1: \documentclass[10pt,twoside,a4paper]{amsart}
2: \usepackage{amsmath}
3: \usepackage{amsfonts}
4: \usepackage{amssymb}
5: \usepackage{amsthm}
6: \usepackage{newlfont}
7: \usepackage{graphicx}
8: \usepackage{amscd}
9:
10: \textwidth 6in
11: \textheight 8.5in
12: \topmargin -1cm
13: \leftmargin -3cm
14: \oddsidemargin=0.8cm
15: \evensidemargin=0.8cm
16: % Fuzz -------------------------------------------------------------------
17: \hfuzz5pt % Don't bother to report over-full boxes if over-edge is < 5pt
18: %\setlength{\tclineskip}{1.05\baselineskip}
19: %%% ----------------------------------------------------------------------
20: %\include{/home/doliwa/texfiles/mydef}
21: % THEOREMS ---------------------------------------------------------------
22: \theoremstyle{plain}
23: \newtheorem{Th}{Theorem}%[section]
24: \newtheorem{Cor}[Th]{Corollary}
25: \newtheorem{Lem}[Th]{Lemma}
26: \newtheorem{Prop}[Th]{Proposition}
27: %
28: \theoremstyle{definition}
29: \newtheorem{Def}{Definition}%[section]
30: \newtheorem{Ex}{Example}%[section]
31: %
32: \theoremstyle{remark}
33: \newtheorem*{Rem}{Remark}%[section]
34: %
35: \numberwithin{equation}{section}
36: %%% ----------------------------------------------------------------------
37: \newcommand{\PP}{{\mathbb P}}
38: \newcommand{\ZZ}{{\mathbb Z}}
39: \newcommand{\RR}{{\mathbb R}}
40: %%%-----------------------------------------------------------------------
41: \newcommand{\bx}{\boldsymbol{x}}
42: \newcommand{\bX}{\boldsymbol{X}}
43: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
44:
45: \begin{document}
46:
47: \title[B-quadrilateral lattice]
48: {The B-quadrilateral lattice, its
49: transformations \\
50: and the algebro-geometric construction}
51:
52: \author[Adam Doliwa]{Adam Doliwa$^\ddagger$}
53: \thanks{$\ddagger$ Supported in part by the DFG Research Center MATHEON}
54:
55: \address{Wydzia{\l} Matematyki i Informatyki,
56: Uniwersytet Warmi\'{n}sko-Mazurski w Olsztynie,
57: ul. \.{Z}o{\l}nierska 14, 10-561 Olsztyn, Poland}
58:
59: %\curraddr
60: %\address
61: %{DFG Research Center MATHEON,
62: %Institut f\"{u}r Mathematik, Technische Universit\"{a}t Berlin,
63: %Stra\ss{}e des 17. Juni 136, 10623 Berlin, Germany}
64:
65: \email{doliwa@matman.uwm.edu.pl}
66:
67: %
68: %\date{\today}
69: \keywords{integrable discrete geometry; discrete BKP equation; Prym varieties;
70: Darboux transformations}
71: \subjclass[2000]{37K10, 37K20, 37K25, 37K35, 37K60, 39A10}
72:
73: \begin{abstract}
74: The B-quadrilateral lattice (BQL) provides geometric interpretation of Miwa's
75: discrete BKP equation within the quadrialteral lattice (QL)
76: theory. After discussing the projective-geometric properties of the lattice we
77: give the algebro-geometric construction of the BQL
78: ephasizing the role of Prym varieties and the corresponding theta functions.
79: We also present the reduction of the vectorial fundamental transformation of the
80: QL to the BQL case.
81: \end{abstract}
82: %{\it 2001 PACS:} 02.30.Ik, 02.40.Dr, 05.45.Yv, 04.60.Nc, 02.40.Hw,
83: \maketitle
84:
85: \tableofcontents
86:
87: \section{Introduction}
88: During last years the many results describing the well known
89: connection between integrable partial
90: differential equations and the differential geometry of submanifolds have been
91: transfered to the discrete (difference) level,
92: see for example \cite{BobSur,DS-EMP,Schief-JNMP}. The interest in such a
93: research is stimulated from various fields, like computer visualization,
94: combinatorics, lattice models in statistical mechanics and quantum field theory,
95: and recent developments in quantum gravity.
96:
97:
98: A succesful general approach towards description of the relation between
99: integrability and geometry is provided by the theory of
100: multidimensional quadrilateral lattices (QLs) \cite{MQL}. These are just
101: maps $x:\ZZ^N\to\PP^M$ ($3\leq N\leq M$) with planar elementary quadrilaterals.
102: The integrable partial difference equation counterpart of the QLs are
103: the discrete Darboux equations (see Section \ref{sec:tau} for details), being found
104: first \cite{BoKo} as the most general
105: difference system integrable by the
106: $\bar\partial$ method. It should be metioned that the (differential)
107: Darboux equations \cite{Darboux-OS} play
108: an important role \cite{BoKo-N-KP} in the multicomponent
109: Kadomtsev--Petviashvilii (KP) hierarchy,
110: which is commonly considered \cite{DKJM,KvL}
111: as the fudamental system of equations in integrability theory.
112:
113: \begin{figure}[h!]
114: \begin{center}
115: \includegraphics{TiTjTkx.eps}
116: \end{center}
117: \caption{The geometric integrability scheme}
118: \label{fig:TiTjTkx}
119: \end{figure}
120: It turns out that integrability of the discrete Darboux system is encoded in a
121: very simple geometric statement (see Fig.~\ref{fig:TiTjTkx}).
122: \begin{Lem}[The geometric integrability scheme] \label{lem:gen-hex}
123: Consider points $x_0$, $x_1$, $x_2$ and $x_3$ in general position in $\PP^M$,
124: $M\geq 3$. On
125: the plane $\langle x_0, x_i, x_j \rangle$, $1\leq i < j \leq 3$ choose a point
126: $x_{ij}$ not on the lines $\langle x_0, x_i \rangle$, $\langle x_0,x_j
127: \rangle$ and $\langle x_i, x_j \rangle$. Then there exists the
128: unique point $x_{123}$
129: which belongs simultaneously to the three planes
130: $\langle x_3, x_{13}, x_{23} \rangle$,
131: $\langle x_2, x_{12}, x_{23} \rangle$ and
132: $\langle x_1, x_{12}, x_{13} \rangle$.
133: \end{Lem}
134: Integrable reductions of the quadrilateral lattice (and thus of the discrete
135: Darboux equations) arise from additional constraints which are compatible with
136: geometric integrability scheme (see, for example \cite{q-red,DS-sym}).
137: One of the most important reductions of the KP hierarchy of nonlinear equations
138: is the so called BKP hierarchy \cite{DJKM-BKP}
139: (here "B" appears in the context of the classification theory of simple
140: Lie algebras). In \cite{Miwa} it was shown that the $\tau$-function of the
141: BKP hiererchy satisfies certain bilinear discrete equation
142: (the transformation between the infinite
143: sequence of times of the hierarchy and the corresponding discrete variables
144: is called the Miwa transformation)
145: \begin{equation} \label{eq:Hirota-Miwa}
146: \tau\, \tau_{(123)} = \tau_{(12)}\tau_{(3)} - \tau_{(13)}\tau_{(2)} +
147: \tau_{(23)}\tau_{(1)},
148: \end{equation}
149: which is known as the discrete BKP or the
150: Miwa equation. Here and in all the paper, given a fuction $F$ on
151: $\ZZ^N$, we denote its shift in the $i$th direction in a standard manner:
152: $F_{(i)}(n_1,\dots, n_i, \dots , n_N) = F(n_1,\dots, n_i + 1, \dots , n_N)$.
153:
154: The linear problem and the Darboux-type (Moutard)
155: transformations for the discrete BKP equation \eqref{eq:Hirota-Miwa} were
156: constructed in \cite{NiSchief}. In literature there are known several geometric
157: interpretations of the discrete BKP equation in terms of the reciprocal
158: figures and inversive geometry~\cite{KoSchiefSBKP,Schief-JNMP},
159: or in terms of the trapezoidal nets \cite{BobSur}. It should be also
160: mentioned that
161: the discrete BKP equation has been recently investigated, under the name of the
162: \emph{cube recurrence}, in combinatorics
163: \cite{Propp,FominZelevinsky,CarrollSpeyer}.
164:
165: In this paper we propose (see Section~\ref{sec:BQL}) another geometric
166: interpretation of the discrete BKP
167: equation, which we consider from the point of view of the quadrilateral lattice
168: theory. This new reduction of the quadrilateral lattice, which we call the
169: \emph{B-quadrilateral lattice} (BQL), is projectively invariant and is based on
170: additional local linear constraint. Section~\ref{sec:BQL} is of rather
171: elementary geometric nature, but the results obtained there have far reaching
172: consequences. In fact, the paper gives new arguments supporting the
173: conjecture that
174: basic integrability features are consequences of incidence geometry statements
175: (see also introductory remarks in \cite{DS-EMP}).
176:
177: More involved techniques are used in Section~\ref{sec:BQL-Prym},
178: where we elaborate the
179: algebro-gemetric method to produce large classes of the B-quadrilateral lattices
180: and the corresponding solutions of the discrete BKP system of equations.
181: In doing that we start from recent results of
182: \cite{DGNS}, where restrictions on the algebro-geometric data of the
183: discrete Darboux system \cite{AKV}
184: compatible with the discussed reduction were given. Then we proceed to
185: formulas for the wave and $\tau$-functions of the BQL in terms
186: of the Prym theta functions related with the algebraic curves
187: used in the construction. We transfer this way the algebro-geometric method
188: of construction of solutions of the BKP hierarchy \cite{DJKM-Prym-BKP} and of its
189: two-component generalization \cite{VeselovNovikov} to the discrete level.
190:
191: We also present, in
192: Section~\ref{sec:B-transf}, the corresponding reduction
193: of the vectorial fundamental transformation of the quadrilateral lattice
194: \cite{TQL} and establish its link with the Pfaffian form of the vectorial
195: Moutard transformation found in \cite{NiSchief}.
196: In Appendix we give alternative proof of a crucial auxilliary
197: result of the
198: paper, and we summarize basic properties of Pfaffians.
199:
200: \section{The B-quadrilateral lattice}
201: \label{sec:BQL}
202: We start with discussing the gometric constraint, which imposed on the
203: quadrilateral lattice
204: allows to define its new integrable reduction. Then we proceed to the algebraic
205: description of such reduced lattice showing its connection with the discrete
206: BKP equation. Finally, we discuss relation between the
207: $\tau$-function of the
208: B-quadrilateral lattice, and the $\tau$-function of the quadrilateral lattice.
209: \subsection{Geometric definition of the BQL}
210: \begin{Prop} \label{lem:BKP-hex}
211: Under hypothesess of Lemma \ref{lem:gen-hex}, assume that the points
212: $x_0$, $x_{12}$, $x_{13}$, $x_{23}$ are coplanar, then the points
213: $x_1$, $x_2$, $x_3$, and $x_{123}$ are coplanar as well (see
214: Figure~\ref{fig:moutard}).
215: \end{Prop}
216: \begin{figure}
217: \begin{center}
218: \includegraphics{moutard}
219: \end{center}
220: \caption{Elementary hexahedron of the B-quadrilateral lattice}
221: \label{fig:moutard}
222: \end{figure}
223: It can be shown by the standard linear algebra (for a synthetic-geometry proof
224: see a remark below). We perform however the calculations, because the way we are
225: going to do it will be important in next Sections
226: in showing connection of the BQL with
227: the discrete BKP equation.
228: \begin{Lem} \label{lem:BQL-gauge-initial}
229: Under hypotheses of Proposition \ref{lem:BKP-hex}, for fixed initially
230: homogeneous coordinates $\bx_0$ and $\bx_1$ (gauges) of $x_0$ and $x_1$,
231: there exist a gauge
232: such that the following linear relations hold
233: \begin{equation} \label{eq:BQL-gauge-initial}
234: \bx_{ij} - \bx_0 = f^{ij} (\bx_{i} - \bx_{j}) , \quad 1\leq i< j\leq 3,
235: \end{equation}
236: where the coefficients $f^{ij}$ depend on the actual positions of the points
237: $x_{ij}$.
238: \end{Lem}
239: \begin{proof}
240: The coplanarity of the four points $x_0$, $x_1$, $x_2$ and $x_{12}$ can be
241: algebraically expressed as the
242: linear relation
243: \begin{equation*}
244: \alpha\bx_{0} + \beta\bx_{1} + \gamma\bx_{2} + \delta\bx_{12} = 0,
245: \end{equation*}
246: where, by the genericity assumption (no three of the points are collinear),
247: all the
248: coefficients do not vanish. By plaing with rescalling the homogeneous
249: coordinates of $x_2$ and $x_{12}$ we can transfer above equation
250: to the form \eqref{eq:BQL-gauge-initial}
251: \begin{equation} \label{eq:BQL-gauge-initial-12}
252: \bx_{12} - \bx_0 = f^{12} (\bx_{1} - \bx_{2}) .
253: \end{equation}
254: Similarly, we can rescale the homogeneous
255: coordinates of $x_3$ and $x_{13}$ to express planarity of the corresponding
256: elementary quadrilateral as
257: \begin{equation} \label{eq:BQL-gauge-initial-13}
258: \bx_{13} - \bx_0 = f^{13} (\bx_{1} - \bx_{3}) .
259: \end{equation}
260: However, with fixed gauges $\bx_0$, $\bx_2$ and $\bx_3$ the coplanarity of
261: $x_0$, $x_2$, $x_3$ and $x_{23}$ can be
262: expressed, by plaing with the gauge of $\bx_{23}$, at most as
263: \begin{equation} \label{eq:gauge-23-bad}
264: \bx_{23} - \bx_0 = a\bx_{2} - b\bx_{3}.
265: \end{equation}
266: Then
267: \begin{equation}\label{eq:gauge-23-wedge}
268: \bx_0\wedge\bx_{12}\wedge\bx_{13}\wedge\bx_{23} = f^{12}f^{13}(a-b)
269: \bx_0\wedge\bx_{1}\wedge\bx_{2}\wedge\bx_{3},
270: \end{equation}
271: and at this moment we use coplanarity of $x_0$, $x_{2}$, $x_{13}$, $x_{23}$,
272: which is equivalent to $a=b$.
273: \end{proof}
274: \begin{Rem}
275: Notice that the whole reasoning
276: can be applied even if the gauges of $\bx_0$ and of
277: $\bx_1$ are not fixed initially (we could then achieve $f^{12}=1$). However we will
278: need this additional restriction in next Sections.
279: \end{Rem}
280: \begin{proof}[Proof of Proposition \ref{lem:BKP-hex}]
281: By the linear algebra, the homogeneous coordinates of the point
282: $x_{123}\in \langle x_3, x_{13}, x_{23} \rangle \cap
283: \langle x_2, x_{12}, x_{23} \rangle \cap
284: \langle x_1, x_{12}, x_{13} \rangle$,
285: in the gauge of Lemma~\ref{lem:BQL-gauge-initial} read
286: \begin{equation*}
287: \frac{1}{\rho}\bx_{123} =
288: \left( f^{13} - f^{12} + \frac{f^{12}f^{13}}{f^{23}} \right)\bx_1 -
289: \left(f^{23} + f^{12} - \frac{f^{12}f^{23}}{f^{13}} \right)\bx_2 +
290: \left( f^{13} - f^{23} + \frac{f^{13}f^{23}}{f^{12}} \right)\bx_3,
291: \end{equation*}
292: (we still keep the undetermined yet factor $\rho$)
293: which gives coplanarity of $x_{123}$, $x_1$, $x_2$ and $x_3$.
294: \end{proof}
295: \begin{Cor} \label{cor:x123-x1}
296: By fixing the gauge function $\rho$ to
297: \begin{equation*}
298: \rho = 2 f^{13} - 2 f^{12} - f^{23} + \frac{f^{13}f^{23}}{f^{12}} +
299: \frac{f^{12}f^{13}}{f^{23}} + \frac{f^{12}f^{23}}{f^{13}},
300: \end{equation*}
301: we find that the
302: linear relations on the new facets (containing $x_{123}$)
303: of the cube are of
304: the form \eqref{eq:BQL-gauge-initial} again, for example
305: \begin{equation*}
306: \bx_{123} - \bx_1 = f^{23}_1
307: (\bx_{12} - \bx_{13}),
308: \end{equation*}
309: where
310: \begin{equation*}
311: f^{23}_1 = \frac{f^{23}}{f^{12}f^{13} - f^{12}f^{23} + f^{13}f^{23}}.
312: \end{equation*}
313: \end{Cor}
314: \begin{Rem}
315: For a geometrically oriented Reader we would like to comment
316: on another interpretation of Proposition~\ref{lem:BKP-hex}. It
317: is related to the notion (see, for example \cite{Coxeter})
318: of the \emph{quadrangular set of points} which are the
319: intersection points of the lines of a complete quadrilateral (add the
320: diagonals). Such a configuration is usually denoted by $\mathsf{Q}(ABC,DEF)$,
321: where the
322: first three points $A,B,C$ lie on sides through one vertex while the remaining
323: three $D,E,F$ lie on the respectively opposite sides, which form a triangle.
324: It is known that $\mathsf{Q}(ABC,DEF)$ implies $\mathsf{Q}(DEF,ABC)$.
325: \begin{figure}
326: \begin{center}
327: \includegraphics{geomBKP}
328: \end{center}
329: \caption{Two quadrangles}
330: \label{fig:quadrangle}
331: \end{figure}
332: In notation of Proposition \ref{lem:BKP-hex}, denote by $\ell$ the intersection line
333: of the plane $\langle x_{1}, x_{2}, x_{3} \rangle $ with the plane
334: $\langle x_{12}, x_{23}, x_{13} \rangle $ (containing also
335: the point $x_{0}$). Denote by $A$, $B$, $C$, $D$, $E$, $F$ intersections of sides
336: of the complete quadrilateral with vertices $x_0, x_{12}, x_{23}, x_{13}$ with
337: $\ell$ (see Figure \ref{fig:quadrangle}), i.e. $\mathsf{Q}(DEF,ABC)$. The
338: statement of the Lemma is equivalent to the fact that the lines
339: $\langle A, x_{1} \rangle $, $\langle B, x_{2} \rangle $,
340: $\langle C, x_{3} \rangle $ intersect in one point (which is $x_{123}$); see
341: Excercise 1 of Section 2.4 of \cite{Coxeter}.
342: \end{Rem}
343: \begin{Rem}
344: As it was pointed to me by Yuri Suris, Proposition \ref{lem:BKP-hex} is
345: equivalent to the M\"{o}bius theorem \cite{Moebius} on mutually inscribed
346: tetrahedra. Indeed, vertices $x_0$, $x_1$, $x_2$, $x_3$ of the tetrahedron
347: $\{x_0, x_1, x_2, x_3 \}$ are contained in the facial planes of the tetrahedron
348: $\{x_{12}, x_{13}, x_{23}, x_{123} \}$, and \emph{vice versa}. In fact,
349: Figure~\ref{fig:quadrangle} appears in M\"{o}bius' original proof of the theorem.
350: \end{Rem}
351: \begin{Rem}
352: Proposition \ref{lem:BKP-hex} can be also considered as a special version of the
353: Miquel theorem \cite{Pedoe} (used in \cite{CDS} to show integrability of the
354: circular lattice) in the same way
355: like Pappus' hexagon theorem is a special case of
356: the Pascal theorem.
357: \end{Rem}
358:
359: We conclude this Section by defining new reduction of the
360: quadrilateral lattice.
361: \begin{Def} \label{def:BQL}
362: A quadrilateral lattice $x:\ZZ^N\to\PP^M$ is called the \emph{B-quadrilateral
363: lattice} if for any triple of different indices $i,j,k$
364: the points $x$, $x_{(ij)}$,
365: $x_{(jk)}$ and $x_{(ik)}$ are coplanar.
366: \end{Def}
367: \begin{Cor} \label{cor:BKP-impl}
368: In the B-quadrilateral
369: lattice, for any triple of different indices $i,j,k$ the points
370: $x_{(i)}$, $x_{(j)}$, $x_{(k)}$ and $x_{(ijk)}$ are coplanar.
371: \end{Cor}
372:
373: \subsection{Multidimensional consistency of the BQL constraint}
374: As it was shown in \cite{MQL} the planarity condition, which allows to construct
375: the point $x_{123}$ as in Lemma \ref{lem:gen-hex}, does not lead to any further
376: restrictions if we increase dimension of the lattice. This is the consequence of
377: the of the following geometric observation.
378: \begin{Lem} \label{lem:4D-consist-QL}
379: Consider points $x_0$, $x_1$, $x_2$, $x_3$ and $x_4$
380: in general position in $\PP^M$, $M\geq 4$. Choose generic points
381: $x_{ij}\in\langle x_0, x_i, x_j \rangle$, $1\leq i < j \leq 4$,
382: on the corresponding planes, and using
383: the planarity condition construct the points
384: $x_{ijk}\in\langle x_0, x_i, x_j , x_k\rangle$, $1\leq i < j < k \leq 4$ -- the
385: remaining vertices of the four (combinatorial) cubes.
386: Then the intersection point $x_{1234}$ of the three planes
387: \[\langle x_{12}, x_{123}, x_{124} \rangle, \;
388: \langle x_{13}, x_{123}, x_{134} \rangle, \;
389: \langle x_{14}, x_{124}, x_{134} \rangle \quad \text{in} \quad
390: \langle x_{1}, x_{12}, x_{13}, x_{14} \rangle,
391: \]
392: coincides with
393: the intersection point of the three planes
394: \[\langle x_{12}, x_{123}, x_{124} \rangle, \;
395: \langle x_{23}, x_{123}, x_{234} \rangle, \;
396: \langle x_{24}, x_{124}, x_{234} \rangle, \quad \text{in} \quad
397: \langle x_{2}, x_{12}, x_{23}, x_{24} \rangle,
398: \]
399: which is the same as
400: the intersection point of the three planes
401: \[\langle x_{13}, x_{123}, x_{134} \rangle, \;
402: \langle x_{23}, x_{123}, x_{234} \rangle, \;
403: \langle x_{34}, x_{134}, x_{234} \rangle, \quad \text{in} \quad
404: \langle x_{3}, x_{13}, x_{23}, x_{34} \rangle,
405: \]
406: and
407: the intersection point of the three planes
408: \[\langle x_{14}, x_{124}, x_{134} \rangle, \;
409: \langle x_{24}, x_{124}, x_{234} \rangle, \;
410: \langle x_{34}, x_{134}, x_{234} \rangle, \quad \text{in} \quad
411: \langle x_{4}, x_{14}, x_{24}, x_{34} \rangle.
412: \]
413: \end{Lem}
414: \begin{Rem}
415: In fact, the point $x_{1234}$ is the unique
416: intersection point of the four three dimensional subspaces
417: $\langle x_{1}, x_{12}, x_{13}, x_{14} \rangle$,
418: $\langle x_{2}, x_{12}, x_{23}, x_{24} \rangle$,
419: $\langle x_{3}, x_{13}, x_{23}, x_{34} \rangle$,
420: and
421: $\langle x_{4}, x_{14}, x_{24}, x_{34} \rangle$ of the four dimensional subspace
422: $\langle x_{0}, x_{1}, x_{2}, x_{3} , x_{4} \rangle$. This observation
423: generalizes naturally to the case of more dimensional hypercube with the
424: planar facets.
425: \end{Rem}
426: The goal of this Section is to show an analogous result for B-quadrilateral
427: lattice. Notice, that in previously known reductions of the quadrilateral
428: lattice, such as the symmetric \cite{DS-sym} or the quadratic \cite{q-red}
429: reduction, the additional constraint was
430: imposed on initial quadrilaterals. Then the multidimensional consistency of the
431: reduction was the
432: result of the three dimensional consistency of the constraint and the
433: multidimensional consistency of the quadrilateral lattice.
434:
435: In the BQL case,
436: the constraint is imposed on the level of elementary cubes. Therefore
437: its four dimensional consistency
438: is crucial for integrability of the B-quadrilateral lattice, and once proven,
439: implies consistency of the reduction in more dimensions.
440: \begin{Prop} \label{lem:4D-consist-BQL}
441: Under hypotheses of Lemma \ref{lem:4D-consist-QL}, assume that the BQL
442: condition holds for the initial data, i.e., the point $x_0$ belongs to the four
443: planes $\langle x_{ij}, x_{ik}, x_{jk} \rangle$, $1\leq i<j<k \leq 4$. Then all
444: the three dimensional (combinatorial) cubes obtained in the construction satisfy
445: the BQL constaint, i.e.,
446: \begin{equation*}
447: x_1 \in \langle x_{123}, x_{124}, x_{134} \rangle, \quad
448: x_2 \in \langle x_{123}, x_{124}, x_{234} \rangle, \quad
449: x_3 \in \langle x_{123}, x_{134}, x_{234} \rangle, \quad
450: x_4 \in \langle x_{124}, x_{134}, x_{234} \rangle.
451: \end{equation*}
452: \end{Prop}
453: \begin{proof}
454: Consider the gauge of Lemma~\ref{lem:BQL-gauge-initial}.
455: If we add into the construction points $x_4$ and $x_{14}$,
456: then by fixing suitably
457: their gauges $\bx_4$ and $\bx_{14}$, we
458: can rewrite the coplanarity condition of $x_0$, $x_1$, $x_4$ and $x_{14}$
459: in the form \eqref{eq:BQL-gauge-initial}.
460: The same argument like in the proof of Lemma~\ref{lem:BQL-gauge-initial} implies
461: that the algebraic coplanarity conditions of
462: $x_0$, $x_i$, $x_4$ and $x_{i4}$, $i=2,3$ take the form of equation
463: \eqref{eq:BQL-gauge-initial}.
464:
465: By fixing gauges of points $x_{ijk}$ like in Corollary~\ref{cor:x123-x1},
466: we obtain the relations
467: \begin{equation} \label{eq:x_ijk-x_i}
468: \bx_{ijk} - \bx_{i} = f^{jk}_{i} (\bx_{ij} - \bx_{ik} ), \qquad i,j,k
469: \quad \text{disctinct},
470: \end{equation}
471: where
472: \begin{equation} \label{eq:star-triangle}
473: f^{jk}_{i} = \frac{f^{jk}}{f^{ij}f^{ik} - f^{ij}f^{jk} + f^{ik}f^{jk}},
474: \end{equation}
475: with $f^{ji}=-f^{ij}$.
476:
477: In equations \eqref{eq:x_ijk-x_i} let us fix $i=1$ and consider the three
478: pairs $(j,k)$: $(2,3)$, $(2,4)$ and $(3,4)$. Then after simple calculation
479: we obtain
480: the following linear relation
481: \begin{equation} \label{eq:4D-BQL-x1}
482: f^{24}_1f^{34}_1 (\bx_{123} - \bx_1) - f^{23}_1 f^{34}_1 (\bx_{124} - \bx_1) +
483: f^{23}_1 f^{24}_1 (\bx_{134} - \bx_1) =0,
484: \end{equation}
485: which shows that $x_1 \in \langle x_{123}, x_{124}, x_{134} \rangle$. Other
486: cases are similar.
487: \end{proof}
488: \begin{Cor}
489: Under assumptions of Proposition \ref{lem:4D-consist-BQL}, the point $x_{1234}$
490: belongs to the
491: four planes: $\langle x_{12}, x_{13}, x_{14} \rangle$,
492: $\langle x_{12}, x_{23}, x_{24} \rangle$,
493: $\langle x_{13}, x_{23}, x_{34} \rangle$ and
494: $\langle x_{14}, x_{24}, x_{34} \rangle$.
495: \end{Cor}
496: \begin{Rem}
497: The same procedure can be applied when we increase
498: dimension of the hypercube keeping the BQL constraint.
499: \end{Rem}
500:
501: \subsection{BQL and the discrete BKP equation}
502: \begin{Prop}
503: A quadrilateral lattice $x:\ZZ^N\to\PP^M$ is a B-quadrilateral lattice if and
504: only if
505: it allows for a homogoneous representation $\bx:\ZZ^N\to\RR^{M+1}_{*}$
506: satisfying the system of discrete Moutard equations (the
507: discrete BKP linear problem)
508: \begin{equation} \label{eq:BKP-lin}
509: \bx_{(ij)} - \bx = f^{ij} (\bx_{(i)} - \bx_{(j)}) , \quad 1\leq i< j\leq N,
510: \end{equation}
511: for suitable functions $f^{ij}:\ZZ^N\to\RR$.
512: \end{Prop}
513: \begin{proof}
514: As we have shown above (we present an alternative difference-equation theory
515: proof in the Appendix), the B-quadrilateral lattices indeed allow for such a
516: gauge; here the Remark after Lemma~\ref{lem:BQL-gauge-initial} turns out to be
517: important. Conversely, three equations \eqref{eq:BKP-lin} for the
518: pairs $(i,j)$, $(i,k)$, $(j,k)$ imply the linear relation
519: \begin{equation} \label{eq:M-BQL}
520: f^{jk}f^{ik} (\bx_{(ij)} - \bx) + f^{ij}f^{ik} (\bx_{(jk)} - \bx) -
521: f^{ij}f^{jk}(\bx_{(ik)} - \bx) =0, \qquad1\leq i < j < k \leq N,
522: \end{equation}
523: expressing coplanarity of the
524: four points $x$, $x_{(ij)}$, $x_{(jk)}$ and $x_{(ik)}$.
525: \end{proof}
526: The system \eqref{eq:BKP-lin} is well known in the literature \cite{NiSchief}.
527: Its compatibility leads to the following set of nonlinear equations
528: \begin{equation} \label{eq:nonl-BQL-f}
529: 1 + f^{jk}_{(i)}(f^{ij} - f^{ik}) = f^{ik}_{(j)} f^{ij}= f^{ij}_{(k)}f^{ik},
530: \qquad i,j,k \quad \text{distinct},
531: \end{equation}
532: with $f^{ji}=-f^{ij}$.
533:
534: \begin{Rem}
535: The system \eqref{eq:M-BQL}
536: can be actually solved, see for example \cite{BobSur},
537: (we have used this fact in a hidden form already).
538: Simply replace the lower index $i$ in the "star-triangle" relation
539: \eqref{eq:star-triangle} by the shift $(i)$.
540: \end{Rem}
541:
542: On the other side, the second equality in the compatibility condition
543: implies existence of the potential $\tau:\ZZ^N\to\RR$,
544: in terms of which the functions $f^{ij}$ can be written as
545: \begin{equation} \label{eq:tau}
546: f^{ij} = \frac{\tau_{(i)}\tau_{(j)}}{\tau \, \tau_{(ij)}}, \qquad i\ne j .
547: \end{equation}
548: The first equality can be then rewritten in the form of the
549: system of the discrete
550: BKP equations \cite{Miwa}
551: \begin{equation} \label{eq:BKP-nlin}
552: \tau\, \tau_{(ijk)} = \tau_{(ij)}\tau_{(k)} - \tau_{(ik)}\tau_{(j)} +
553: \tau_{(jk)}\tau_{(i)}, \quad 1\leq i< j < k \leq N.
554: \end{equation}
555:
556: %
557: \begin{Rem}
558: Two dimensional quadrilateral lattices whose homogeneous coordinates satisfy
559: (up to a gauge)
560: equation \eqref{eq:BKP-lin} are characterized geometrically \cite{DGNS}
561: by condition that any point $x$ and its four
562: second-order neighbours $x_{(\pm 1 \pm 2)}$ are contained in a subspace of
563: dimension three. Obviously, any two dimensional slide of the B-quadrilateral
564: lattice fulfills this property, which can therefore serve as definition of a
565: two dimensional BQL. However the example of
566: the standard injection $\ZZ^N\to\RR^N\subset\PP^N$ shows that without additional
567: requirements this property does not characterize completly multidimensional BQL.
568: \end{Rem}
569: %
570: \subsection{The $\tau$-functions} \label{sec:tau}
571: In this Section we present relation between the $\tau$-function of the
572: quadrilateral lattice \cite{DS-sym}, which we donote here by $\tilde{\tau}$,
573: and the
574: above $\tau$-function of the B-quadrilateral lattice. From the relation
575: between
576: the KP and BKP hierarchies \cite{DKJM} we expect that
577: within the class of the
578: B-quadrilateral lattices $\tilde{\tau}$ should be equal to the square of $\tau$.
579:
580: Let us recall briefly the algebraic construction of the $\tau$-function of the
581: quadrilateral lattice (the
582: geometric meaning is presented in \cite{DS-sym}).
583: The non-homogeneous coordinates $\mathrm{\bx}:\ZZ^N\to\RR^M$
584: (we restrict our attention to the affine
585: geometric aspects of the theory) of the quadrilateral lattice satisfy
586: the system of Laplace equations
587: \begin{equation} \label{eq:Laplace-a}
588: \bx_{(ij)} - \bx = a^{ij}(\bx_{(i)} - \bx) + a^{ji}(\bx_{(j)}-\bx), \quad i\ne
589: j.
590: \end{equation}
591: The functions $a^{ij}:\ZZ^N \to\RR$ are not arbitrary (the system
592: \eqref{eq:Laplace-a} must be compatible), in particular they
593: can be parametrized in terms of the potentials $h^i$ (the Lam\'{e} coefficients)
594: as follows
595: \begin{equation} \label{eq:a-h}
596: a^{ij} = \frac{h_{i(j)}}{h_i}, \quad i\ne j.
597: \end{equation}
598: Define the so called rotation coefficients $\beta_{ij}$ from equations
599: \begin{equation} \label{eq:lin-h}
600: \Delta_i h_j = h_{i(j)}\beta_{ij}, \quad i\ne j,
601: \end{equation}
602: and the normalized tangent vectors $\bX_i$ from
603: \begin{equation} \label{eq:x-hX}
604: \Delta_i \bx = h_i\bX_i .
605: \end{equation}
606: Then the Laplace system \eqref{eq:Laplace-a} takes the first
607: order form
608: \begin{equation} \label{eq:lin-X}
609: \Delta_j \bX_i = \beta_{ij}\bX_j ,\quad i\ne j,
610: \end{equation}
611: and its compatibility reads
612: \begin{equation} \label{eq:nlin-beta}
613: \Delta_j\beta_{ik} = \beta_{ij(k)}\beta_{jk}, \quad i,j,k \quad
614: \text{distinct}.
615: \end{equation}
616: The discrete Darboux equations \eqref{eq:nlin-beta} imply existence of the
617: potential $\tilde\tau$ (the $\tau$-function of the quadrilateral lattice)
618: \begin{equation} \label{eq:tau-QL}
619: \frac{\tilde{\tau}\tilde{\tau}_{(ij)}}{\tilde{\tau}_{(i)}\tilde{\tau}_{(j)}} =
620: 1 - \beta_{ij}\beta_{ji},
621: \quad i\ne j.
622: \end{equation}
623:
624: In looking for the Lam\'{e} coefficients $h_i$ in the reduction from QL to BQL
625: we can
626: compare both linear systems
627: \eqref{eq:BKP-lin} and \eqref{eq:Laplace-a}, and the
628: expresions \eqref{eq:a-h} and \eqref{eq:tau} to
629: obtain
630: \begin{equation} \label{eq:h-tau}
631: h_i = (-1)^{\sum_{k<i}m_k}\frac{\tau}{\tau_{(i)}}.
632: \end{equation}
633: The corresponding rotation coefficients are then given by (below we assume $i<j$)
634: \begin{align} \label{eq:beta-BQL-ij}
635: \beta_{ij} & = - (-1)^{\sum_{i\leq k<j}m_k}
636: \left( \frac{\tau_{(i)}}{\tau} + \frac{\tau_{(ij)}}{\tau_{(j)}}\right)
637: \frac{\tau}{\tau_{(j)}},\\
638: \label{eq:beta-BQL-ji}
639: \beta_{ji} & = - (-1)^{\sum_{i\leq k<j}m_k}
640: \left( \frac{\tau_{(j)}}{\tau} - \frac{\tau_{(ij)}}{\tau_{(i)}}\right)
641: \frac{\tau}{\tau_{(i)}},
642: \end{align}
643: which implies (compare with formula \eqref{eq:tau-QL})
644: \begin{equation}
645: 1 - \beta_{ij}\beta_{ji} = \left(
646: \frac{{\tau}{\tau}_{(ij)}}{{\tau}_{(i)}{\tau}_{(j)}} \right)^2,
647: \qquad i\ne j.
648: \end{equation}
649: Therefore, we can summarize the above considerations as follows.
650: \begin{Prop}
651: Given B-quadrilateral lattice $x$ with the $\tau$-function $\tau$. Then,
652: formally on the level of the reduction of the system of
653: discrete affine Laplace equations
654: \eqref{eq:Laplace-a} to the system of discrete Moutard equations
655: \eqref{eq:BKP-lin}, the Lam\'{e} functions and the rotation coefficients
656: are given in terms of $\tau$ by equations
657: \eqref{eq:h-tau}-\eqref{eq:beta-BQL-ji}, and the corresponding $\tau$-function
658: $\tilde{\tau}$ of the quadrilateral lattice is given as
659: \begin{equation} \label{eq:tau-QL-BQL}
660: \tilde{\tau} = \tau^2.
661: \end{equation}
662: \end{Prop}
663: \begin{Rem}
664: We should be aware that the $\tau$-function of the quadrilateral lattice
665: is defined with
666: respect to the coefficients of the affine Laplace equation. Equation
667: \eqref{eq:BKP-lin},
668: although formally written in the affine form, is a consequence of the
669: projectively-invariant definition of the B-quadrilateral lattice.
670: Therefore, in this formal correspondence the
671: geometric meaning of the rotation coefficients
672: in the BQL reduction has been lost. We mention that the
673: affine geometric meaning of equation
674: \eqref{eq:BKP-lin}, and therefore also of the rotation coefficients
675: \eqref{eq:beta-BQL-ij}-\eqref{eq:beta-BQL-ji}, can be
676: provided within the context of the trapezoidal lattices
677: \cite{BobSur}
678: (see also \cite{Sauer,KoSch-trap}).
679: \end{Rem}
680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
681: \section{Algebro-geometric construction of the BQL} \label{sec:BQL-Prym}
682: Below we apply the algebro-geometric approach, well known in the theory if
683: integrable systems \cite{BBEIM}, to the B-quadrilateral lattice reduction.
684: It is known \cite{DGNS} that in the BQL
685: reduction case, the generic algebraic curve, used to generate solutions of the
686: discrete Darboux equations \cite{AKV}, should be replaced by a curve admitting
687: a holomorphic involution with two fixed points. Such curves have been already
688: used in construction of solutions of the BKP hierarchy \cite{DJKM-Prym-BKP}
689: and of its two-component generalization \cite{VeselovNovikov}. We develop the
690: corresponding results of \cite{DGNS} and we present the explicit formulas for
691: the lattice points and the solutions of the discrete BKP equation in terms of
692: the Prym theta functions related to such special curves.
693:
694: \subsection{Curves with involution and their Prym varieties}
695: Let us first summarize some fact from theory of Riemann surfaces (see
696: \cite{FarkasKra,Fay}).
697: Consider $\hat{\Gamma} \stackrel{\pi}{\to}\Gamma$ a ramified double
698: covering of genus $\hat{g}=2g$ of a compact Riemann surface
699: $\Gamma$ of
700: genus $g$ with exactly two branch points $Q_0$, $Q_\infty$.
701: Denote by $\sigma:\hat{\Gamma}\to\hat{\Gamma}$ the holomorphic involution
702: permuting
703: sheets of the covering, i.e., $\Gamma =\hat{\Gamma}/\sigma$.
704:
705: The map $\pi^*:\mathcal{J}({\Gamma})\to\mathcal{J}({\hat{\Gamma}})$ lifting
706: divisor classes of degree $0$ is an injection.
707: The holomorphic involution $\sigma$ extends to
708: $\mathcal{J}({\hat{\Gamma}})$ and allows to define the Prym variety
709: \begin{equation} \label{def:Prym-variety}
710: \mathcal{P}_\sigma({\hat{\Gamma}}) = \{ A- \sigma(A) \, |A\in
711: \mathcal{J}(\hat{\Gamma}) \}.
712: %= \{ B\in\mathcal{J}(\hat{\Gamma}) \, | \sigma(B) = -B \}.
713: \end{equation}
714: The natural epimorphism
715: $i:\mathcal{J}({\Gamma})\times \mathcal{P}_\sigma({\hat{\Gamma}})
716: \to\mathcal{J}({\hat{\Gamma}})$ has finite kernel consisting of $4^g$
717: half-periods in $\mathcal{J}({\Gamma})$.
718:
719: There exists a basis of cycles $a_i,b_i$, $1\leq i \leq 2g$ on $\hat{\Gamma}$
720: with the canonical intersection matrix
721: such that $\pi(a_k),\pi(b_k)$, $1\leq k \leq g$, is a canonical basis of cycles on
722: $\Gamma$ and
723: \begin{equation*}
724: \sigma(a_k) = - a_{g+k}, \quad \sigma(b_k) = - b_{g+k} ,
725: \qquad 1\leq k \leq g .
726: \end{equation*}
727: The corresponding normalized holomorphic differentials $\omega_i$
728: \begin{equation} \label{eq:a}
729: \oint_{a_j}\omega_i = \delta_{ij}, \qquad 1\leq i,j \leq 2g,
730: \end{equation}
731: satisfy
732: \begin{equation}
733: \sigma^*(\omega_k) =-\omega_{g+k},\quad
734: \sigma^*(\omega_{g+k}) =-\omega_{k}, \qquad 1\leq k \leq g.
735: \end{equation}
736: The differentials
737: \begin{equation*}
738: u_k = \omega_k - \omega_{g+k}, \qquad \sigma^*(u_k) = u_k,
739: \qquad 1\leq k \leq g,
740: \end{equation*}
741: form a basis of normalized holomorphic differentials on
742: $\Gamma$, while the
743: odd differentials
744: \begin{equation} \label{eq:w}
745: w_k = \omega_k + \omega_{g+k}, \qquad \sigma^*(w_k) = - w_k,
746: \qquad 1\leq k \leq g,
747: \end{equation}
748: are called normalized holomorphic
749: Prym differentials. Then the Riemann matrix
750: \begin{equation}
751: \hat{B}_{jk} = \oint_{b_j}\omega_k, \qquad \qquad 1\leq j,k \leq \hat{g}
752: \end{equation}
753: for
754: $\hat{\Gamma}$
755: has the form
756: \begin{equation}
757: \hat{B} = \frac{1}{2}\left( \begin{array}{cc}
758: \Pi + B & \Pi - B \\
759: \Pi - B & \Pi + B
760: \end{array} \right),
761: \end{equation}
762: where
763: \begin{equation}
764: B_{jk} = \oint_{b_j}u_k, \qquad \qquad 1\leq j,k \leq g
765: \end{equation}
766: is the corresponding Riemann matrix for $\Gamma$,
767: and $\Pi$ is the matrix of the $b$-periods of the Prym differentials
768: \begin{equation} \label{eq:Pi}
769: \Pi_{jk} = \oint_{b_j}w_k, \qquad \qquad 1\leq j,k \leq g.
770: \end{equation}
771: The matrix $\Pi$ is symmetric and has positively defined imaginary part, and
772: defines the Prym theta function $\theta(\mathbf{z} ;\Pi)$,
773: $\mathbf{z}\in\mathbb{C}^{g}$,
774: \begin{equation}
775: \theta(\mathbf{z};\Pi) = \sum_{\mathbf{n}\in{\ZZ^{{g}}}}
776: \exp \left\{ \pi i \langle \mathbf{n},{\Pi} \mathbf{n} \rangle +
777: 2 \pi i\langle \mathbf{n}, \mathbf{z} \rangle \right\},
778: \end{equation}
779: where $\langle \cdot , \cdot \rangle$ denotes the standard bilinear
780: form in $\mathbb{C}^{{g}}$.
781:
782:
783: With the above choice of the period matrix $B$ for $\hat{\Gamma}$
784: and with $Q_\infty$ as the base-point of the Abel map
785: \begin{equation}
786: \mathbf{A}(P)=\int_{Q_\infty}^P \boldsymbol{\omega},
787: \qquad \boldsymbol{\omega}=(\omega_1,\dots ,\omega_{2g}),
788: \end{equation}
789: the lift of $\sigma$ from
790: $\mathcal{J}(\hat{\Gamma})=\mathbb{C}^{2g}/(I_{2g},\hat{B})$ to
791: $\mathbb{C}^{2g}$ reads
792: \begin{equation*}
793: \sigma( z_1, \dots , z_{2g}) =
794: -(z_{g +1},\dots , z_{2g}, z_1,\dots ,z_{g}),
795: \end{equation*}
796: while the map $\pi^*:\mathcal{J}(\Gamma)\to\mathcal{J}(\hat{\Gamma})$
797: is represented by
798: \begin{equation*}
799: \pi^*(z_1,\dots ,z_{g}) = (z_1,\dots ,z_{g},-z_1,\dots ,-z_{g}).
800: \end{equation*}
801: The Prym variety is a principally polarized abelian variety isomorphic to
802: $\mathcal{P}=\mathbb{C}^{g}/(I_{g},\Pi)$, and the injection
803: $\phi: \mathbb{C}^{g}/(I_{g},\Pi) \to \mathbb{C}^{2g}/(I_{2g},\hat{B})$
804: is given by
805: \begin{equation*}
806: \phi(z_1,\dots ,z_{g}) = (z_1,\dots ,z_{g},z_1,\dots ,z_{g}) .
807: \end{equation*}
808: There holds the following analog of the Riemann theorem.
809: \begin{Lem}[\cite{Fay}] \label{lem:R-P}
810: Given $\mathbf{e} \in \mathbb{C}^{g}$ such that
811: $\theta\left(\mathbf{e};\Pi\right) \neq 0$, then the zero divisor $Z$ of
812: $\theta\left(\int_{Q_\infty}^P\boldsymbol{w} - \mathbf{e};\Pi\right)$ on
813: $\hat{\Gamma}$
814: is of degree $2g$ and satisfies the relation
815: \begin{equation} \label{eq:R-P}
816: \phi(\mathbf{e}) =
817: \mathbf{A}(Z)-\mathbf{A}( Q_0 ) + \pi^* \mathbf{K}_\Gamma,
818: \qquad \mathrm{mod} \; (I_{2g},\hat{B}),
819: \end{equation}
820: where $\mathbf{K}_\Gamma$ is the Riemann constants vector of $\Gamma$.
821: Moreover $Z + \sigma(Z) -Q_0 - Q_\infty$ is a canonical divisor on
822: $\hat{\Gamma}$.
823: \end{Lem}
824: %
825: By $\omega_{ST}$, ($S,T\in \hat{\Gamma}$, $S\neq T$) denote the unique
826: meromorphic
827: differential holomorphic in $\hat{\Gamma}\setminus\{S,T\}$ with
828: poles of the first
829: order in $S$, $T$ with residues, correspondingly, $1$ and $-1$,
830: and normalized by conditions
831: \begin{equation} \label{eq:omST-a}
832: \oint_{a_j}\omega_{ST} = 0, \qquad 1\leq j \leq 2g.
833: \end{equation}
834: It is known that the $b$-periods of such differentials are given by
835: \begin{equation} \label{eq:omST-b}
836: \oint_{b_j}\omega_{ST} = 2\pi i \int_{T}^S \omega_j, \qquad 1\leq j \leq 2g,
837: \end{equation}
838: with the integral being taken along a curve joining $S$ to $T$ in
839: $\hat{\Gamma}\setminus \bigcup_{j=1}^{2g}a_j \setminus \bigcup_{j=1}^{2g}b_j$.
840: Moreover, the following relation between two such differentials holds
841: (with the paths of integration being appropriately choosen \cite{FarkasKra})
842: \begin{equation}\label{eq:omST-PQ}
843: \int_{P}^Q \omega_{ST} = \int_{S}^T \omega_{PQ}.
844: \end{equation}
845: %
846: \begin{Rem}
847: Such a form can be expressed in terms of the theta function on $\hat{\Gamma}$
848: as
849: \begin{equation*}
850: \omega_{ST} = d_P\log\frac{\theta(\mathbf{A}(P) - \mathbf{A}(S) -
851: \boldsymbol{\xi};\hat{B})}
852: {\theta(\mathbf{A}(P) - \mathbf{A}(T) - \boldsymbol{\xi};\hat{B}))},
853: \end{equation*}
854: where $\boldsymbol{\xi}$ is a general point of the divisor $\Theta$ of zeros of
855: the theta function in
856: $\mathcal{J}(\hat{\Gamma})$.
857: \end{Rem}
858: %
859: \subsection{Explicit algebro-geometric formulas}
860: Given $N$ points $Q_i\in\Gamma$, different from $Q_0, Q_\infty$,
861: and an effective
862: non-special divisor ${D}$ of degree $\hat{g}$ such that
863: \begin{equation} \label{eq:D-C}
864: D + \sigma(D) - Q_0 - Q_\infty \stackrel{\mathcal{J}_{2\hat{g}-2}({\hat{\Gamma}})}{=}
865: C_{\hat{\Gamma}}
866: \end{equation}
867: is a canonical divisor. For an arbitrary $m\in\ZZ^N$ there exists \cite{DGNS}
868: the unique function
869: $\psi(m)$ meromorphic on $\Gamma$ having in points $Q_i$ (in points
870: $\sigma(Q_i)$) poles
871: (corrspondingly, zeros) of the order $m_i$, no other singularities except for possible
872: simple poles in points of the divisor $D$, and normalized to $1$ at $Q_\infty$.
873: In \cite{DGNS} it was shown that, as a function of the discrete parameter $m$,
874: the wave function $\psi$ satisies the system of the discrete Moutard equations
875: \begin{equation} \label{eq:d-Moutard-ag}
876: \psi_{(ij)}(P) - \psi(P) = f^{ij}(\psi_{(i)}(P) - \psi_{(j)}(P) ),
877: \qquad 1\leq i < j \leq N,
878: \end{equation}
879: where
880: \begin{equation} \label{eq:f-ij-alg-def}
881: f^{ij} = \lim_{P\to Q_i}\frac{\psi_{(ij)}(P)}{\psi_{(i)}(P)} =
882: -\lim_{P\to Q_j}\frac{\psi_{(ij)}(P)}{\psi_{(j)}(P)} ,
883: \qquad i < j .
884: \end{equation}
885: To obtain B-quadrilateral lattices we pick up $M+1$ points $P_1,\dots
886: ,P_{M+1}\in\Gamma$ of the Riemann surface. Then $x_j(m) = \Psi(m|P_j)$, $1\leq j \leq
887: M+1$ serve as homogeneous coordinates of the lattice. However in this way we obtain
888: B-quadrilateral lattices in complex projective space. In order to get real
889: lattices certain additional restrictions, which were also given in \cite{DGNS},
890: should be imposed on the algebro-geometric data.
891:
892: In \cite{DGNS} the multidimensional aspects of the system
893: \eqref{eq:d-Moutard-ag} were not of particular importance. Also the role of the
894: Prym variety and of the corresponding theta function was not fully exploited.
895: Our goal here is to fill this point. We start from the
896: imediate consequence of
897: Definition \eqref{def:Prym-variety} of the Prym variety.
898: \begin{Cor}
899: Denote by $D(m)$ the divisor of additional zeros of $\psi(m)$, then
900: \begin{equation}
901: D(m) - D \stackrel{\mathcal{J}({\hat{\Gamma}})}{=}
902: \sum_{k=1}^N m_k \left(\sigma(Q_k) - Q_k \right)
903: \end{equation}
904: moves linearly within the Prym variety.
905: \end{Cor}
906:
907: An important part of the algebro-geometric theory of integrable systems consists
908: on providing the explicit formulas, in terms of the Riemann theta functions of
909: the corresponding Jacobi varieties, for the wave functions and the soliton
910: fields. In the case of the special Riemann surfaces used in the paper, there
911: exist \cite{Fay} formulas connecting the theta functions of $\hat\Gamma$,
912: $\Gamma$ and $\mathcal{P}_\sigma$. However, instead of reducing
913: the explicit expressions given in \cite{AKV,DGNS} for the generic curves,
914: we will follow the reasoning of \cite{DJKM-Prym-BKP}.
915: In order to present the explicit
916: formulas, in terms of the (Riemann--) Prym theta function, for the wave function
917: and other relevant data we will use Lema~\ref{lem:R-P}.
918:
919: Let us define
920: \begin{equation*}
921: \mathbf{V}_k = \int_{Q_\infty}^{Q_k}\!\!\!\boldsymbol{w} \in\mathbb{C}^g,
922: \qquad 1\leq k \leq N, \qquad \boldsymbol{w} = (w_1,\dots , w_g),
923: \end{equation*}
924: then equations \eqref{eq:w} and \eqref{eq:omST-b} imply that
925: \begin{equation} \label{eq:V-om}
926: \phi(\mathbf{V}_k) = \int_{\sigma(Q_k)}^{Q_k}\!\!\! \boldsymbol{\omega} =
927: -\frac{1}{2\pi i}\oint_{\boldsymbol{b}}\omega_{\sigma(Q_k) Q_k},
928: \qquad \boldsymbol{b} = (b_1,\dots , b_{2g}).
929: \end{equation}
930: \begin{Prop}
931: The BQL wave function $\psi(m)$ can be written down
932: with the help of the Prym theta functions as follows
933: \begin{equation} \label{eq:psi}
934: \psi(m|P)= \frac{\theta ( \int_{Q_\infty}^P\!\! \boldsymbol{w}-
935: \sum_{k=1}^N m_k \mathbf{V}_k - \mathbf{e} ;\Pi )\, \theta( \mathbf{e};\Pi )}
936: {\theta ( \sum_{k=1}^N m_k \mathbf{V}_k + \mathbf{e} ;\Pi) \,
937: \theta( \int_{Q_\infty}^P\!\! \boldsymbol{w}- \mathbf{e};\Pi )}
938: \exp \left( \sum_{k=1}^N m_k \int_{Q_\infty}^P \!\!\!
939: \omega_{\sigma(Q_k) Q_k} \right),
940: \end{equation}
941: where
942: \begin{equation} \label{eq:Z}
943: \phi(\mathbf{e}) = \mathbf{A}(D) -\mathbf{A}(Q_0) + \pi^*(\mathbf{K}_\Gamma) .
944: %\qquad \mathrm{mod} \; (I_{\hat{g}},\hat{B}).
945: \end{equation}
946: \end{Prop}
947: \begin{proof}
948: Using the property \eqref{eq:D-C} of the divisor $D$, the Hurwitz formula
949: \begin{equation}
950: C_{\hat{\Gamma}} \stackrel{\mathcal{J}_{2\hat{g} -2}(\Gamma)}{=}
951: \pi^* C_\Gamma + Q_0 + Q_\infty
952: \end{equation}
953: relating canonical divisors on $\hat{\Gamma}$ and $\Gamma$, and the relation
954: \begin{equation}
955: \mathbf{A}(C_{\hat{\Gamma}}) = - 2 \mathbf{K}_{\hat{\Gamma}}
956: \end{equation}
957: between the canonical divisor and the Riemann constants vector, we obtain
958: \begin{equation*}
959: \sigma(\mathbf{A}(D) -\mathbf{A}(Q_0) + \pi^*(\mathbf{K}_\Gamma))
960: =-\mathbf{A}(D) +\mathbf{A}(Q_0) - \pi^*(\mathbf{K}_\Gamma),
961: \end{equation*}
962: which asserts that the definition of the vector $\mathbf{e}$ in \eqref{eq:Z}
963: is meaningful.
964:
965: To show that the right hand side of equation \eqref{eq:psi} is single valued on
966: $\hat{\Gamma}$ we check that it is independent on the integration path in
967: the integrals $\int_{Q_\infty}^P$. When two paths differ by an elementary cycle
968: we use the properties \eqref{eq:a}-\eqref{eq:Pi} of the holomorphic
969: differentials,
970: the quasi-periodicity properties of the theta
971: functions
972: \begin{equation*}
973: \theta(\mathbf{z}+\mathbf{e}_k;{\Pi}) = \theta(\mathbf{z};{\Pi}),\qquad
974: \theta(\mathbf{z}+\Pi \mathbf{e}_k;{\Pi}) = \exp(-\pi i \Pi_{kk} - 2\pi i z_k)
975: \theta(\mathbf{z};{\Pi}),
976: \end{equation*}
977: where $\mathbf{e}_k$ are vectors of the standard basis in $\mathbb{C}^g$,
978: and the relations \eqref{eq:omST-a} and \eqref{eq:V-om}.
979: From now on our path of integration avoids the cuts, like in fromulas
980: \eqref{eq:omST-b}-\eqref{eq:omST-PQ}.
981:
982: As the normalizaton condition at $Q_\infty$ is obvious (the theta function is
983: even) we are left with the analyticity properties. Lemma \ref{lem:R-P} implies
984: that the right hand side has simple poles at points of the divisor $D$. Apart
985: from the zeros of the theta function in the nominator (which may
986: eventually cancel with the poles at $D$), the only other
987: poles and zeros are consequences of the analytical properties of the integral in
988: the exponential part. Let us choose a local parameter $z_k(P)$ at $Q_k$, then
989: \begin{equation*}
990: \omega_{\sigma(Q_k) Q_k}(P) \stackrel{P\to Q_k}{=}\left(-\frac{1}{z_k(P)} +
991: \dots \right)dz_k(P),
992: \end{equation*}
993: which implies that
994: \begin{equation*}
995: \int_{Q_\infty}^P \!\!\!
996: \omega_{\sigma(Q_k) Q_k} \stackrel{P\to Q_k}{=} - \log z_k(P) + O(1),
997: \end{equation*}
998: and, in consequence, the right hand side in equation \eqref{eq:psi} has pole
999: of order $m_k$ at $Q_k$. Similarly, since
1000: $z_k(\sigma(P))$ is a local parameter at
1001: $\sigma(Q_k)$, we have
1002: \begin{equation*}
1003: \int_{Q_\infty}^P \!\!\!
1004: \omega_{\sigma(Q_k) Q_k}
1005: \stackrel{P\to \sigma(Q_k)}{=} \log z_k(\sigma(P)) + O(1),
1006: \end{equation*}
1007: and the right hand side in equation \eqref{eq:psi} has zero
1008: of order $m_k$ at $\sigma(Q_k)$.
1009: \end{proof}
1010: \begin{Cor}
1011: The potentials read
1012: \begin{equation} \label{eq:f-ij-alg}
1013: f^{ij}(m) =
1014: \frac{\theta (\mathbf{V}_i +\sum_{k=1}^N m_k \mathbf{V}_k +\mathbf{e} ;\Pi )
1015: \theta (\mathbf{V}_j +\sum_{k=1}^N m_k \mathbf{V}_k +\mathbf{e} ;\Pi )}
1016: {\theta ( \sum_{k=1}^N m_k \mathbf{V}_k + \mathbf{e} ;\Pi) \,
1017: \theta (\mathbf{V}_i + \mathbf{V}_j +
1018: \sum_{k=1}^N m_k \mathbf{V}_k +\mathbf{e} ;\Pi )} \lambda_{ij}^{-1}, \quad i<j
1019: \end{equation}
1020: where
1021: \begin{equation}
1022: \lambda_{ij} =
1023: \exp \left( \int_{Q_\infty}^{Q_i} \!\!\! \omega_{\sigma(Q_j) Q_j} \right),
1024: \qquad i<j.
1025: \end{equation}
1026: The BQL (the discrete BKP) $\tau$-function within the above class of
1027: solutions reads
1028: \begin{equation} \label{eq:tau_alg}
1029: \tau(m) = \theta ( \sum_{k=1}^N m_k \mathbf{V}_k + \mathbf{e} ;\Pi)
1030: \prod_{i<j} \lambda_{ij}^{m_i m_j}.
1031: \end{equation}
1032: \end{Cor}
1033: \begin{proof}
1034: Expression \eqref{eq:f-ij-alg} for the potentials is a direct consequence of
1035: their algebro-geometric definition \eqref{eq:f-ij-alg-def} and of equation
1036: \eqref{eq:psi}. Then the formula \eqref{eq:tau_alg}
1037: for $\tau$-function follows easily from its
1038: definiton \eqref{eq:tau}.
1039: \end{proof}
1040: \begin{Rem}
1041: From general considerations of \cite{DGNS} we know that
1042: \begin{equation} \label{eq:lambda}
1043: \lambda_{ij} =
1044: \exp \left( \int_{Q_\infty}^{Q_i} \!\!\! \omega_{\sigma(Q_j) Q_j} \right) =
1045: -\exp \left( \int_{Q_\infty}^{Q_j} \!\!\! \omega_{\sigma(Q_i) Q_i} \right),
1046: \quad i<j,
1047: \end{equation}
1048: which reflects the second equality in \eqref{eq:f-ij-alg-def}.
1049: \end{Rem}
1050: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1051: \section{Transformations of the B-quadrilateral lattice}
1052: \label{sec:B-transf}
1053: Below we present the reduction of the vectorial fundamental transformation
1054: compatible with the B-quadrilateral lattice constraint.
1055: In literature
1056: \cite{NiSchief} there is known the direct vectorial
1057: Moutard transformation between solutions of the BQL linear problem
1058: \eqref{eq:BKP-lin} providing thus the corresponding transformation between
1059: solutions
1060: of the discrete BKP equation \eqref{eq:BKP-nlin}. Our goal will be to find the
1061: transition to the Pfaffian expressions of \cite{NiSchief} starting from the BQL
1062: reduction of the fundamental transformation.
1063: In describing this connection we follow the
1064: ideas of \cite{LiuManas-BKP}, where similar problem between the Grammian
1065: expressions for binary Darboux transformation of the KP hierarchy has
1066: been transformed, in the BKP reduction, into the Pfaffian form
1067: \cite{Hirota-BKP-Pf} (see also \cite{Hirota-BKP-Pf,TsujimotoHirota} for other
1068: aspects of the relation of Pfaffians with
1069: the BKP hierarchy and the discrete BKP
1070: equation).
1071:
1072: \subsection{The fundamental transformation of the QL}
1073: %
1074: Let us first recall some basic facts concerning the vectorial fundamental
1075: transformation of the quadrilateral lattice.
1076: Geometrically, the (scalar) fundamental transformation is the relation
1077: between two
1078: quadrilateral
1079: lattices $x$ and $\hat{x}$ such that for each direction $i$ the
1080: points $x$, $\hat{x}$, $x_{(i)}$ and $\hat{x}_{(i)}$ are coplanar.
1081:
1082: We
1083: present below the algebraic description of its vectorial extension (see
1084: \cite{MDS,TQL,MM} for details) in the affine formalism.
1085: Given the solution $\boldsymbol{Y}_i:\mathbb{Z}^N\to\mathbb{V}$,
1086: $\mathbb{V}$ being a
1087: linear space, of the linear system \eqref{eq:lin-X}, and given the solution
1088: $\boldsymbol{Y}^*_i:\mathbb{Z}^N\to\mathbb{V}^*$, $\mathbb{V}^*$ being the
1089: dual of $\mathbb{V}$, of the linear system \eqref{eq:lin-h}. These allow to
1090: construct the linear operator valued potential
1091: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*):
1092: \mathbb{Z}^N\to L(\mathbb{V})$,
1093: defined by
1094: \begin{equation} \label{eq:Omega-Y-Y}
1095: \Delta_i \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1096: \boldsymbol{Y}_i \otimes\boldsymbol{Y}^*_i,
1097: \qquad i = 1,\dots , N;
1098: \end{equation}
1099: similarly, one defines
1100: $\boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*):
1101: \mathbb{Z}^N\to L(\mathbb{V},\mathbb{R}^M)$ and
1102: $\boldsymbol{\Omega}(\boldsymbol{Y},h):
1103: \mathbb{Z}^N\to \mathbb{V}$ by
1104: \begin{align} \label{eq:Omega-X-Y}
1105: \Delta_i \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*) & =
1106: \boldsymbol{X}_i \otimes\boldsymbol{Y}^*_i, \\
1107: \Delta_i\boldsymbol{\Omega}(\boldsymbol{Y},h) & =
1108: \boldsymbol{Y}_i \otimes h_i.
1109: \end{align}
1110: If $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ is invertible then
1111: the (vectorial) fundamental transform of the lattice $\bx$ is given by
1112: \begin{equation} \label{eq:fund-vect}
1113: \hat{\bx} = \bx -
1114: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*)
1115: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1116: \boldsymbol{\Omega}(\boldsymbol{Y},h).
1117: \end{equation}
1118: The corresponding transformation of the $\tau$-function $\tilde\tau$ of the quadrilateral
1119: lattice reads
1120: \begin{equation} \label{eq:fund-vect-tau}
1121: \hat{\tilde\tau} = \tilde\tau \det
1122: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*).
1123: \end{equation}
1124: When $\dim\mathbb{V}=1$ we obtain the formula relating the quadrilateral
1125: lattice $\bx$ and its fundmental transform $\hat{\bx}$. The vectorial
1126: fundamental transformation can be considered as superposition of
1127: $\dim\mathbb{V}$ (scalar) fundamental transformations; on intermediate stages
1128: the rest of the transformation data should be suitably transformed as well.
1129: Such a description contains already the principle of permutability of such
1130: transformations, which follows from the following observation~\cite{TQL}.
1131: \begin{Lem}
1132: Assume the following splitting of the data of the vectorial fundamental
1133: transformation
1134: \begin{equation}
1135: \boldsymbol{Y}_i = \left( \begin{array}{c}
1136: \boldsymbol{Y}_i^a \\ \boldsymbol{Y}_i^b \end{array} \right),\qquad
1137: \boldsymbol{Y}_i^* = \left( \begin{array}{cc}
1138: \boldsymbol{Y}_{ai}^{*}\; , & \boldsymbol{Y}_{b i}^{*} \end{array} \right),
1139: \end{equation}
1140: associated with the partition $\mathbb{V} = \mathbb{V}_a \oplus \mathbb{V}_b$,
1141: which implies the following splitting of the potentials
1142: \begin{equation} \label{eq:split-fund-1}
1143: \boldsymbol{\Omega}(\boldsymbol{Y},h) = \left( \begin{array}{c}
1144: \boldsymbol{\Omega}(\boldsymbol{Y}^a,h) \\
1145: \boldsymbol{\Omega}(\boldsymbol{Y}^b,h) \end{array} \right), \qquad
1146: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) = \left( \begin{array}{cc}
1147: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_a^*) &
1148: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_b^*) \\
1149: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}_a^*) &
1150: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}_b^*)\end{array} \right),
1151: \end{equation}
1152: \begin{equation} \label{eq:split-fund-2}
1153: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*) = \left( \begin{array}{cc}
1154: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_a^*)\; , &
1155: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_b^*)\end{array} \right).
1156: \end{equation}
1157: Then the vectorial fundamental tansformation is equivalent to the following
1158: superposition of vectorial fundamental transformations:\\
1159: 1) Transformation $\bx\to\hat{\bx}^{\{a\}}$ with the potentials
1160: $\boldsymbol{\Omega}(\boldsymbol{Y}^a,h)$,
1161: $\boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_a^*)$,
1162: $\boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_a^*)$
1163: \begin{equation} \label{eq:fund-vect-a}
1164: \hat{\bx}^{\{a\}} = \bx -
1165: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1166: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1167: \boldsymbol{\Omega}(\boldsymbol{Y}^a,h).
1168: \end{equation}
1169: 2) Application on the result the vectorial fundamental transformation with the
1170: transformed potentials
1171: \begin{equation} \label{eq:fund-vect-a-b}
1172: \hat{\bx}^{\{a,b\}} = \hat{\bx}^{\{a\}} -
1173: \hat{\boldsymbol{\Omega}}(\boldsymbol{X},\boldsymbol{Y}^*_b)^{\{a\}}
1174: [\hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b)^{\{a\}}]^{-1}
1175: \hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,h)^{\{a\}},
1176: \end{equation}
1177: where
1178: \begin{align}
1179: \hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,h)^{\{a\}} & =
1180: \boldsymbol{\Omega}(\boldsymbol{Y}^b,h) -
1181: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1182: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1183: \boldsymbol{\Omega}(\boldsymbol{Y}^a,h),
1184: \\
1185: \hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b)^{\{a\}} & =
1186: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b) -
1187: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1188: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1189: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_b),
1190: \\
1191: \hat{\boldsymbol{\Omega}}(\boldsymbol{X},\boldsymbol{Y}^*_b)^{\{a\}} & =
1192: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_b) -
1193: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1194: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1195: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_b).
1196: \label{eq:fund-vect-potentials-slit}
1197: \end{align}
1198: \end{Lem}
1199: \begin{Cor}
1200: The normalized tangent vectors $\boldsymbol{X}_i$ and the Lam\'{e} coefficients
1201: $h_i$ are transformed, at the intermediate step, according
1202: to formulas
1203: \begin{align}
1204: \hat{\boldsymbol{X}}_i^{\{a\}} & = \boldsymbol{X}_i -
1205: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1206: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1207: \boldsymbol{Y}^a_i,
1208: \\
1209: \hat{h}_i^{\{a\}} & = h_i - \boldsymbol{Y}^*_{i a}
1210: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1211: \boldsymbol{\Omega}(\boldsymbol{Y}^a,h),
1212: \end{align}
1213: which also give the corresponding transforms of the second set of
1214: transformation data
1215: $\boldsymbol{Y}^b$ and $\boldsymbol{Y}^*_b$
1216: \begin{align}
1217: \hat{\boldsymbol{Y}}_i^{b\{a\}} & = \boldsymbol{Y}_i^b -
1218: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1219: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1220: \boldsymbol{Y}^a_i,
1221: \\
1222: \hat{\boldsymbol{Y}}_{i b}^{*\{a\}} & = \boldsymbol{Y}_{i b}^* -
1223: \boldsymbol{Y}^*_{i a}
1224: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1225: \boldsymbol{\Omega}(\boldsymbol{Y}^a, \boldsymbol{Y}_{b}^*);
1226: \end{align}
1227: which agree with the transformation rules
1228: \eqref{eq:fund-vect-potentials-slit} for the potentials, i.e.,
1229: \begin{align*}
1230: \hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,h)^{\{a\}} &=
1231: \boldsymbol{\Omega}(\hat{\boldsymbol{Y}}^{b\{a\}},\hat{h}^{\{a\}}), \\
1232: \hat{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b)^{\{a\}} &=
1233: \boldsymbol{\Omega}(\hat{\boldsymbol{Y}}^{b\{a\}},\hat{\boldsymbol{Y}}_b^{*\{a\}}),
1234: \\
1235: \hat{\boldsymbol{\Omega}}(\boldsymbol{X},\boldsymbol{Y}^*_b)^{\{a\}} & =
1236: \boldsymbol{\Omega}(\hat{\boldsymbol{X}}^{\{a\}},\hat{\boldsymbol{Y}}_b^{*\{a\}}).
1237: \end{align*}
1238: \end{Cor}
1239: \begin{Rem}
1240: The same result $\hat\bx = \hat{\bx}^{\{a,b\}}=\hat{\bx}^{\{b,a\}}$
1241: is obtained exchanging the order of transformations, exchanging also the indices
1242: $a$ and $b$ in formulas
1243: \eqref{eq:fund-vect-a}-\eqref{eq:fund-vect-potentials-slit}.
1244: \end{Rem}
1245:
1246: \begin{Rem}
1247: If we denote by $\hat{\bx}^{\{1,2\}}$ the
1248: quadrilateral lattice obtained by superposition of two (scalar) fundamental
1249: transforms from $\bx$ to $\hat{\bx}^{\{1\}}$ and $\hat{\bx}^{\{2\}}$, then the points
1250: $\bx$, $\hat{\bx}^{\{1\}}$, $\hat{\bx}^{\{2\}}$ and $\hat{\bx}^{\{1,2\}}$ are coplanar
1251: again, i.e., the fundamental transformations reproduce the planarity constraint
1252: responsible for integrability of the quadrilateral lattice.
1253: \end{Rem}
1254:
1255: \subsection{The BQL (Moutard) reduction of the fundamental transformation}
1256: In this section we describe restrictions on the data of the fundamental
1257: transformation in order to preserve the reduction from QL to BQL.
1258: As usuall (see, for example \cite{TQL,q-red,MM}) the reduction of the
1259: fundamental transformation for the special quadrilateral lattices
1260: mimics the geometric properties of the lattice. Because the basic geometric
1261: property of the (scalar) fundamental transformation can be interpreted as
1262: construction of a "new level" of the quadrilateral lattice, then it is natural
1263: to define the reduced transformation in a similar spirit. Our definition of
1264: \emph{the BQL reducion of the fundamental transformation} is therefore based on
1265: the following observation.
1266: \begin{Lem} \label{lem:BQL-fund}
1267: Given B-quadrilateral lattice $x:\ZZ^N\to\PP^M$ and its fundamental transform
1268: $\hat{x}$ constructed under additional assumption that for any point
1269: $x$ of the lattice and any pair $i,j$ of different directions, the four points
1270: $x$, $x_{(ij)}$, $\hat{x}_{(i)}$ and $\hat{x}_{(j)}$ are coplanar. Then the
1271: lattice $\hat{x}:\ZZ^N\to\PP^M$ is B-quadrilateral lattice as well.
1272: \end{Lem}
1273: \begin{proof}
1274: The result is equivalent to the $4$-dimensional consistency of the BQL lattice.
1275: Indeed, in Proposition~\ref{lem:4D-consist-BQL} let the forth direction be
1276: identified withe the transformation direction (the three first directions are
1277: the lattice directions $i$, $j$, $k$). Then the implication
1278: $x_4 \in \langle x_{124}, x_{134}, x_{234} \rangle$ is rewritten in the form
1279: $\hat{x} \in \langle \hat{x}_{(ij)}, \hat{x}_{(ik)}, \hat{x}_{(jk)} \rangle$.
1280: \end{proof}
1281: \begin{Def}
1282: The fundamental transform $\hat{x}$ of a B-quadrilateral lattice $x:\ZZ^N\to\PP^M$
1283: constructed under additional assumption that for any point
1284: $x$ of the lattice and any pair $i,j$ of different directions, the four points
1285: $x$, $x_{(ij)}$, $\hat{x}_{(i)}$ and $\hat{x}_{(j)}$ are coplanar is called
1286: \emph{the BQL reduction} of the fundamental transformation of $x$.
1287: \end{Def}
1288:
1289: On the algebraic level, the Darboux-type transformation of the solutions
1290: of the linear problem \eqref{eq:BKP-lin} was introduced and studied by Nimmo
1291: and Schief in \cite{NiSchief} as \emph{discretization of the Moutard
1292: transformation}. We will derive their results from the general theory of
1293: transformations of the quadrilateral lattice.
1294: %
1295: \begin{Lem} \label{lem:Y-Y*-BQL}
1296: Given a scalar solution $Y_i$ of the linear problem \eqref{eq:lin-X} with the
1297: rotation coefficients restricted by the BQL
1298: reduction \eqref{eq:beta-BQL-ij}-\eqref{eq:beta-BQL-ji},
1299: denote by $\theta = \Omega(Y,h)$ the corresponding potential, where the
1300: Lam\'{e} coefficients are given by equation \eqref{eq:h-tau}. Then
1301: the functions
1302: \begin{equation} \label{eq:Yi*-BQL}
1303: Y_i^* = (-1)^{\sum_{k<i}m_k}\frac{\tau}{\tau_{(i)}} (\theta + \theta_{(i)})
1304: \end{equation}
1305: are solutions of the adjoint linear problem \eqref{eq:lin-h}
1306: in the BQL reduction, and
1307: the function $\theta^2$ can be taken as the corresponding potential
1308: $\Omega(Y,Y^*)$
1309: \begin{equation} \label{eq:Om-BQL}
1310: \theta^2 = \Omega(Y,Y^*).
1311: \end{equation}
1312: \end{Lem}
1313: \begin{proof}
1314: By direct calculation one verifies that the functions defined in
1315: \eqref{eq:Yi*-BQL} satisfy the reduced system \eqref{eq:lin-h}. Similarly one
1316: checks validity of equation \eqref{eq:Om-BQL}.
1317: \end{proof}
1318: \begin{Rem}
1319: Notice that equation \eqref{eq:Om-BQL} implies, under assumptions of
1320: Lemma~\ref{lem:Y-Y*-BQL}, the form \eqref{eq:Yi*-BQL} of the solution of the
1321: adjoint linear problem.
1322: \end{Rem}
1323: \begin{Prop} \label{prop:BQL-fund}
1324: Given BQL lattice $x$ with homogeneous representation $\bx$ in the gauge of the
1325: linear problem \eqref{eq:BKP-lin}, then the transform of $\bx$
1326: constructed using equation~\eqref{eq:fund-vect}
1327: with the data described in Lemma~\ref{lem:Y-Y*-BQL} data satisfies
1328: the conditions of the BQL reduction.
1329: \end{Prop}
1330: \begin{proof}
1331: The fundamental transform of
1332: $\bx$ constructed with such a data reads
1333: \begin{equation} \label{eq:Omega-XY-BQL}
1334: \hat\bx = \bx - \boldsymbol{\Omega}(\bX,Y^*)/\theta.
1335: \end{equation}
1336: Equation \eqref{eq:Omega-X-Y}, with $\boldsymbol{\Omega}(\bX,Y^*)$ given above,
1337: can be rewritten then in the following
1338: form
1339: \begin{equation} \label{eq:Moutard-transf}
1340: \hat\bx_{(i)} - \bx = \frac{\theta}{\theta_{(i)}}\left( \hat\bx -
1341: \bx_{(i)}\right),
1342: \end{equation}
1343: which, together with the linear problem \eqref{eq:BKP-lin}, implies the
1344: linear relation
1345: \begin{equation}
1346: \frac{\theta}{\theta_{(i)}}\left(\hat\bx_{(i)} - \bx \right) -
1347: \frac{\theta}{\theta_{(j)}}\left(\hat\bx_{(j)} - \bx \right) +
1348: \frac{1}{f^{ij}}\left(\bx_{(ij)} - \bx \right) =0, \quad i< j,
1349: \end{equation}
1350: between the homogeneous coordinates of the points $x$, $x_{(ij)}$
1351: of the lattice and the points $\hat{x}_{(i)}$ and $\hat{x}_{(j)}$
1352: of its fundamental transform, which is the
1353: algebraic expression of their coplanarity.
1354: \end{proof}
1355: In the approach of \cite{NiSchief} the Moutard transform $\hat\bx$ of $\bx$ was
1356: defined in terms of the system \eqref{eq:Moutard-transf}. Then $\hat\bx$
1357: satisfies new Moutard equations \eqref{eq:BKP-lin} with new potential
1358: \begin{equation}
1359: \hat{f}^{ij} = f^{ij}\frac{\theta_{(i)}\theta_{(j)}}{\theta\theta_{(ij)}},
1360: \qquad i<j,
1361: \end{equation}
1362: and new $\tau$-function
1363: \begin{equation} \label{eq:transf-Mout-tau}
1364: \hat\tau = \theta\tau.
1365: \end{equation}
1366: Another important ingredient of \cite{NiSchief} was the existence of the
1367: potential $S(\theta|\bx) = \theta \hat\bx $ which satisfies
1368: the system
1369: \begin{equation}
1370: \Delta_{i}S(\theta|\bx) = \theta_{(i)} \bx - \theta \bx_{(i)}.
1371: \end{equation}
1372:
1373: We have shown that the algebraic reduction, described in
1374: Lemma~\ref{lem:Y-Y*-BQL}, of the data of the fundamental transformation
1375: can be interpreted as a BQL reduction of the transformation.
1376: We close this Section by showing that the above algebraic description
1377: holds generally.
1378: \begin{Prop}
1379: Any BQL-reduction of the fundamental transformation can be
1380: algebraically described as
1381: in Proposition~\ref{prop:BQL-fund}.
1382: \end{Prop}
1383: \begin{proof}
1384: We will follow the
1385: reasoning of \cite{BobSur} used to the same linear problem \eqref{eq:BKP-lin}
1386: but in different geometric context.
1387: Because the BQL-reduced
1388: fundamental transformation can be considered as construction of the new level of
1389: the B-quadrilateral lattice, its algebraic representation
1390: should be (in appropriate gauge) in the form of the BQL linear problem
1391: \eqref{eq:BKP-lin}
1392: \begin{equation} \label{eq:BQL-transf-lin}
1393: \hat\bx_{(i)} - \bx = f^{0i}(\hat\bx - \bx_{(i)}),
1394: \end{equation}
1395: where we can label the transformation direction by the index $0$. The
1396: compatibility of system \eqref{eq:BQL-transf-lin} gives
1397: the following equations (compare with
1398: \eqref{eq:nonl-BQL-f})
1399: \begin{equation}
1400: f^{0i}_{(j)}f^{0j} = f^{0j}_{(i)}f^{0i}, \qquad
1401: 1 - f^{0j}_{(i)}(f^{0i} + f^{ij}) = - f^{0i}_{(j)}f^{ij}.
1402: \end{equation}
1403: First of them implies the existence of a potential $\theta$ such that
1404: \begin{equation*}
1405: f^{0i} = \frac{\theta}{\theta_{(i)}},
1406: \end{equation*}
1407: thus equations \eqref{eq:Moutard-transf}. The second equation rewritten in terms
1408: of the potential implies that
1409: $\theta$ satisfies linear problem \eqref{eq:BKP-lin}, i.e. $\theta=\Omega(Y,h)$,
1410: where $h$ given by \eqref{eq:h-tau} and $Y_i$ is a solution
1411: of the linear problem
1412: \eqref{eq:lin-X} with the
1413: rotation coefficients restricted by the BQL
1414: reduction \eqref{eq:beta-BQL-ij}-\eqref{eq:beta-BQL-ji}. By equation
1415: \eqref{eq:Yi*-BQL} we define the corresponding solution of the adjoint linear
1416: problem. Finally, direct calculation with the help of the Moutard transformation
1417: formulas \eqref{eq:Moutard-transf} show that the potential
1418: \begin{equation*}
1419: \boldsymbol{\Omega}(\bX,Y^*) = \theta (\bx - \hat\bx),
1420: \end{equation*}
1421: (compare with equation \eqref{eq:Omega-XY-BQL}) does satisfy
1422: equation \eqref{eq:Omega-X-Y}, thus $\Omega(Y,Y^*)$ is of the form given in
1423: equation \eqref{eq:Om-BQL}.
1424: \end{proof}
1425:
1426: \subsection{The BQL reduction of the vectorial fundamental transformation}
1427: In this Section we propose the restrictions on the data of the vectorial
1428: fundamental transformation, which are compatible with the BQL reduction.
1429: \begin{Prop} \label{prop:vect-fund-red-BQL}
1430: Given solution $\boldsymbol{Y}_i:\ZZ^N\to\mathbb{V}$ of the linear problem
1431: \eqref{eq:lin-X} corresponding to the BQL linear problem \eqref{eq:BKP-lin}
1432: satisfied by the homogeneous coordinates $\bx$ of the BQL lattice
1433: $x:\ZZ^N\to\PP^M$.
1434: Denote by $\boldsymbol{\Theta} = \boldsymbol{\Omega}(\boldsymbol{Y},h)$ the
1435: corresponding potential, which is also new
1436: vectorial solution of the BQL linear problem \eqref{eq:BKP-lin}.\\
1437: 1) Then
1438: \begin{equation} \label{eq:BQL-Y*}
1439: \boldsymbol{Y}^*_i = (-1)^{\sum_{k<i}m_k}\frac{\tau}{\tau_{(i)}}
1440: (\boldsymbol{\Theta}^{\mathrm{t}} + \boldsymbol{\Theta}_{(i)}^{\mathrm{t}})
1441: \end{equation}
1442: provides a vectorial
1443: solution of the adjoint linear problem, and the corresponding
1444: potential
1445: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ allows for the following
1446: constraint
1447: \begin{equation} \label{eq:vect-Moutard-constr}
1448: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) +
1449: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{\mathrm{t}} =
1450: 2 \boldsymbol{\Theta}\otimes \boldsymbol{\Theta}^{\mathrm{t}}.
1451: \end{equation}
1452: 2) The fundamental vectorial transform $\hat\bx$ of $\bx$, given by
1453: \eqref{eq:fund-vect} with the potentials $\boldsymbol{\Omega}$ restricted as
1454: above can be considered as the superposition of $\dim\mathbb{V}$ (scalar)
1455: discrete BQL reduced fundamental transforms.
1456: \end{Prop}
1457: \begin{proof}
1458: The point 1) can be checked by direct calculation.
1459: To prove the point 2) notice that when $\dim\mathbb{V}=1$ we obtain the
1460: BQL reduction of the fundamental transformation in the
1461: setting of Proposition~\ref{prop:BQL-fund}. For $\dim\mathbb{V}>1$
1462: the statement follows from the standard reasoning applied to superposition of
1463: two reduced vectorial fundamental transformations (compare with \cite{TQL,q-red}).
1464:
1465: Assume the splitting
1466: $\mathbb{V}=\mathbb{V}_a\oplus\mathbb{V}_b$ of the vectorial space $\mathbb{V}$,
1467: and the induced splitting of the basic data $\boldsymbol{Y}_i$ of the
1468: transformation. Then we have also (in shorthand notation, compare equations
1469: \eqref{eq:split-fund-1}-\eqref{eq:split-fund-2})
1470: \begin{equation}
1471: \boldsymbol{\Theta} = \left( \begin{array}{c}
1472: \boldsymbol{\Theta}^a \\ \boldsymbol{\Theta}^b \end{array} \right),
1473: \qquad
1474: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1475: \left( \begin{array}{cc}
1476: \boldsymbol{\Omega}^a_a & \boldsymbol{\Omega}^a_b \\
1477: \boldsymbol{\Omega}^b_a & \boldsymbol{\Omega}^b_b
1478: \end{array} \right),
1479: \end{equation}
1480: and the constraint \eqref{eq:vect-Moutard-constr} reads
1481: \begin{equation} \label{eq:vect-Moutard-constr-split}
1482: \left( \begin{array}{cc}
1483: \boldsymbol{\Omega}^a_a & \boldsymbol{\Omega}^a_b \\
1484: \boldsymbol{\Omega}^b_a & \boldsymbol{\Omega}^b_b
1485: \end{array} \right)
1486: +
1487: \left( \begin{array}{cc}
1488: \boldsymbol{\Omega}^{a \mathrm{t}}_a & \boldsymbol{\Omega}^{b \mathrm{t}}_a \\
1489: \boldsymbol{\Omega}^{a \mathrm{t}}_b & \boldsymbol{\Omega}^{b \mathrm{t}}_b
1490: \end{array} \right)
1491: = 2
1492: \left( \begin{array}{cc}
1493: \boldsymbol{\Theta}^a \otimes \boldsymbol{\Theta}^{a \mathrm{t}}&
1494: \boldsymbol{\Theta}^a \otimes \boldsymbol{\Theta}^{b \mathrm{t}} \\
1495: \boldsymbol{\Theta}^b \otimes \boldsymbol{\Theta}^{a \mathrm{t}} &
1496: \boldsymbol{\Theta}^b \otimes \boldsymbol{\Theta}^{b \mathrm{t}}
1497: \end{array} \right).
1498: \end{equation}
1499: By straightforward algebra, using equations
1500: \eqref{eq:vect-Moutard-constr-split}, one checks that the transformed potentials
1501: (compare equations \eqref{eq:fund-vect-potentials-slit})
1502: \begin{align}
1503: \boldsymbol{\Omega}^{b \{ a\}}_b & = \boldsymbol{\Omega}^b_b -
1504: \boldsymbol{\Omega}^b_a [\boldsymbol{\Omega}^a_a]^{-1} \boldsymbol{\Omega}^a_b,
1505: \\
1506: \boldsymbol{\Theta}^{b \{ a\}} & = \boldsymbol{\Theta}^b -
1507: \boldsymbol{\Omega}^b_a [\boldsymbol{\Omega}^a_a]^{-1} \boldsymbol{\Theta}^a,
1508: \end{align}
1509: satisfy the BQL constraint \eqref{eq:vect-Moutard-constr} as well, i.e.,
1510: \begin{equation}
1511: \boldsymbol{\Omega}^{b \{ a\}}_b +
1512: \boldsymbol{\Omega}^{b \{ a\}\, \mathrm{t} }_b =
1513: 2 \, \boldsymbol{\Theta}^{b \{ a\}} \otimes
1514: \boldsymbol{\Theta}^{b \{ a\}\, \mathrm{t} },
1515: \end{equation}
1516: which concludes the proof.
1517: \end{proof}
1518: \begin{Rem}
1519: Because the BQL-reduced fundamental transformation can be cosidered as
1520: construction of new levels of the B-quadrilateral lattice, then
1521: if we denote by $\hat{x}^{\{1,2\}}$ the
1522: B-quadrilateral lattice obtained by superposition of two (scalar) such
1523: transforms from $x$ to $\hat{x}^{\{1\}}$ and $\hat{x}^{\{2\}}$, then
1524: for each direction $i$ of the lattice the points
1525: $x$, $\hat{x}^{\{1\}}_{(i)}$, $\hat{x}^{\{2\}}_{(i)}$ and
1526: $\hat{x}^{\{1,2\}}$ are coplanar as well as the points
1527: $x_{(i)}$, $\hat{x}^{\{1\}}$, $\hat{x}^{\{2\}}$ and $\hat{x}^{\{1,2\}}_{(i)}$.
1528: Similarly, if we consider superpositions of three (scalar) transforms of the
1529: B-quadrilateral lattice $x$ then
1530: the points $x$, $\hat{x}^{\{1,2\}}$, $\hat{x}^{\{1,3\}}$ and
1531: $\hat{x}^{\{2,3\}}$ are coplanar as well as the points
1532: $\hat{x}^{\{1\}}$, $\hat{x}^{\{2\}}$, $\hat{x}^{\{3\}}$ and
1533: $\hat{x}^{\{1,2,3\}}$.
1534: \end{Rem}
1535:
1536: \subsection{The Pfaffian form of the transformation}
1537: \label{sec:fund-Pf}
1538: Finally, we are going to show that the Pfaffian formulas of the
1539: vectorial discrete Moutard transformation obtained in \cite{NiSchief}
1540: can be derived from
1541: the corresponding formulas of the fundamental transformation subjected to
1542: the BQL
1543: reduction.
1544:
1545: Denote by $S(\boldsymbol{\Theta} | \boldsymbol{\Theta})$ the antisymmetric part of
1546: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ then equations
1547: \eqref{eq:Omega-Y-Y}, \eqref{eq:BQL-Y*} and \eqref{eq:vect-Moutard-constr} imply
1548: \begin{equation}
1549: \Delta_i S(\boldsymbol{\Theta} | \boldsymbol{\Theta}) =
1550: \boldsymbol{\Theta}_{(i)}\otimes \boldsymbol{\Theta}^{\mathrm{t}} -
1551: \boldsymbol{\Theta}\otimes \boldsymbol{\Theta}^{\mathrm{t}}_{(i)}.
1552: \end{equation}
1553: \begin{Lem}
1554: For $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ and
1555: $S(\boldsymbol{\Theta} |\boldsymbol{\Theta})$ as above we have
1556: \begin{equation} \label{eq:det-Omega}
1557: \det\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1558: \det S(\boldsymbol{\Theta} | \boldsymbol{\Theta}) + \left| \begin{array}{cc}
1559: 0 & - \boldsymbol{\Theta}^{\mathrm{t}} \\
1560: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta})
1561: \end{array} \right|.
1562: \end{equation}
1563: \end{Lem}
1564: \begin{proof}
1565: Notice that the $j$th column $\Omega_j$ of
1566: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ is of the form
1567: \begin{equation} \label{eq:Omega-j}
1568: \Omega_j = \theta^j \boldsymbol{\Theta} + S_j,
1569: \end{equation}
1570: where $\theta^j$ is the $j$th component of $\boldsymbol{\Theta}$, and
1571: $S_j$ is the $j$th column of $S(\boldsymbol{\Theta} | \boldsymbol{\Theta})$.
1572: Then the basic properties of determinants imply that
1573: \begin{equation} \label{eq:det-Omega-2}
1574: \det \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1575: \det S(\boldsymbol{\Theta} | \boldsymbol{\Theta}) + \sum_{j=1}^{\dim\mathbb{V}}
1576: \theta^j S(j),
1577: \end{equation}
1578: where by $S(j)$ we denote the matrix $S(\boldsymbol{\Theta} |
1579: \boldsymbol{\Theta})$ with $j$th column replaced by $\boldsymbol{\Theta}$. The
1580: second summand in \eqref{eq:det-Omega-2} is the Laplace expansion of that in
1581: \eqref{eq:det-Omega}.
1582: \end{proof}
1583:
1584: The standard properties of determinants of antisymmetric matrices (see
1585: Appendix~\ref{app:Pf}) imply the following result derived in \cite{NiSchief}
1586: directly on the level of vectorial Moutard transformation.
1587: \begin{Cor}
1588: The transformation formula \eqref{eq:fund-vect-tau} of the QL $\tau$-function
1589: and the relation \eqref{eq:tau-QL-BQL} between both $\tau$-functions imply
1590: the following transformation formula for the BQL $\tau$-function
1591: \begin{equation}
1592: \hat\tau = \begin{cases}
1593: \tau \: \mathrm{Pf} \: S(\boldsymbol{\Theta} | \boldsymbol{\Theta}) , &
1594: \dim\mathbb{V} \; \text{even} \\
1595: \tau \: \mathrm{Pf} \left( \begin{array}{cc}
1596: 0 & - \boldsymbol{\Theta}^{\mathrm{t}} , \\
1597: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta})
1598: \end{array} \right), \qquad & \dim\mathbb{V} \; \text{odd}.
1599: \end{cases}
1600: \end{equation}
1601: \end{Cor}
1602: \begin{Rem}
1603: Notice \cite{NiSchief} that
1604: \begin{equation}
1605: \left( \begin{array}{cc}
1606: 0 & - \boldsymbol{\Theta}^{\mathrm{t}} , \\
1607: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta})
1608: \end{array} \right) = S(\boldsymbol{\tilde\Theta} | \boldsymbol{\tilde\Theta}),
1609: \qquad \text{where} \qquad \boldsymbol{\tilde\Theta} = \left( \begin{array}{c}
1610: 1 \\ \boldsymbol{\Theta} \end{array} \right),
1611: \end{equation}
1612: which allows to define
1613: \begin{equation}
1614: \mathcal{P}(\boldsymbol{\Theta} ) = \begin{cases}
1615: \mathrm{Pf} \: S(\boldsymbol{\Theta} | \boldsymbol{\Theta}), &
1616: \dim\mathbb{V} \; \text{even} ,\\
1617: \mathrm{Pf}\: S(\boldsymbol{\tilde\Theta} | \boldsymbol{\tilde\Theta}),
1618: \qquad & \dim\mathbb{V} \; \text{odd},
1619: \end{cases}
1620: \end{equation}
1621: and gives
1622: \begin{equation}
1623: \hat\tau = \tau \:\mathcal{P}(\boldsymbol{\Theta} ) .
1624: \end{equation}
1625: \end{Rem}
1626:
1627: Finally, we will connect the formula of the vectorial fundamental
1628: transformation \eqref{eq:fund-vect} in the BQL reduction
1629: with the Pffafian form of the vectorial Moutard transformation \cite{NiSchief}.
1630: \begin{Cor}
1631: The homogeneous coordinates (in the gauge of the linear problem
1632: \eqref{eq:BKP-lin}) of the BQL lattice $\hat{x}$ obtained from the BQL lattice
1633: $x$ via vectorial transform with the solution $\boldsymbol{\Theta}$ of the
1634: linear problem \eqref{eq:BKP-lin} are given by
1635: \begin{equation}
1636: \hat{x}^i =
1637: \frac{\mathcal{P}(\boldsymbol{\Theta},x^i)}{\mathcal{P}(\boldsymbol{\Theta})},
1638: \end{equation}
1639: where
1640: \begin{equation}
1641: \mathcal{P}(\boldsymbol{\Theta},x^i) =
1642: \mathcal{P}\left(\left(\begin{array}{c}\boldsymbol{\Theta}\\x^i \end{array}
1643: \right)\right).
1644: \end{equation}
1645: \end{Cor}
1646: \begin{proof}
1647: We will work using assumptions and notation of
1648: Proposition~\ref{prop:vect-fund-red-BQL}.
1649: Let us define
1650: \begin{equation}
1651: S(\bx | \boldsymbol{\Theta}) =
1652: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*) -
1653: \bx\otimes \boldsymbol{\Theta}^{\mathrm{t}} ,
1654: \end{equation}
1655: then, due to equations \eqref{eq:x-hX}, \eqref{eq:Omega-X-Y} and
1656: \eqref{eq:BQL-Y*} we have
1657: \begin{equation}
1658: \Delta_i S(\bx | \boldsymbol{\Theta}) =
1659: \bx_{(i)}\otimes \boldsymbol{\Theta}^{\mathrm{t}} -
1660: \bx\otimes \boldsymbol{\Theta}^{\mathrm{t}}_{(i)}.
1661: \end{equation}
1662: By the Cramer rule and equation \eqref{eq:Omega-j}, formula \eqref{eq:fund-vect}
1663: in the considered reduction case can be brought to the form
1664: \begin{equation} \label{eq:Mout-vect}
1665: \hat\bx = \bx - \left( \bx\otimes \boldsymbol{\Theta}^{\mathrm{t}} +
1666: S(\bx | \boldsymbol{\Theta}) \right)\frac{1}{\det
1667: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) }\left( \begin{array}{c}
1668: \det S(1) \\ \vdots \\ \det S(\dim \mathbb{V}) \end{array} \right),
1669: \end{equation}
1670: moreover we have
1671: \begin{equation} \label{eq:theta-Sj}
1672: \boldsymbol{\Theta}^{\mathrm{t}} \left( \begin{array}{c}
1673: \det S(1) \\ \vdots \\ \det S(\dim \mathbb{V}) \end{array} \right) =
1674: \left| \begin{array}{cc}
1675: 0 & - \boldsymbol{\Theta}^{\mathrm{t}} \\
1676: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta})
1677: \end{array} \right|.
1678: \end{equation}
1679:
1680: Our further analysis splits in the cases of $\dim\mathbb{V}$ being even or odd.
1681: In the first case the right hand side of
1682: equation \eqref{eq:theta-Sj} vanishes giving
1683: \begin{equation}
1684: \hat\bx = \bx -
1685: \frac{S(\bx | \boldsymbol{\Theta})}
1686: {\mathrm{Pf} S(\boldsymbol{\Theta}|\boldsymbol{\Theta}) }
1687: \left( \begin{array}{c}
1688: \mathrm{Pf} S[1] \\ \vdots \\ \mathrm{Pf} S [\dim\mathbb{V}] \end{array} \right),
1689: \end{equation}
1690: where we used equation \eqref{eq:det-Omega} and the Pfaffian analogue
1691: \eqref{eq:Pfaff-Cramer} of the Cramer rule for
1692: for solutions of the equation
1693: $S(\boldsymbol{\Theta}|\boldsymbol{\Theta}) \boldsymbol{y} =
1694: \boldsymbol{\Theta}$. Then the expansion rule for Pfaffians
1695: \eqref{eq:Laplace-Pfaff} implies that the
1696: $i$th coordinates of the B-quadrilateral lattice $\bx$ and its transform
1697: $\hat\bx$ can be put in the form
1698: \begin{equation}
1699: \hat{x}^i = \frac{1}{\mathrm{Pf} S(\boldsymbol{\Theta}|\boldsymbol{\Theta})}
1700: \mathrm{Pf}\left( \begin{array}{ccc}
1701: 0 & x^i & S(x^i| \boldsymbol{\Theta} ) \\
1702: -x^i & 0 & -\boldsymbol{\Theta}^{\mathrm{t}} \\
1703: -S(x^i | \boldsymbol{\Theta} )^{\mathrm{t}} & \boldsymbol{\Theta} &
1704: S(\boldsymbol{\Theta} | \boldsymbol{\Theta} )
1705: \end{array} \right) =
1706: \frac{\mathcal{P}(\boldsymbol{\Theta},x^i)}{\mathcal{P}(\boldsymbol{\Theta})}.
1707: \end{equation}
1708:
1709: For $\dim\mathbb{V}$ odd, by equations \eqref{eq:det-Omega} and
1710: \eqref{eq:theta-Sj}, equation \eqref{eq:Mout-vect} reduces to
1711: \begin{equation}
1712: \hat{\boldsymbol{x}} = - \left| \begin{array}{cc}
1713: 0 & - \boldsymbol{\Theta}^{\mathrm{t}} \\
1714: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta})
1715: \end{array} \right|^{-1} S(\bx|\boldsymbol{\Theta})
1716: \left( \begin{array}{c}
1717: \det S(1) \\ \vdots \\ \det S(\dim \mathbb{V}) \end{array} \right).
1718: \end{equation}
1719: Expanding $\det S(j)$ with respect to its $j$th column, and using Pfaffian
1720: expressions \eqref{eq:minors-odd-Pf} for the minors of
1721: $S(\boldsymbol{\Theta} | \boldsymbol{\Theta} ) $ we obtain
1722: \begin{equation}
1723: \hat{x}^i = \frac{\mathrm{Pf}\left( \begin{array}{cc}
1724: 0 & S(x^i| \boldsymbol{\Theta} ) \\
1725: -S(x^i | \boldsymbol{\Theta} )^{\mathrm{t}} &
1726: S(\boldsymbol{\Theta} | \boldsymbol{\Theta} )
1727: \end{array} \right) }
1728: {\mathrm{Pf}\left( \begin{array}{cc}
1729: 0 & -\boldsymbol{\Theta}^{\mathrm{t}} \\
1730: \boldsymbol{\Theta} & S(\boldsymbol{\Theta} | \boldsymbol{\Theta} )
1731: \end{array} \right) } =
1732: \frac{\mathcal{P}(\boldsymbol{\Theta},x^i)}{\mathcal{P}(\boldsymbol{\Theta})},
1733: \end{equation}
1734: which concludes the proof.
1735: \end{proof}
1736:
1737: %
1738: \section{Conclusion and remarks}
1739: We presented new geometric interpretation of the discrete BKP equation within
1740: the theory of quadrilateral lattices. This new integrable
1741: lattice should be considered, together with the
1742: symmetric lattice \cite{DS-sym} and the quadrilateral
1743: lattices subject to quadratic
1744: constraints \cite{q-red}, as one of basic reductions of the quadrilateral
1745: lattice. In the forthcomong paper \cite{AD-isoth} we show, for example, that the
1746: discrete isothermic surfaces \cite{BP2} are lattices subjected simultaneously to
1747: the BQL and quadratic (in this case the quadric is the M\"{o}bius sphere)
1748: reductions.
1749:
1750: As in the case of the Hirota (the discrete KP) equation, also the Miwa (the
1751: discrete BKP)
1752: equation can be considered in the finite fields (or the finite geometry)
1753: setting. In paricular, the main algebro-geometric way of reasoning (see
1754: \cite{DBK,BD} for the former
1755: discrete KP case) leading to the Prym varieties should be also transferable
1756: for fields of
1757: characteristic different from two (see \cite{Mumford} for general theory of
1758: Prym varieties).
1759:
1760: The explicit Prym-theta functional formalae (it is enough to consider the case
1761: $N=2$) for the wave function and potentials of
1762: the discrete Moutard equation can be also used to provide
1763: characterization of the Prym varieties among all principally polarized abelian
1764: varieties (the
1765: Prym--Schottky problem) in the spirit of \cite{Krichever-Prym}.
1766:
1767: \section*{Acknowledgements}
1768: The author would like to thank Jaros{\l}aw Kosiorek and Andrzej
1769: Matra\'{s} for discussions on projective geometry.
1770:
1771: The main part of the paper was
1772: prepared during author's work at DFG Research Center MATHEON in
1773: Institut f\"{u}r Mathematik of the Technische Universit\"{a}t Berlin.
1774: The paper was supportet also in part by the Polish Ministry of
1775: Science and Higher Education research grant 1~P03B~017~28.
1776:
1777: \appendix
1778:
1779:
1780: \section{An alternative proof of the existence of the BQL gauge}
1781:
1782: The planarity condition of elementary
1783: quadrilaterals of QL
1784: can be expressed in terms of generic homogoneous representation
1785: as the following system of discrete Laplace equations \cite{MQL}
1786: \begin{equation} \label{eq:Laplace}
1787: \bx_{(ij)} = a^{ij}\bx_{(i)} - a^{ji}\bx_{(j)} + c^{ij}\bx,
1788: \quad 1\leq i < j \leq N
1789: \end{equation}
1790: whose compatibility are equations
1791: \begin{align} \label{eq:MQL1}
1792: &a^{ij}_{(k)}c^{ik} - a^{ji}_{(k)}c^{jk} =
1793: a^{ik}_{(j)}c^{ij} - a^{ki}_{(j)}c^{jk} =
1794: a^{jk}_{(i)}c^{ij} - a^{kj}_{(i)}c^{ik}, \\
1795: \label{eq:MQL2}
1796: &a^{ij}_{(k)}a^{ik} = a^{ik}_{(j)}a^{ij} =
1797: c^{jk}_{(i)} + a^{jk}_{(i)}a^{ij} - a^{kj}_{(i)}a^{ik} \\
1798: \label{eq:MQL3}
1799: &a^{ji}_{(k)}a^{jk} = a^{jk}_{(i)}a^{ji} =
1800: - c^{ik}_{(j)} + a^{ik}_{(j)}a^{ji} + a^{ki}_{(j)} a^{jk} , \\
1801: \label{eq:MQL4}
1802: &a^{ki}_{(j)}a^{kj} = a^{kj}_{(i)}a^{ki} =
1803: c^{ij}_{(k)} + a^{ji}_{(k)}a^{kj}
1804: - a^{ij}_{(k)}a^{ki} ,
1805: \end{align}
1806: where $1\leq i < j < k \leq N$.
1807: Because
1808: \begin{equation}
1809: \bx\wedge\bx_{(ij)}\wedge\bx_{(ik)}\wedge\bx_{(jk)} =
1810: \bx\wedge\bx_{(i)}\wedge\bx_{(j)}\wedge\bx_{(k)}
1811: (a^{ij}a^{jk}a^{ki} - a^{ji}a^{kj}a^{ik}),
1812: \end{equation}
1813: then the BQL reduction condition is equivalent to
1814: \begin{equation} \label{eq:BKP-cond}
1815: a^{ij}a^{jk}a^{ki} - a^{ji}a^{kj}a^{ik} =0 .
1816: \end{equation}
1817:
1818: We will show that equation \eqref{eq:BKP-cond}
1819: implies existence the gauge function $\rho:\ZZ^N\to\RR$ such that
1820: \begin{align} \label{eq:rho-1}
1821: a^{ij}\rho_{(i)} = a^{ji}\rho_{(j)}, \quad
1822: a^{ik}\rho_{(i)} & = a^{ki}\rho_{(k)}, \quad
1823: a^{jk}\rho_{(j)} = a^{kj}\rho_{(k)}, \\
1824: \label{eq:rho-2}
1825: \rho_{(ij)} = c^{ij}\rho, \quad
1826: \rho_{(ik)} & = c^{ik}\rho, \quad
1827: \rho_{(jk)} = c^{jk}\rho.
1828: \end{align}
1829: Then, after rescaling $\bx \to \bx/\rho$, the new homogeneous coordinates
1830: satisfy the system \eqref{eq:BKP-lin}.
1831:
1832: Let us consider equations \eqref{eq:rho-1}-\eqref{eq:rho-2} as a difference
1833: system, which allows to calculate from $\rho$ and (say) $\rho_{(i)}$ values of
1834: the gauge function in remainig vertices of the hexahedron.
1835: Notice first that the condition \eqref{eq:BKP-cond} and the system
1836: \eqref{eq:MQL1}-\eqref{eq:MQL4} imply
1837: (it can be verified directly, but actually it follows from Corollary
1838: \ref{cor:BKP-impl}) that
1839: \begin{equation} \label{eq:BKP-impl}
1840: a^{ij}_{(k)}c^{ik} = a^{ji}_{(k)}c^{jk} , \quad
1841: a^{ik}_{(j)}c^{ij} = a^{ki}_{(j)}c^{jk} , \quad
1842: a^{jk}_{(i)}c^{ij} = a^{kj}_{(i)}c^{ik}.
1843: \end{equation}
1844:
1845: The condition \eqref{eq:BKP-cond} assures self-consistency of equations
1846: \eqref{eq:rho-1}. Then equations \eqref{eq:BKP-impl} imply consistency of
1847: equations \eqref{eq:rho-2} with the following consequence of equations
1848: \eqref{eq:rho-1}
1849: \begin{equation*}
1850: a^{ij}_{(k)}\rho_{(ik)} = a^{ji}_{(k)}\rho_{(jk)}, \quad
1851: a^{ik}_{(j)}\rho_{(ij)} = a^{ki}_{(j)}\rho_{(jk)}, \quad
1852: a^{jk}_{(i)}\rho_{(ij)} = a^{kj}_{(i)}\rho_{(ik)}. \\
1853: \end{equation*}
1854: Finally, the self-consistency of equations \eqref{eq:rho-2} in finding
1855: $\rho_{(ijk)}$ follows from equations \eqref{eq:rho-1} and the system
1856: \eqref{eq:MQL2}-\eqref{eq:MQL4}.
1857:
1858: \begin{Rem}
1859: Notice that the system
1860: \eqref{eq:MQL1}-\eqref{eq:MQL4} of \emph{nonlinear} equations can be
1861: considered as a system of eight \emph{linear} equations allowing for
1862: transition
1863: \begin{equation*}
1864: (a^{ij}, a^{ji}, a^{ik}, a^{ki}, a^{jk},
1865: a^{kj}, c^{ij}, c^{ik}, c^{jk}) \to
1866: (a^{ij}_{(k)}, a^{ji}_{(k)}, a^{ik}_{(j)}, a^{ki}_{(j)},
1867: a^{jk}_{(i)}, a^{kj}_{(i)}, c^{ij}_{(k)}, c^{ik}_{(j)}, c^{jk}_{(i)});
1868: \end{equation*}
1869: the difference in
1870: the number of unknowns and equations reflects the homogeneous nature of
1871: the linear system \eqref{eq:Laplace}.
1872: \end{Rem}
1873:
1874: \section{Pfaffians} \label{app:Pf}
1875: We recall basic properties of Pfaffians \cite{Muir,Proskuryakov}, which we use in
1876: Section~\ref{sec:fund-Pf}.
1877: Let $A=(a_{ij})_{1\leq i,j \leq 2r}$ be a skew symmetric matrix (i.e.,
1878: $a_{ji} = - a_{ij}$) of the even order $2r$. Consider the form
1879: \begin{equation}
1880: \omega = \sum_{i<j}a_{ij}e_i\wedge e_j,
1881: \end{equation}
1882: then the Pfaffian $\mathrm{Pf}(A)$ of $A$ is defined by
1883: \begin{equation}
1884: \omega^{\wedge r}=
1885: %\omega \stackrel{r \; \text{factors}}{\wedge \dots \wedge}\omega
1886: (r!)\;\mathrm{Pf}(A) \; e_1\wedge \dots \wedge e_{2r}.
1887: \end{equation}
1888: For each permutation $\pi$ of $\{1,\dots,2r\}$, put $A^\pi=(a_{\pi(i)\pi(j)})$,
1889: then
1890: \begin{equation}
1891: \mathrm{Pf}(A^\pi) = \mathrm{sgn} \; \pi \; \mathrm{Pf}(A).
1892: \end{equation}
1893: Notice the analogy with the determinant $\det B$ of an arbitrary square matrix
1894: $B=(b_{ij})_{1\leq i,j \leq n}$ expressed in terms of the forms
1895: \begin{equation}
1896: \omega_i = \sum_j b_{ij}e_j
1897: \end{equation}
1898: as follows
1899: \begin{equation}
1900: \omega_1\wedge \dots \wedge \omega_n = \det (B) \; e_1\wedge \dots \wedge e_{n}.
1901: \end{equation}
1902: It turns out that the determinant of any skew symmetric matrix of an even order
1903: equals to the square of its Pfaffian
1904: \begin{equation}
1905: \det (A) = (\mathrm{Pf}(A))^2.
1906: \end{equation}
1907:
1908: For any two subsets $I,J\subset\{1,\dots,n \}$ denote by $B(I,J)$
1909: the sub-matrix of $B$
1910: obtained by removing all the $i$th$\in I$ rows and all the
1911: $j$th$\in J$ columns of $B$.
1912: Then in analogy to the Laplace expansion of determinants
1913: \begin{equation}
1914: \delta_{ij} \det (B) = \sum_{k=1}^n b_{kj}(-1)^{k+i}\det (B(\{k\},\{i\})).
1915: \end{equation}
1916: we have the following
1917: expansion formula for Pfaffians
1918: \begin{equation} \label{eq:Laplace-Pfaff}
1919: \delta_{ij}\mathrm{Pf}(A) = \sum_{k=1}^{2r} a_{kj}(-1)^{k+i-1}
1920: \mathrm{Pf}(A(\{k,i\},\{k,i\})).
1921: \end{equation}
1922: Both formulas imply
1923: \begin{equation}
1924: \det (A(\{i\},\{j\})) = - \mathrm{Pf}(A)\; \mathrm{Pf}(A(\{i,j\},\{i,j\})),
1925: \end{equation}
1926: which leads to the following Pfaffian-Cramer rule
1927: for solutions of the linear
1928: system
1929: \begin{equation}
1930: A\boldsymbol{y} = \boldsymbol{b}
1931: \end{equation}
1932: with non-degenerate skew symmetrix matrix od the even order:
1933: \begin{equation} \label{eq:Pfaff-Cramer}
1934: y^j = \frac{\mathrm{Pf} (A[j])}{\mathrm{Pf} (A)},
1935: \end{equation}
1936: where by $A[j]$ is denoted the matrix $A$ whose $j$th column is replaced by
1937: \begin{equation}
1938: \boldsymbol{b^\prime} = \left( b^1,\dots , b^{j-1},
1939: 0 , - b^{j+1} , \dots , -b^{2r} \right)^{\mathrm{t}},
1940: \end{equation}
1941: and whose $j$th row is replaced by $-\boldsymbol{b^\prime}^{\mathrm{t}}$.
1942:
1943: When the order of the skew symmetric matrix $A$ is odd we have $\det (A) =0$,
1944: but the following formula holds
1945: \begin{equation} \label{eq:minors-odd-Pf}
1946: \det (A(\{i\},\{j\})) = \mathrm{Pf}(A(\{i\},\{i\})) \;\mathrm{Pf}(A(\{j\},\{j\})).
1947: \end{equation}
1948: \bibliographystyle{amsplain}
1949:
1950: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1951: \begin{thebibliography}{10}
1952:
1953: \bibitem{AKV}
1954: A. A. Akhmetishin, I. M. Krichever and Y. S. Volvovski, \emph{Discrete analogues
1955: of the Darboux--Egoroff metrics}, Proc. Steklov Inst. Math. \textbf{225} 16-39.
1956:
1957: \bibitem{BD}
1958: M.~Bia{\l}ecki, A.~Doliwa, \emph{Algebro-geometric solution of the discrete
1959: KP equation over a finite field out of a hyperelliptic curve},
1960: Comm. Math. Phys. \textbf{253} (2005), 157--170.
1961:
1962:
1963: \bibitem{BP2}
1964: A. I. Bobenko and U. Pinkall, \emph{Discrete isothermic surfaces}, J. Reine Angew.
1965: Math. \textbf{475} (1996) 187--208.
1966:
1967: \bibitem{BobSur}
1968: A.~I. Bobenko and Yu.~Suris, \emph{Discrete differential geometry. Consistency
1969: as integrability}, {\tt math.DG/0504358}.
1970:
1971: \bibitem{BBEIM}
1972: E. D. Belokolos, A. I. Bobenko, V. Z. Enol'skii, A. R. Its and V. B. Matveev,
1973: \emph{Algebro-geometric approach to nonlinear integrable equations}, Sronger,
1974: Berlin, 1994.
1975:
1976: \bibitem{BoKo}
1977: L. V. Bogdanov and B. G. Konopelchenko, \emph{Lattice and $q$-difference
1978: Darboux--Zakharov--Manakov systems via $\bar\partial$ method}, J. Phys. A: Math.
1979: Gen. \textbf{28} L173--L178.
1980:
1981: \bibitem{BoKo-N-KP}
1982: L. V. Bogdanov and B. G. Konoelchenko, \emph{Analytic-bilinear approach to
1983: integrable hiererchies II. Multicomponent KP and 2D Toda hiererchies},
1984: J. Math. Phys. \textbf{39} (1998) 4701--4728.
1985:
1986: \bibitem{CarrollSpeyer}
1987: G. D. Carroll and D. Speyer, \emph{The cube recurrence},
1988: Electron. J. Combin. \textbf{11} (2004), no. 1, Research Paper 73, 31 pp.
1989: (electronic).
1990:
1991: \bibitem{CDS}
1992: J. Cie\'{s}li\'{n}ski, A. Doliwa and P. M. Santini,
1993: {\it The Integrable Discrete Analogues of Orthogonal Coordinate Systems
1994: are Multidimensional Circular Lattices},
1995: Phys. Lett. A {\bf 235} (1997) 480--488.
1996:
1997:
1998: \bibitem{Coxeter}
1999: H.~S.~M.~Coxeter, \emph{Projective Geometry}, Springer, Berlin, 1974.
2000:
2001: \bibitem{Darboux-OS}
2002: G. Darboux, \emph{Le\c{c}ons sur les syst\'{e}mes orthogonaux et les
2003: coordonn\'{e}es curvilignes}, Gauthier-Villars, Paris, 1910.
2004:
2005: \bibitem{DKJM}
2006: E.~Date, M.~Kashiwara, M. Jimbo and T.~Miwa, \emph{Transformation groups for
2007: soliton equations}, [in:] Proceedings of RIMS Symposioum on Non-Linear
2008: Integrable Systems --- Classical Theory and Quantum Theory (M. Jimbo and T.
2009: Miwa, eds.) World Science Publishing Co., Singapore, 1983, pp. 39--119.
2010:
2011:
2012: \bibitem{DJKM-BKP}
2013: E.~Date, M. Jimbo, M.~Kashiwara and T.~Miwa, \emph{A new hierarchy of soliton
2014: equations of KP-type},
2015: Physica \textbf{4} D (1982) 343--365.
2016:
2017: \bibitem{DJKM-Prym-BKP}
2018: E.~Date, M.~Jimbo, M.~Kashiwara and T.~Miwa, \emph{Quasi-periodic solutions of
2019: the orthognal KP equation -- Transformation groups for soliton equations V},
2020: Publ. RIMS, Kyoto Univ. \textbf{18} (1982) 1111--1119.
2021:
2022: \bibitem{q-red}
2023: A.~Doliwa, \emph{Quadratic reductions of quadrilateral lattices}, J. Geom.
2024: Phys.
2025: \textbf{30} (1999), 169--186.
2026:
2027: \bibitem{AD-isoth}
2028: A. Doliwa, \emph{Generalized isothermic lattice}, in preparation.
2029:
2030: \bibitem{MQL}
2031: A.~Doliwa and P.~M. Santini, \emph{Multidimensional quadrilateral lattices
2032: are integrable}, Phys. Lett. A \textbf{233} (1997), 365--372.
2033:
2034: \bibitem{DS-sym}
2035: A.~Doliwa and P.~M. Santini, \emph{The symmetric, {D}-invariant and {E}gorov
2036: reductions of the
2037: quadrilateral lattice}, J. Geom. Phys. \textbf{36} (2000) 60--102.
2038:
2039: \bibitem{DS-EMP}
2040: A.~Doliwa and P.~M.~Santini, \emph{Integrable systems and discrete
2041: geometry}, [in:] Encyclopedia of Mathematical Physics, J. P. Fran\c{c}ois,
2042: G. Naber and T. S. Tsun (eds.), Elsevier, 2006, Vol. III, pp. 78-87.
2043:
2044: \bibitem{DBK}
2045: A.~Doliwa, M.~Bia{\l}ecki, P.~Klimczewski, \emph{The Hirota equation over finite
2046: fields: algebro-geometric approach and multisoliton solutions} J. Phys. A
2047: \textbf{36} (2003) 4827--4839.
2048:
2049: \bibitem{TQL}
2050: A. Doliwa, P. M. Santini and M. Ma{\~n}as,
2051: \emph{Transformations of Quadrilateral Lattices}, J. Math. Phys. \textbf{41}
2052: (2000) 944--990.
2053:
2054:
2055: \bibitem{DGNS}
2056: A.~Doliwa, P.~Grinevich, M.~Nieszporski, and P.~M. Santini, \emph{Integrable
2057: lattices and their sub-lattices:
2058: from the discrete Moutard (discrete Cauchy--Riemann) 4-point equation to the
2059: self-adjoint 5-point scheme},
2060: {\tt nlin.SI/0410046}.
2061:
2062: \bibitem{Fay}
2063: J.~D.~Fay, \emph{Theta functions on Riemann surfaces}, Springer, Berlin, 1973.
2064:
2065: \bibitem{FarkasKra}
2066: H. M. Farkas and I. Kra, \emph{Riemann surfaces}, Springer, Berlin -- New York, 1992.
2067:
2068: \bibitem{FominZelevinsky}
2069: S. Fomin and A. Zelevinsky, \emph{The Laurent phenomenon}, Adv. Appl. Math.
2070: \textbf{28} (2002) 119--144.
2071:
2072: \bibitem{Hirota-BKP-Pf}
2073: R. Hirota, \emph{Soliton solutions to the BKP equations. I. The Pfaffian
2074: technique}, J. Phys. Soc. Japan \textbf{58} (1989) 2285--2296.
2075:
2076: \bibitem{KvL}
2077: V. G. Kac and J. van de Leur, \emph{The $n$-component KP hiererchy and
2078: representation theory}, [in:] Important developments in soliton theory, (A. S.
2079: Fokas and V. E. Zakharov, eds.) Springer, Berlin, 1993, pp. 302--343.
2080:
2081: \bibitem{KoSch-trap}
2082: B.~G. Konopelchenko and W.~K. Schief, \emph{Trapezoidal discrete surfaces:
2083: geometry and integrability}, J. Geom. Phys. \textbf{31} (1999) 75--95.
2084:
2085: \bibitem{KoSchiefSBKP}
2086: B.~G. Konopelchenko and W.~K. Schief, \emph{Reciprocal figures, graphical
2087: statics and inversive geometry of the Schwarzian BKP hierarchy},
2088: Stud. Appl. Math. \textbf{109} (2002) 89--124.
2089:
2090: \bibitem{Krichever-Prym}
2091: I.~Krichever, \emph{A characterization of Prym varieties},
2092: \texttt{math.AG/0506238}.
2093:
2094: \bibitem{LiuManas-BKP}
2095: Q. P. Liu and M. Ma\~{n}as, \emph{Pfaffian form of Grammian determinant
2096: solutions of the BKP hierarchy}, {\tt solv-int/9806004}, Chinese Ann. Math. Ser.
2097: A \textbf{23} (2002) 693--698.
2098:
2099: \bibitem{MM}
2100: M. Ma\~{n}as, \emph{Fundamental transformation for quadrilateral lattices:
2101: first potentials and $\tau$-functions, symmetric and pseudo-Egorov reductions},
2102: J.~Phys. A \textbf{34} (2001) 10413--10421.
2103:
2104: \bibitem{MDS}
2105: M. Ma\~{n}as, A. Doliwa and P.M. Santini, \emph{Darboux transformations for
2106: multidimensional quadrilateral lattices. I}, Phys. Lett. A \textbf{232}
2107: (1997) 99--105.
2108:
2109:
2110: \bibitem{Miwa} T.~Miwa, \emph{On Hirota's difference equations},
2111: Proc. Japan Acad. \textbf{58} (1982) 9--12.
2112:
2113: \bibitem{Moebius}
2114: F. A. M|"{o}bius, \emph{Kann von zwei dreiseitigen Pyramiden eine jede in
2115: Bezug auf die andere um- und eingeschrieben zugleich heissen?},
2116: J. reine angew. Math. \textbf{3} (1828) 273-278.
2117:
2118: \bibitem{Muir}
2119: T. Muir, \emph{A treatise on the theory of determinants}, revised and enlarged
2120: by W. H. Metzler, Dover Publ., New York, 1960.
2121:
2122:
2123: \bibitem{Mumford}
2124: D.~Mumford, \emph{Prym varieties. I}, Contributions to analysis (a collection of
2125: papers dedicated to Lipman Bers), pp. 325--350, Academic Press, New York, 1974.
2126:
2127:
2128: \bibitem{NiSchief}
2129: J.~J.~C. Nimmo and W.~K. Schief, \emph{Superposition principles associated with
2130: the {Moutard} transformation. {An} integrable discretisation of a
2131: (2+1)-dimensional sine-{Gordon} system}, Proc. R. Soc. London A \textbf{453}
2132: (1997), 255--279.
2133:
2134: \bibitem{Pedoe}
2135: D. Pedoe, \emph{Geometry, a comprehensive course}, Dover Publications, New York,
2136: 1988.
2137:
2138: \bibitem{Propp}
2139: J. Propp, \emph{The many faces of alternating-sign matrices}, Discrete
2140: Mathematics and
2141: Theoretical Computer Science Proceedings \textbf{AA (DM-CCG)} (2001) 43--58.
2142:
2143: \bibitem{Proskuryakov}
2144: I. V. Proskuryakov, \emph{Problems in linear algebra}, Mir Publishers, Moscow,
2145: 1978.
2146:
2147:
2148: \bibitem{Sauer}
2149: R.~Sauer, \emph{Differenzengeometrie}, Springer, Berlin, 1970.
2150:
2151: \bibitem{Schief-JNMP}
2152: W.~K. Schief, \emph{Lattice geometry of the discrete Darboux, KP, BKP and CKP
2153: equations. Menelaus' and Carnot's theorems}, J. Nonl. Math. Phys. \textbf{10}
2154: Supplement 2 (2003) 194--208.
2155:
2156: \bibitem{Shiota-Prym}
2157: T.~Shiota, \emph{Prym varieties and soliton equations},
2158: Infinite dimensional Lie algebras and groups, (V.~G.~Kac, ed.), World
2159: Scientific, Singapore, 1989, pp. 407--448.
2160:
2161: \bibitem{Taimanov}
2162: I.~A.~Taimanov, \emph{Prym varieties of branch covers and nonlinear
2163: equations}, Matem. Sbornik \textbf{181} (1990) 934--950.
2164:
2165:
2166: \bibitem{TsujimotoHirota}
2167: S. Tsujimoto and R. Hirota, \emph{Pfaffian representation of solutions to the
2168: discrete BKP hierarchy in bilinear form}, J. Phys. Soc. Japan \textbf{66} (1997)
2169: 2797--2806.
2170:
2171: \bibitem{VeselovNovikov}
2172: A.~P.~Veselov and S.~P.~Novikov, \emph{Finite-gap two-dimensional potential
2173: Schr\"{o}dinger operators: explicit formulae and evolution equations}, Doklady
2174: AN USSR \textbf{279} (1984) 20--24.
2175:
2176: \end{thebibliography}
2177:
2178:
2179:
2180: \end{document}
2181:
2182:
2183: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2184: