+1} )c_{n} & = &
2166: \bar a_{n+1}(\tilde z + \gamma_{n+1}) c_{n+1} + r_{n} c_{n-1}.
2167: \end{array}\right.
2168: \ee
2169: We have defined $\bar \gamma_n = \frac{u_n}{a_n} - u^{0}_n$.
2170: The equations are compatible if $c_{n-1}$ and $c_{n+1}$ found
2171: through $c_n$
2172: differ by the shift $n\to n+2$. We have
2173: \be \nonumber
2174: c_{n+1} = \frac{c_n}{d_n} \det \left | \begin{array}{cc}
2175: z+r_{n-1} a_{n-1} -u^{(0)}_{n}& a_{n-1}(z+\bar \gamma_{n-1}) \\
2176: \tilde z + r_n\bar a_{n} - \bar u^{(0)}_{n+1} & r_{n}
2177: \end{array} \right | = c_n \frac{\widetilde{{\mathcal D}}_{n}}{d_n},
2178: \ee
2179: \be \nonumber
2180: c_{n-1} = \frac{c_n}{d_n} \det \left | \begin{array}{cc}
2181: r_{n} & z + r_{n-1} a_{n-1} -u^{(0)}_{n} \\
2182: \bar a_{n+1}( \tilde z +
2183: \gamma_{n+1}) & \tilde z + r_n\bar a_{n} - \bar u^{(0)}_{n+1}
2184: \end{array} \right | = c_n \frac{{\mathcal D}_n}{d_n},
2185: \ee
2186: where
2187: \be \nonumber
2188: d_n =
2189: \det \left | \begin{array}{cc}
2190: r_{n} & a_{n-1}(z+\bar \gamma_{n-1}) \\
2191: \bar a_{n+1} (\bar z +\gamma_{n+1}) & r_{n}.
2192: \end{array} \right |
2193: \ee
2194: This yields the curve
2195: \be\la{811}
2196: \widetilde{\mathcal D}_n\cdot {\mathcal D}_{n+1}=d_nd_{n+1}.
2197: \ee
2198: Comparing the two forms of the curve
2199: (\ref{8}) and (\ref{811}),
2200: we obtain the conservation laws of growth:
2201: \be \nonumber
2202: \gamma =\gamma_n = \frac{\bar u_n}{\bar a_n}-\bar u^{0}_n,
2203: \ee
2204: \be\nonumber
2205: \beta = \frac{r_n}{\bar a_{n+1}} + u^{(0)}_{n+1} + \frac{u^{(0)}_{n+1}
2206: a_n \bar a_{n+1}}{1-a_{n}\bar a_{n+1}},
2207: \ee
2208: \be \nonumber
2209: n \hbar -\bar \alpha= r_n r_{n+1} - \frac{r_{n+1}u_{n+1}}{a_{n}a_{n+1}} .
2210: \ee
2211: They are the quantum version of (\ref{27}).
2212:
2213:
2214: The poles of the vector potential field
2215: determine most elements of the matrix
2216: ${\mathcal L}_n$. We have
2217: \be \nonumber
2218: {\mathcal L}_n
2219: \underline\chi_n=\Big({\bf A}_n+\frac{{\bf B}_n}{z}+
2220: \frac{{\bf C}_n}{z-\beta}\Big)
2221: \underline\chi_n,
2222: \quad\quad \underline\chi_n=(\chi_n,\chi_{n+1})^{t},
2223: \ee
2224: where $\bf {A, B}$ and ${\bf C}$ are $z$-independent $2\times 2$
2225: matrices. Their
2226: elements are
2227: $A_{11}=\bar\beta,\,A_{21}=\bar\beta a_n,\quad A_{12}=A_{22}=0,\quad
2228: B_{11}=t-\bar\alpha,\,B_{12}=
2229: \frac{r_{n-1}r_n}{a_{n-1}},\quad B_{21}=B_{22}=0,
2230: \quad\quad \mbox{tr}{\bf C }=-\alpha,\,\mbox{det}{\bf C}=0,\, \hbar n=
2231: -C_{11}-\bar\beta a_n(B_{12}+C_{12})$.
2232: The undetermined matrix elements are
2233: found similarly to the case of the ellipse.
2234: We treat the recurrence relation (the first line of (\ref{94}))
2235: as a $2\times 2$ representation of the shift operator
2236: $$
2237: {\mathcal W}_n=
2238: \left (
2239: \begin{array}{cc}
2240: 0 & 1 \\
2241: -z\frac{a_n}{r_{n+1}} &
2242: \frac{z+r_na_n-u^{(0)}_{n+1})}{r_{n+1}}
2243: \end{array}
2244: \right)
2245: $$
2246: and compute ${\mathcal L}_n$ from the second line of (\ref{94}). The
2247: matrix elements
2248: of ${\bf C}_n$ appear to be rather cumbersome.
2249:
2250:
2251:
2252:
2253: \subsection{Genus three -- hypotrochoid}
2254: The potential is $V(z) = t_3 z^3$. A general
2255: polynomial potential,
2256: including this example,
2257: has been analyzed in Ref. \cite{KM03}. We briefly review
2258: it here to illustrate features of the curve with multiple poles
2259: of the Schwarz function.
2260: In this case, $A(z)$ has a double pole at infinity. The number of
2261: sheets is three.
2262: The behavior of the
2263: Schwarz function at infinity is $S(z)\sim 3t_3z^2$ and the
2264: symmetry of the potential under $2\pi/3$-rotation restricts the
2265: equation of the curve to the form
2266: \be\la{321}
2267: f_n(z,\bar z)=(z \bar z)^2 - \frac{\bar z^3}{3t_3} - \frac{z^3}{3
2268: \bar t_3} + k_nz \bar z + h_n
2269: =0,
2270: \ee
2271: with two unknown coefficients. Since $S(z)$ has a double pole at infinity
2272: on the physical sheet,
2273: it has a branching point at infinity on the other sheets.
2274: This is the point where two unphysical sheets meet. Eq.(\ref{321})
2275: suggests that
2276: there are 9 more branching points.
2277: Therefore, genus of the curve is $g=3$. There are four droplets,
2278: but only one belongs to the physical sheet. $2\pi/3$ rotations leave
2279: the physical
2280: droplet invariant and exchange unphysical droplets. Therefore, the
2281: filling factors of
2282: unphysical droplets must be equal. This number, together with the
2283: filling factor of
2284: the physical droplet, determine the coefficients in (\ref{321}).
2285:
2286: Let us start from the semiclassical analysis. Assume that the
2287: filling factor of the
2288: physical droplet is $n$. We then have an algebraic domain
2289: bounded by a hypotrochoid, Fig.~\ref{hypotrochoid}.
2290: Six critical points sitting outside the
2291: droplet collapse two by two to three double points.
2292: The function
2293: \begin{equation} \label{hypotrochoid1}
2294: z(w) = rw + \frac{u}{w^2}
2295: \end{equation}
2296: maps the exterior of the unit disk to the exterior of the
2297: hypotrochoid. The inverse function
2298: has three branches $w_i(z),\quad i=1,2,3$.
2299: At infinity, the three sheets have the leading terms $w_1(z)\to z/r$,
2300: $w_{2,3}(z) \to \pm (u/z)^{1/2}$. Therefore, the inverse
2301: of the map
2302: (\ref{hypotrochoid1}) is $w_1(z)$.
2303: The branch $S(z) = \bar z(w^{-1}_1(z))$ is the Schwarz function.
2304: It is a meromorphic function outside the domain with one double
2305: pole at infinity,
2306: $S(z)\sim 3t_3z^2+ (n\hbar)/z$.
2307: This gives $\bar u=3t_3r^2$, while the area
2308: is $n\hbar=r^2-2|u|^2=
2309: r^2-18|t_3|^2 r^4$. The coefficients of the equation of the curve
2310: are determined by
2311: \be\la{341}
2312: k = \frac{(1-9|t_3|^2 r^2)(1+18|t_3|^2 r^2)}{9|t_3|^2},
2313: \quad
2314: h = - \frac{r^2(1-9|t_3|^2r^2)^3}{9|t_3|^2}.
2315: \ee
2316: Four remaining branching points are $z=\infty$ and
2317: three points solving $d z =
2318: (r-2uw^{-3})d w =0$, at $w_{*}^3 = 2u/r$. They are situated inside the
2319: droplet.
2320: Three double points solve
2321: $S^{(i)}(z) = S^{(j)}(z)$.
2322: They are outside the droplet:
2323: $$
2324: z_*^3 = \frac{(r^2-|u|^2)^3}{|u|^2\bar u}.
2325: $$
2326: At a critical degeneration the double points merge on the boundary to
2327: the branching
2328: point inside the droplet, $z_* = z(w_{*}), |w_{*}|=1$
2329: (Fig.~\ref{evolutionhyp}).
2330:
2331:
2332: Quantum analysis starts from recurrence relations guided by the
2333: classical case.
2334: Eqs. (\ref{M11}, \ref{M}) read
2335: \be\la{591}
2336: z\psi_n=r_n\psi_{n+1}+u_n\psi_{n-2},\quad\quad
2337: (L^{\dag}\psi)_n=r_{n-1}\psi_{n-1}+\bar u_{n+2}\psi_{n+2}.
2338: \ee
2339: Acting on the eigenvector of ${\mathcal L}_n$, we get
2340: \begin{equation*}
2341: \left \{
2342: \begin{array}{ccc}
2343: zc_n & = & r_n c_{n+1} + u_nc_{n-2} \\
2344: \tilde z c_{n} & = & r_{n-1}c_{n-1} + \bar u_{n+2}c_{n+2}.
2345: \end{array}
2346: \right .
2347: \end{equation*}
2348: Writing the first equation for $n\to n\pm 1$, we express $c_{n+2}$
2349: and $c_{n-1}$
2350: through $ c_n$ and $c_{n+1}$ to obtain
2351: \be \nonumber
2352: \tilde z c_n =z \frac{ \bar
2353: u_{n+2}}{r_{n+1}} c_{n+1} +
2354: \frac{1}{r_{n+1}}( r_{n-1}r_{n+1} -
2355: u_{n+1}\bar u_{n+2}) c_{n-1},
2356: \ee
2357: \be \nonumber
2358: zc_{n}= \left (r_n - \frac{u_n \bar u_{n+1}}{r_{n-2}} \right ) c_{n+1} +
2359: \bar z\frac{ u_n}{r_{n-2}} c_{n-1}.
2360: \ee
2361: Equations are compatible if $\tilde z$ belongs to the curve
2362: \be \nonumber
2363: \det
2364: \left |
2365: \begin{array}{cc}
2366: \left (r_{n+1} - \frac{u_{n+1} \bar u_{n+2}}{r_{n-1}} \right ) & z \\
2367: \frac{z \bar u_{n+3}}{r_{n+2}} & \tilde z
2368: \end{array}
2369: \right |
2370: \cdot
2371: \det
2372: \left |
2373: \begin{array}{cc}
2374: z & \frac{\tilde z u_n}{r_{n-2}} \\
2375: \tilde z & \left (r_{n-1} -
2376: \frac{u_{n+1}\bar u_{n+2}}{r_{n+1}} \right )
2377: \end{array}
2378: \right |= d_{n} d_{n+1},
2379: \ee
2380: where
2381: \be \nonumber
2382: d_n = \det
2383: \left |
2384: \begin{array}{cc}
2385: \left (r_n - \frac{u_n \bar u_{n+1}}{r_{n-2}} \right )
2386: & \frac{\tilde z u_n}{r_{n-2}} \\
2387: \frac{z \bar u_{n+2}}{r_{n+1}} &
2388: \left (r_{n-1} - \frac{u_{n+1}\bar u_{n+2}}{r_{n+1}}
2389: \right )
2390: \end{array}
2391: \right |.
2392: \ee
2393: Comparing with (\ref{321}), we find the quantum version of
2394: (\ref{341}):
2395: \be \nonumber
2396: k_n = r^2_n \left [ 1-9|t_3|^2 (r^2_{n-1}+r^2_{n+1} ) +
2397: \frac{1}{9|t_3|^2r^2_n} \right ],
2398: \ee
2399: \be \nonumber
2400: h_n = -
2401: \frac{r^2_n}{9|t_3|^2}(1-9|t_3|^2r^2_{n-1})
2402: (1-9|t_3|^2r^2_{n})(1-9|t_3|^2r^2_{n+1}),
2403: \ee
2404: where we used
2405: the conservation law and a quantum analog of the area
2406: formula
2407: \be \nonumber
2408: 3t_3 = \frac{\bar u_{n+1}}{r_{n-1}r_{n+1}},
2409: \quad
2410: r_n^2-(|u_{n+2}|^2+
2411: |u_{n+1}|^2)=\hbar n.
2412: \ee
2413: Together, these equations
2414: give a closed equation for $r_{n}^{2}$,
2415: $$
2416: r_n^2(1-|3t_3|^2(r_{n+1}^2+r_{n-1}^2))=\hbar n,
2417: $$
2418: which is a discrete analog
2419: of the Painlev\'e I equation.
2420:
2421:
2422:
2423:
2424: The matrix form of $L^{\dag}$ can be obtained in a similar manner.
2425: It is a $3\times 3$ matrix acting on the 3-vector
2426: $(\chi_{n},\chi_{n+1},\chi_{n+2})^{{\rm t}}$.
2427: Its analytic structure is
2428: $${\mathcal L}_{n}={\bf A}_n+z{\bf B}_n+z^2 {\bf C}_n,$$
2429: and the matrices ${\bf A},\,{\bf B},\,{\bf C}$ are
2430: $$\displaystyle
2431: {\bf A}_n=\begin{matrix}
2432: \left(
2433: \begin{array}{ccc}
2434: 0 &
2435: 0 & \frac{r_n}{r_{n-1}}(3\bar
2436: t_3)^{-1}(1-|3t_3|^2r_{n-1}^2) \\
2437: r_{n-1}(1-|3t_3|^2r_n^2)& 0 & 0
2438: \\
2439: 0 & r_{n}(1-|3t_3|^2r_{n+1}^2) & 0
2440: \end{array}
2441: \right )
2442: \end{matrix},
2443: $$
2444: $$\displaystyle
2445: {\bf B}_n=\begin{matrix}
2446: \left(
2447: \begin{array}{ccc}
2448: 0 &
2449: \frac{1}{ r_{n-1}}(3\bar t_3)^{-1} & 0 \\
2450: 0 & 0 & 3t_3r_n
2451: \\
2452: -|3t_3|^2r_nr_{n-1} &0 & 0
2453: \end{array}
2454: \right ) \end{matrix}
2455: ,\quad\quad\displaystyle
2456: {\bf C}_n=\begin{matrix}
2457: \left(
2458: \begin{array}{ccc}
2459: 0 &
2460: 0 & 0\\
2461: 0& 0 & 0
2462: \\
2463: 0 & 0 & 3t_3
2464: \end{array}
2465: \right ). \end{matrix}
2466: $$
2467:
2468: \section{Semiclassical wave function}\la{Semi2}
2469: In the semiclassical limit, $\hbar\to 0$, while
2470: $t=\hbar N$, $t^{(\alpha )}=\hbar\nu_\alpha$, and
2471: the potential $V(z)$ are kept fixed.
2472: The quantum curve and evolution equations go to their
2473: semiclassical counterparts.
2474: Also, the semiclassical
2475: wave functions are written through the objects of the
2476: semiclassical theory, i.e., through differentials on
2477: the semiclassical curve.
2478:
2479: A technical difficulty is that wave
2480: functions with close indices, say $\psi_n$ and
2481: $\psi_{n+1}$, are not necessarily close to each other
2482: as $n \to \infty$.
2483: Choosing a sequence of states
2484: which has a semiclassical limit
2485: is sensitive to the potential, and reflects the
2486: configuration of semiclassical droplets.
2487: This problem does not arise in the
2488: case of a single droplet in the algebraic case, where
2489: states with close numbers are close. As we discussed above, this case
2490: corresponds
2491: to degenerate Riemann surfaces.
2492: Hereafter,
2493: we assume that states with close $n$'s are close, and proceed to the
2494: semiclassical limit.
2495: The result obtained for this case can be generalized to
2496: any smooth Riemann surface.
2497:
2498:
2499: First, we rewrite the $L$ operator (\ref{M11}) in
2500: a more symmetric form
2501: $L=\hat w^{-1/2} r \hat w^{1/2}+\sum \hat w^{-k/2} u^{(k)} \hat w^{-k/2}$,
2502: by redefining $r$ and $u^{(k)}$. Then we search for a semiclassical expansion
2503: \be \nonumber
2504: \psi(z)
2505: \sim e^{\frac{1}{\hbar}{\mathcal A}_0 (z)+{\mathcal A}_1 (z) +\dots}
2506: \ee
2507: (in what follows, we suppress the index $n$).
2508:
2509:
2510: In the first two leading orders, the $L$-operator
2511: (\ref{M11}) reads
2512: $$L \longrightarrow z(w)+\frac{\hbar}{2}\{w
2513: \p_wz(w),\p_t\}{\mathcal A}_1+\dots,$$
2514: where $\log w=\p_t{\mathcal A}_0$ and $\{\,,\,\}$ means anticommutator.
2515: The function
2516: \be\la{681}
2517: z(w)=r w+\sum_{k\geq 0}u^{(k)} w^{-k}
2518: \ee
2519: is obtained by replacing the shift
2520: operator by its classical value $\hat w\to w$.
2521:
2522: The recurrence relation
2523: $L \psi(z)=z \psi(z)$ in two leading orders reads
2524: \be\nonumber
2525: z(w)=z,\quad
2526: \p_t{\mathcal A}_1=-\frac{1}{2} \frac{\p_t \p_w z(w)}{\p_w z(w)},\quad
2527: \log w=\p_t{\mathcal A}_0.
2528: \ee
2529: It gives
2530: \be\la{70}
2531: {\mathcal A}_1(z)=\frac{1}{2}\log w'(z),
2532: \ee
2533: where $w(z)$ is a multivalued function inverse to (\ref{681}).
2534:
2535:
2536: In the leading orders, (\ref{M}) reads
2537: \be\la{711}
2538: \hbar\p_z\psi(z)=\Big(-\frac{\bar z}{2}+\bar
2539: z(w^{-1})+\frac{\hbar}{2}\{w\p_w\bar z(w^{-1}),\p_t\}+{\mathcal
2540: A}_1+\dots\Big)\psi(z).
2541: \ee
2542: The function $\bar z(w^{-1})$ is the Schwarz reflection of $z(w)$
2543: with respect to the
2544: contour whose exterior is mapped to the exterior of the unit disk
2545: by the function $w(z)$. This map is discussed in Sec.~\ref{C}.
2546: The contour is the boundary of a semiclassical droplet. We observe
2547: that in the simplest semiclassical limit
2548: the conformal map and its Schwarz reflection are
2549: limits of the $L$ and $L^{\dag}$
2550: operators of Sec.~\ref{1A}.
2551: The complex curve
2552: in the semiclassical limit
2553: is given by $f(z,\bar z)=a(z)\prod (\bar z-\bar z(w^{-1}(z))$, where
2554: the product is taken over
2555: all possible branches of the multivalued function $w(z)$.
2556:
2557:
2558:
2559: Introducing the Schwarz function $S(z)=\bar z(w^{-1}(z))$,
2560: as in (\ref{ss}), we obtain
2561: \be\nonumber
2562: {\mathcal A}_0=-\frac{|z|^2}{2}+\Omega(z),
2563: \ee
2564: where $\Omega(z)=\int^z_{\xi_0} d\Omega$ is
2565: the primitive function of the
2566: generating differential
2567: introduced in Sec.~\ref{Diff}.
2568: The point $\xi_0$ is such that $\log w(\xi_0)=0$. It is
2569: a boundary point. The integral goes along the
2570: physical sheet.
2571:
2572: Summing up, we obtain the result:
2573: \beq\label{sp1}
2574: \psi (z)\simeq (2\pi^3 \hbar )^{-1/4}\sqrt{w'(z)}\,
2575: e^{\frac{1}{\hbar}(-\frac{1}{2}|z|^2 +\Omega (z))}.
2576: \eeq
2577:
2578: Formula (\ref{sp1}) was reported in \cite{ABWZ02}. A similar
2579: formula for the two-matrix model
2580: was obtained in \cite{Eynard97}. Being
2581: applied to the Hermite or Laguerre
2582: polynomials
2583: (see example in Sec.~\ref{classicalellipse}), the formula gives
2584: the asymptotical behaviour found by Tricomi in 1941 \cite{BE}.
2585:
2586: The overall numerical factor can be fixed either by comparison with
2587: asymptotics of biorthogonal
2588: polynomials for the Ginibre ensemble, or
2589: through normalization of the semiclassical wave function.
2590: The latter is described below.
2591:
2592: An extension of this formula
2593: to the case of a general complex curve and
2594: multiple droplets is
2595: straightforward.
2596: The function $\Omega (z)$ should be again
2597: understood as an integral of the generating
2598: differential $d\Omega$. The integration path
2599: starts at a point on the boundary
2600: of some ``reference" droplet and ends at the point $z$.
2601: The correction factor, $w'(z)$ in
2602: (\ref{sp1}), is replaced by the
2603: meromorphic Abelian differential of the third kind
2604: $W^{(\infty,\bar\infty )}(z)$, introduced
2605: in Sec.~\ref{Diff}. As a
2606: result, we obtain:
2607: \be\la{75}
2608: \psi (z)\sqrt{dz}\sim \sqrt{W^{(\infty,\bar\infty )}(z)}\,
2609: e^{-\frac{1}{\hbar}\big(\frac{|z|^2}{2}-\int^z_{\xi_0} S(z) dz\big)}.
2610: \ee
2611: By virtue of (\ref{18}), monodromies of this
2612: wave function around droplets are $e^{2\pi i\nu_{\alpha}}$,
2613: where $\nu_{\alpha}$ are filling factors.
2614:
2615: The asymptotical representation (\ref{75}) fails
2616: around cuts or double points. It is valid only around boundaries of
2617: physical droplets and
2618: only if the distance between boundary of a physical droplet
2619: and a cut or a double point
2620: is much larger than $\sqrt\hbar$. In other words, it
2621: is valid for smooth non-critical curves.
2622: The semiclassical asymptotics of the wave functon in
2623: the whole plane is far beyond the scope of this paper.
2624:
2625:
2626:
2627:
2628: The following properties of ${\mathcal A}_0$ visualize the result.
2629: \begin{itemize}
2630: \item [(i)]The real part of ${\mathcal A}_0$
2631: is constant along boundary of any
2632: droplet (see Sec.~\ref{Diff}), generally different for different droplets
2633: \be\nonumber
2634: 2{\mathcal R}e {\mathcal A}_0(z)=\phi_\alpha,
2635: \quad z\in\p D_\alpha.
2636: \ee
2637:
2638: \item[(ii)]Derivatives $\p_z$ and $\p_{\bar z}$
2639: of ${\mathcal R}e\, {\mathcal A}_0$ vanish on the boundary:
2640: \be \nonumber
2641: \p_z {\mathcal R}e {\mathcal A}_0(z)=
2642: \p_{\bar z} {\mathcal R}e {\mathcal A}_0(z)=
2643: 0, \quad z\in\p D.
2644: \ee
2645: This is the
2646: condition for semiclassical extrema (\ref{s}).
2647: \item[(iii)]
2648: In the leading order, the change of
2649: ${\mathcal R}e \, {\mathcal A}_0$ under
2650: small deviations along the normal direction away
2651: from the boundary to both exterior
2652: and interior does not depend on the point of the boundary:
2653: \be \nonumber
2654: \p_n^2 {\mathcal R}e\, {\mathcal A}_0(z)=-2.
2655: \ee
2656: \item [(iv)] The higher order expansion
2657: away from the boundary depends of the curvature of the boundary $k(z),
2658: z\in\p D_\alpha$:
2659: \beq\label{sp61}
2660: {\mathcal R}e{\mathcal A}_0(z +\delta n )=\phi_\alpha - (\delta n)^2 +
2661: \frac{1}{3}k(z)(\delta n)^3
2662: -\frac{1}{4}k^2 (z) (\delta n)^4 +\ldots,
2663: \eeq
2664: where $\delta n $ is a normal deviation from the boundary directed outward.
2665: This expansion fails
2666: at very curved parts of the contour, where $k(z)\sim \hbar^{-1/2}$.
2667: \end{itemize}
2668: Altogether, these properties mean
2669: that the wave function peaks at the
2670: boundary of a physical droplet and
2671: decays as Gaussian
2672: at the rate $\sqrt\hbar$ (see Fig.~\ref{semi}):
2673: \beq\label{sp6}
2674: |\psi (z\!+\!\delta n)|^2
2675: \sim
2676: e^{\frac{\phi_\alpha}{\hbar}}e^{-\frac{2}{\hbar}(\delta_n )^2}\,,
2677: \;\;\;\;z\in \p D_\alpha.
2678: \eeq
2679: The amplitude of the wave function may be different on different
2680: boundaries.
2681:
2682:
2683: Since the
2684: wave function is localized on the boundary, one
2685: can integrate in the normal direction. In the case
2686: of one droplet, we have
2687: \be\la{78}
2688: \int |\psi (z)|^2 d(\delta n)=
2689: \frac{1}{2\pi}
2690: \p_n\log |w(z)|=\frac{1}{2\pi}|w'(z)|,\quad z\in\p D
2691: \ee
2692: (here $\p_n$ is the normal derivative).
2693: Further integration along the boundary is
2694: consistent with the normalization
2695: $\int |\psi_N (z)|^2 d^2z =1$.
2696: This can be used to determine the constant in (\ref{sp1}).
2697:
2698: According to the arguments of Sec.~\ref{F1}, Eq. (\ref{78})
2699: can be interpreted as a rate of growth.
2700:
2701: Note that Eq. (\ref{70}) and the next to leading order
2702: of (\ref{711}) give the relation
2703: \be\la{zz} w(\p_w z\p_t\bar z-\p_tz\p_w\bar z)=2,
2704: \ee
2705: where one differentiates at a fixed $w$. This relation
2706: represents evolution of the growing domain
2707: as a time-dependent conformal map.
2708: The area of the domain grows linearly with time, while
2709: its harmonic moments stay fixed. This type of classical
2710: growth is equivalent
2711: to the Laplacian growth.
2712: Relation (\ref{zz}) has been known in the
2713: literature for a long time \cite{Kochina}.
2714: It is a classical limit of Eq. (\ref{string}).
2715:
2716:
2717:
2718:
2719:
2720: \section{Appendix. Laplacian growth}\label{Appendix_A}
2721: Laplacian growth referres to growth of a planar domain, whose boundary
2722: propagates with a velocity
2723: proportional to the gradient of a harmonic field. The
2724: Hele-Shaw problem is a typical example (see \cite{review} for a
2725: review). The experimental set-up consists (Fig.~\ref{hs1})
2726: of two parallel and horizontal glass plates, separated by a small distance along the
2727: vertical direction, called $the$ $gap$, $b$. The gap is small enough compared
2728: to typical intermolecular distances in fluids, so that the system can be considered
2729: two-dimensional.
2730:
2731:
2732: \begin{figure} \begin{center}
2733: %\centering
2734: \includegraphics*[width=5cm]{hs1.eps}
2735: \caption{\label{hs1}The Hele-Shaw setup.}
2736: \includegraphics*[width=5cm]{hs3.eps}
2737: \caption{\label{hs3}Laminar Hele-Shaw dynamics.}
2738: \end{center} \end{figure}
2739:
2740:
2741: Initially, the space between the plates is filled with fluid 2, of large viscosity
2742: (tipical examples are silicon oil or liquid crystals). Through a small opening in
2743: the middle of the upper glass plate, fluid 1, of much lower viscosity (such as air),
2744: is inserted into the system, at a constant rate. Both fluids are incompressible and
2745: immiscible, so fluid 2 must be evacuated from the system at the same rate as fluid
2746: 1 is being introduced. This is done somewhere at the edge of the glass plates.
2747:
2748: Given that the fluids are incompressible and immiscible, the dynamics of the system
2749: is reduced to that of the interface between them. Depending on the relevant
2750: parameters of the experiment (the coefficients of viscosity, $\mu_{1,2}$ of the
2751: fluids, the coefficient of surface tension $\sigma$, the gap $b$ and the pumping rate
2752: $Q$), the interface dynamics will exhibit very different behaviors. At low pumping
2753: rates $Q$, for given $\sigma$, the interface will remain smooth at all times, regardless
2754: of how much fluid is being inserted (Fig.~\ref{hs3}). However, at high $Q$ (or
2755: equivalently, at very low $\sigma$), the boundary will develop the so-called $fingering$
2756: $instability$, where certain regions grow in the shape of fingers, whose tips split and
2757: form secondary fingers, etc. Real experiments (Fig.~\ref{sw})
2758: show that in this regime, the interface seems to evolve into a self-similar, fractal structure.
2759:
2760: Under the assumption that the system is two dimensional, the equations
2761: governing the dynamics of the interface lead to Darcy's Law:
2762: the velocity field $\vec{v}_{i}$ and pressure $p_i$ in fluid $i$ are related as:
2763: \be
2764: \vec{v}_i = -\frac{b^2}{12\mu_i}\vec{\nabla}p_i,
2765: \ee
2766: and therefore the continuity equation accross the interface leads to
2767: \be
2768: \vec{v}\cdot \vec{n} = -\frac{b^2}{12\mu_1}\vec{n}\cdot\vec{\nabla}p_1
2769: = -\frac{b^2}{12\mu_2}\vec{n}\cdot\vec{\nabla}p_2,
2770: \ee
2771: where $\vec{n}$ is the unit vector normal to the interface at a given point.
2772:
2773: In each fluid, the pressure field is supposed to solve the Laplace equation,
2774: \be
2775: \Delta p_i = 0,
2776: \ee
2777: with appropriate boundary conditions. For instance, since fluid 2 is drained from the
2778: system somewhere away from the origin, fluid 2 has a sink at infinity, which leads to
2779: the condition
2780: \be
2781: p_2(z) \to -\log |z|, \quad z \to \infty.
2782: \ee
2783: The coefficient of surface tension enters the problem through the Laplace formula
2784: relating the pressure jump accross the boundary and the local curvature $\kappa(z)$:
2785: \be
2786: p_2(z) -p_1(z) = \sigma \kappa(z).
2787: \ee
2788:
2789: The first simplification of the problem consists in setting the viscosity of fluid 1 to zero,
2790: $\mu_1=0$. The continuity condition then implies $p_1 = $ constant, and we may therefore
2791: choose $p_1 = 0$ (a redefinition of $p_2$).
2792:
2793: The second important simplification is neglecting the effect of surface tension. This assumption
2794: has important physical meaning, since the resulting problem should necessarily lead to turbulent
2795: dynamics. Mathematically, the two assumptions lead to the following reformulation of the
2796: problem:
2797: \be
2798: \Delta p = 0 \mbox{ on } D_{-}, \quad p = 0, \, \, V_n \sim \p_n p \, \, \, \mbox{ on } \Gamma,
2799: \quad p(z) \stackrel{z \to \infty}{\longrightarrow} -\log|z| ,
2800: \ee
2801: where $V_n$ is the normal component of the velocity, $D_{-}$ is the domain occupied by fluid 2,
2802: and $\Gamma$ is the boundary between fluids. Under these assumptions, in order to describe the
2803: interface dynamics, we must solve an exterior Dirichlet problem for the pressure field.
2804:
2805: \section*{Acknowledgments}
2806:
2807: We are indebted to
2808: A. Kapaev,
2809: V. Ka\-za\-kov, I. Kri\-che\-ver, I. Kos\-tov,
2810: A. Mar\-sha\-kov and M. Mi\-ne\-ev-\-Wein\-stein
2811: for useful discussions, interest in the subject and help.
2812: P.W. and R.T. were supported by the NSF MRSEC Program under
2813: DMR-0213745, NSF DMR-0220198 and by the Humboldt foundation.
2814: A.Z. and P.W. acknowledge support
2815: by the LDRD project 20020006ER ``Unstable
2816: Fluid/Fluid Interfaces" at Los Alamos National Laboratory and M.
2817: Mi\-ne\-ev-\-Wein\-stein for the hospitality
2818: in Los Alamos. A.Z. was also supported in part by RFBR grant
2819: 03-02-17373 and by the grant for support of scientific schools
2820: NSh-1999.2003.2. P.W. is grateful to K.B. Efetov for the hospitality in
2821: Ruhr-Universitaet Bochum and to A.Cappelli for the hospitality in the
2822: University of Florence,
2823: where this work was completed. We are grateful to Harry Swinney
2824: for permitting us to use Fig.~\ref{sw} from \cite{Sw}.
2825:
2826:
2827: \begin{thebibliography}{10}
2828: \bibitem{mwz} Mineev-Weinstein, M., Wiegmann, P.B., \& Zabrodin, A., 2000,
2829: Phys.\ Rev.\ Lett.\ \textbf{84}, 5106.
2830:
2831: \bibitem{ABWZ02} Agam, O., Bettelheim, E., Wiegmann, P.B.,
2832: \& Zabrodin, A., Viscous Fingering and a Shape of an
2833: Electronic Droplet in a Quantum Hall Regime, 2002, Phys. Rev. Lett.
2834: {\bf {88}}, 236802.
2835:
2836: \bibitem{mkwz}Krichever, I., Mineev-Weinstein, M., Wiegmann, P.B., \&
2837: Zabrodin, A., Laplacian growth and Whitham Equations of Soliton Theory,
2838: 2003, nlin.SI/0311005.
2839:
2840: \bibitem{DV03} Dijgraaf, R., \& Vafa, C., ${\mathcal{N}}=1$
2841: Supersymmetry, Deconstruction and Bosonic Gauge Theory, 2003, hep-th/0302011,
2842: A Perturbative Window into Non-Perturbative Physics, 2002, hep-th/0208048,
2843: Matrix Models, Topological Strings and
2844: Supersymmetric Gauge Theories, 2002, hep-th/0206255,
2845: On Geometry and Matrix Models, 2002, hep-th/0207106.
2846:
2847: \bibitem{BEH}
2848: Bertola, M., Eynard, B. \& Harnad, J.,
2849: Partition functions for Matrix Models and Isomonodromic
2850: Tau Functions, 2002, nlin/0204054.
2851:
2852: \bibitem{Kapaev} Kapaev, A., Riemann-Hilbert problem
2853: for bi-orthogonal polynomials, 2003, J. Phys. A
2854: {\bf 36}, 4629-4640.
2855:
2856: \bibitem{gravity} Di Francesco, P., Ginsparg, P., \& Zinn-Justin, J.,
2857: 2-D Gravity and Random Matrices, 1995, Phys. Rept. 254, 1-133.
2858:
2859: \bibitem{Its}Fokas, A.S., Its, A.R., \& Kitaev, A., The isomonodromy
2860: approach to the matrix models in 2-D gravity, 1992, Comm. Math. Phys. 147, 395-430.
2861:
2862:
2863:
2864: \bibitem{KM03} Kazakov, V.A., \& Marshakov, A.,
2865: Complex Curve of the Two Matrix Model and its Tau-function, 2003,
2866: J. Phys. A {\bf {36}}, 3107-3136.
2867:
2868:
2869:
2870: \bibitem{Chekhov}Chekhov, L., \& Mironov, A., Matrix
2871: Models vs. Seiberg-Witten/Whitham theories, 2002,
2872: hep-th/0209085.
2873:
2874: \bibitem{David}David, F., Non-Perturbative Effects
2875: in Matrix Models and Vacua of Two Dimensional
2876: Gravity, 1993, Phys.Lett. B302, 403-410; hep-th/9212106; David, F., Bonnet, G., \&
2877: Eynard, B., Breakdown of universality in multi-cut matrix models, 2000,
2878: J.Phys. A33, 6739.
2879:
2880:
2881: \bibitem{Zaboronsky}Chau, Ling-Lie, \& Zaboronsky, O., On the structure
2882: of Normal Matrix Model, 1998, Commun.
2883: Math. Phys. 196, 203-247, hep-th/9711091.
2884: \bibitem{Mehta91} Mehta, M.L., Random
2885: Matrices, 1991, Academic Press, New York.
2886:
2887: \bibitem{WZ03} Wiegmann, P.B., \& Zabrodin, A., Large
2888: Scale Correlations in Normal and General Non-Hermitian Matrix
2889: Ensembles, 2003, J. Phys. A, {\bf {36}}, 3411-3424.
2890:
2891:
2892:
2893: \bibitem{Ginibre65} Ginibre, J., Statistical
2894: Ensembles of Complex Quaternion and Real Matrices, 1965, Journal of
2895: Math. Phys. {\bf {6}} (3), 440.
2896: Girko, V.L., Elliptic Law, 1986, Theory of Probability and
2897: Its Applications, 30, (4), 677-690.
2898:
2899: \bibitem{Sw}Sharon, E.,
2900: Moore, M.G., McCormick, W.D., \& Swinney, H.L.,
2901: Coarsening of Fractal Viscous Fingering Patterns, 2003,
2902: Phys. Rev. Lett., 91, 205504.
2903:
2904:
2905: \bibitem{mesoscopic}Altshuler, B.L., \& Simons, B.D.,
2906: Universalities: from Anderson Localization to Quantum Chaos,
2907: in: Mesoscopic Quantum Physics, 1994, Les Houches 1994, eds.
2908: Akkermans, E., Montambaux, G., Pichard, J-L., \& Zinn-Justin, J.
2909:
2910:
2911:
2912: \bibitem{Aratyn}Aratyn, H., Integrable Lax Hierarchies, their
2913: Symmetry Reductions
2914: and Multi-Matrix Models, 1995, hep-th/9503211.
2915:
2916:
2917: \bibitem{BEHdual}
2918: Bertola, M., Eynard B., \& Harnad, J.,
2919: Duality of spectral curves arising in two-matrix models, 2001,
2920: nlin/0112006.
2921:
2922: \bibitem{Eynard03} Bertola, M., Eynard, B., \& Harnad, J.,
2923: Differential systems for biorthogonal polynomials appearing in 2-matrix
2924: models and the associated
2925: Riemann-Hilbert problem, 2002, nlin/0208002.
2926:
2927: \bibitem{Krichev-red} Krichever, I.M., Funkt. Analiz i
2928: ego Pril., 1995, {\bf 29/2}, 1-8.
2929:
2930: \bibitem{Wasow}
2931: Wasow, W., Asymptotic Expansions for Ordinary
2932: Differential Equations, 1965, John Wiley \& Sons, New York.
2933:
2934: \bibitem{curve}Dubrovin, B.A., Krichever, I.M., \& Novikov, S.P.,
2935: Topological and
2936: algebraic geometry methods in contemporary mathematical physics II, 1982,
2937: Soviet Scient. Reviews, Math. Phys. Reviews 3 1-150;
2938: Krichever, I.M., Integration of nonlinear equations by the method
2939: of algebraic geometry, 1977, Funct. Anal. Appl., {\bf 11}, 12-26.
2940:
2941:
2942: \bibitem{Alhfors50} Ahlfors, L.V., Complex
2943: Analysis, an Introduction to the Theory of Analytic Functions of One
2944: Complex Variable, 1953, McGraw-Hill, New York.
2945:
2946:
2947: \bibitem{C}
2948: Cohn, H., Conformal Mapping on Riemann Surfaces, 1967, Dover, New York.
2949:
2950:
2951: \bibitem{SS} Schiffer, M., \& Spencer, D.C.,
2952: Functionals of finite Riemann surfaces, 1954,
2953: Princeton University Press.
2954:
2955:
2956: \bibitem{Miwa}Jimbo, M., Miwa, T., \& Ueno, K.,
2957: Monodromy preserving deformations of
2958: linear ordinary differential equations with rational coefficients I, 1981, Physica D, {\bf 2}, 306-52; Jimbo, M., \& Miwa, T.,
2959: Monodromy preserving deformation of linear
2960: ordinary differential equations with rational coefficients II, 1981, Physica D,
2961: {\bf 2}, 407-48; Flaschka, H., \& Newell, A.C.,
2962: Monodromy- and spectrum-preserving
2963: deformations, 1980, Commun. Math. Phys., {\bf 76}, 65-116.
2964:
2965: \bibitem{Takasaki} Takasaki, K.,
2966: Dual Isomonodromic Problems and Whitham Equations, 1998, Lett. Math. Phys.
2967: {\bf 43} 123-135.
2968:
2969:
2970: \bibitem{Ahar-Shap}
2971: Aharonov, D., \& Shapiro, H., 1976,
2972: J. Anal. Math., {\bf 30}, 39-73;\\
2973: Shapiro, H., The Schwarz function
2974: and its generalization to higher dimensions, 1992,
2975: University of Arkansas Lecture Notes in the
2976: Mathematical Sciences, Volume 9, Summers W.H.,
2977: Editor, John Wiley \& Sons.
2978:
2979: \bibitem{Kostov}Kostov, I.K., Conformal field theory techniques in
2980: random matrix models, 1999, arXiv:hep-th/9907060.
2981:
2982:
2983: \bibitem{review} Bensimon, D., Kadanoff, L.P.,
2984: Liang, S., Shraiman, B.I., \& Tang, C., 1986,
2985: Rev.\ Mod.\ Phys.\, \textbf{58}, 977.
2986:
2987: \bibitem{kkmwz}Kostov, I.K., Krichever, I., Mineev-Weinstein, M.,
2988: Wiegmann, P.B., \& Zabrodin, A., $\tau$-Function for Analytic
2989: Curves, in:
2990: Random matrices and their applications, 2000, MSRI publications, vol. 40, 285,
2991: Cambridge University Press.
2992:
2993: \bibitem{Eynard97} Eynard, B., Eigenvalue
2994: Distribution of Large Random Matrices, from One Matrix to Several
2995: Coupled Matrices, 1997, cond-mat/9707005.
2996:
2997:
2998: \bibitem{BE} Bateman, H., \& Erdelyi, A., Higher transcendal
2999: functions, 1953,
3000: v. 2, McGraw-Hill.
3001:
3002:
3003: \bibitem{Morozov}Marshakov, A., Mironov, A., \& Morozov, A.,
3004: Generalized matrix models as conformal
3005: field theories: Discrete case, 1991, Phys. Lett. B, {\bf 265}, 99; Kharchev, S.,
3006: Marshakov, A., Mironov, A., Morozov, A., \& Pakuliak, S.,
3007: Conformal matrix models
3008: as an alternative to conventional multimatrix models, 1993, Nucl. Phys. B
3009: {\bf 404}, 717, [arXiv:hep-th/9208044].
3010:
3011:
3012:
3013: \bibitem{Toda}
3014: Ueno, K., \& Takasaki, K., 1984,
3015: Adv. Stud. Pure Math., {\bf 4}, 1.
3016:
3017:
3018:
3019: \bibitem{Whitham}Whitham, J.B.,
3020: Linear and nonlinear waves, Wiley-Interscience, 1974,
3021: New York;
3022: Flaschka, H., Forest, M.G., \& McLaughlin, D.W.,
3023: Multiphase averaging and the inverse spectral solution of
3024: the Korteweg-de Vries equation, 1980,
3025: Comm. Pure Appl. Math., {\bf 33}, 739-84;
3026: Krichever, I.M.,
3027: Method of averaging for two-dimensional ``integrable"
3028: equations, 1988, Funct. Anal. Appl., {\bf 22}, 200-213.
3029:
3030: \bibitem{Kochina}
3031: Galin, L.A., 1945, Dokl. Akad. Nauk SSSR,
3032: {\bf 47}, 250-253;\\
3033: Polubarinova-Kochina, P.Ya., 1945, Dokl. Akad. Nauk SSSR,
3034: {\bf 47}, 254-257; \\
3035: Kufarev, P.P., 1947, Dokl. Akad. Nauk SSSR
3036: {\bf 57}, 335-348.
3037:
3038:
3039: \bibitem{Akemann02}Akemann, G., The Solution of a Chiral Random
3040: Matrix Model with Complex Eigenvalues, 2002, J. Phys. {\bf A36}, 3363.
3041:
3042:
3043: \bibitem{1}Wiegmann, P.B., \& Zabrodin, A.,
3044: Conformal maps and dispersionless integrable hierarchies, 2000,
3045: Commun.Math.Phys. {\bf 213}, 523-538;
3046: Marshakov, A., Wiegmann, P.B., \& Zabrodin, A.,
3047: Integrable Structure of the Dirichlet Boundary Problem in Two Dimensions, 2002,
3048: Commun. Math. Phys. {\bf 227}, 131-153.
3049: \bibitem{mkz} Krichever, I., Marshakov, A., \& Zabrodin, A.,
3050: Integrable Structure of the Dirichlet Boundary Problem in
3051: Multiply-Connected Domains, 2003, hep-th/0309010.
3052:
3053: \end{thebibliography}
3054:
3055:
3056:
3057: \end{document}
3058:
3059:
3060:
3061: \end{document}
3062:
3063: