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 6.25in
11: \textheight 9in
12: \topmargin -0.7cm
13: \leftmargin -3cm
14: \oddsidemargin=0cm
15: \evensidemargin=0cm
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{\EE}{{\mathbb E}}
39: \newcommand{\CC}{{\mathbb C}}
40: \newcommand{\FF}{{\mathbb F}}
41: \newcommand{\ZZ}{{\mathbb Z}}
42: \newcommand{\RR}{{\mathbb R}}
43: \newcommand{\AAf}{{\mathbb A}}
44: %%%-----------------------------------------------------------------------
45: \newcommand{\bx}{\boldsymbol{x}}
46: \newcommand{\by}{\boldsymbol{y}}
47: \newcommand{\vby}{\vec{\by}}
48: \newcommand{\bt}{\boldsymbol{t}}
49: \newcommand{\vbt}{\vec{\bt}}
50:
51: \newcommand{\bY}{{\boldsymbol Y}}
52: \newcommand{\tbY}{\tilde{\bY}}
53:
54: \newcommand{\bX}{\boldsymbol{X}}
55: \newcommand{\tbX}{\tilde{\bX}}
56:
57: \newcommand{\tP}{\tilde{P}}
58: \newcommand{\tv}{\tilde{v}}
59: \newcommand{\tH}{\tilde{H}}
60: \newcommand{\tQ}{\tilde{Q}}
61:
62: \newcommand{\D}{{\Delta}}
63: \newcommand{\tD}{\tilde{\D}}
64: \newcommand{\tA}{\tilde{A}}
65:
66: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67:
68: \begin{document}
69:
70: \title[The C-quadrilateral lattice]
71: {The C-(symmetric) quadrilateral lattice, its
72: transformations
73: and the algebro-geometric construction}
74:
75: \author{Adam Doliwa}
76:
77:
78: \address{A. Doliwa, Wydzia{\l} Matematyki i Informatyki,
79: Uniwersytet Warmi\'{n}sko-Mazurski w Olsztynie,
80: ul.~\.{Z}o{\l}nierska 14, 10-561 Olsztyn, Poland}
81:
82: \email{doliwa@matman.uwm.edu.pl}
83:
84: %
85: %\date{\today}
86: \keywords{integrable discrete geometry; discrete CKP equation; finite-gap
87: integration; Darboux transformations}
88: \subjclass[2000]{37K10, 37K20, 37K25, 37K35, 37K60, 39A10}
89:
90: \begin{abstract}
91: The C-quadrilateral lattice (CQL), called also the symmetric lattice,
92: provides geometric interpretation of the
93: discrete CKP equation within the quadrilateral lattice (QL) theory. We discuss
94: affine-geometric properties of the lattice emphasizing the role of the Gallucci
95: theorem in the multidimensional consistency of the CQL. Then we give the
96: algebro-geometric construction of the lattice. We also present the reduction of
97: the vectorial fundamental transformation of the QL to the CQL case. In the
98: Appendix we show a relation between the QL and the so called Darboux maps.
99: \end{abstract}
100: %{\it 2001 PACS:} 02.30.Ik, 02.40.Dr, 05.45.Yv, 04.60.Nc, 02.40.Hw,
101: \maketitle
102:
103: \tableofcontents
104:
105: \section{Introduction}
106: The difference equations play an increasing role in science. We have
107: learned to appreciate the inherent discreteness of physical phenomena at the
108: atomic and subatomic level. Also in modern theoretical
109: physics the assumption of
110: a space-time continuum is often being abandoned. Apart from physics
111: difference equations have
112: numerous applications, e.g., in numerical analysis, computer science,
113: mathematical biology and economics.
114:
115: The domain of discrete (difference) systems forms nowadays one of the focal
116: points in integrable systems research. Also the connection between geometry
117: and integrability, well known in the case of integrable differential equations
118: \cite{Sym,RogersSchief,GuHuZhou}, has been transferred to the discrete level
119: (see \cite{DS-EMP} and references therein).
120: A successful general approach towards description of this relation
121: is provided by the theory of
122: multidimensional quadrilateral lattices (QLs) \cite{MQL}. These are just
123: maps $x:\ZZ^N\to\PP^M$ ($3\leq N\leq M$) of multidimensional integer lattice
124: into projective space, with planar elementary quadrilaterals.
125: The integrable partial difference equation counterpart of the QLs are
126: the discrete Darboux equations (see Section \ref{sec:MQL-affine} for details),
127: being found first \cite{BoKo} as the most general
128: difference system integrable by the non-local
129: $\bar\partial$ dressing method.
130: It turns out that integrability of the discrete Darboux system is encoded in a
131: very simple geometric statement (see Fig.~\ref{fig:TiTjTkx}).
132: \begin{Lem}[The geometric integrability scheme] \label{lem:gen-hex}
133: Consider points $x_0$, $x_1$, $x_2$ and $x_3$ in general position in $\PP^M$,
134: $M\geq 3$. On
135: the plane $\langle x_0, x_i, x_j \rangle$, $1\leq i < j \leq 3$ choose a point
136: $x_{ij}$ not on the lines $\langle x_0, x_i \rangle$, $\langle x_0,x_j
137: \rangle$ and $\langle x_i, x_j \rangle$. Then there exists the
138: unique point $x_{123}$
139: which belongs simultaneously to the three planes
140: $\langle x_3, x_{13}, x_{23} \rangle$,
141: $\langle x_2, x_{12}, x_{23} \rangle$ and
142: $\langle x_1, x_{12}, x_{13} \rangle$.
143: \end{Lem}
144: \begin{figure}
145: \begin{center}
146: \includegraphics{TiTjTkx.eps}
147: \end{center}
148: \caption{The geometric integrability scheme}
149: \label{fig:TiTjTkx}
150: \end{figure}
151:
152: Integrable reductions of the quadrilateral lattice (and thus of the discrete
153: Darboux equations) arise from additional constraints which are compatible with
154: geometric integrability scheme (see, for example \cite{q-red,DS-sym,BQL}).
155: Because application of several integrable reductions
156: preserves integrability of the lattice, it is important to isolate the basic
157: ones.
158: Integrable constraints of the quadrilateral lattice
159: may have local or global nature.
160: The global constraints are related with existence of an
161: additional geometric structure in the projective space which allows to impose
162: some restrictions on the lattice.
163: The reduction considered in this paper can be described
164: within the affine geometry approach.
165:
166: The differential
167: Darboux equations, which have appeared first in projective
168: differential geometry of multidimensional conjugate nets \cite{Darboux-OS},
169: play an important role \cite{BoKo-N-KP} in the multicomponent
170: Kadomtsev--Petviashvilii (KP) hierarchy,
171: which is commonly considered \cite{DKJM,KvL}
172: as the fundamental system of equations in integrability theory.
173: One of the most important reductions of the KP hierarchy of nonlinear equations
174: is the so called CKP hierarchy \cite{DJKM-CKP}
175: (here ``C" appears in the context of the classification theory of simple
176: Lie algebras). In \cite{AvL} it was shown that the differential Darboux
177: equations of the multicomponent KP hierarchy should be
178: in this case supplemented by certain symmetry condition.
179: In fact such equations were considered first
180: within the differential geometry context in \cite{Darboux-OS,Bianchi}. Their
181: discrete counterpart was studied in \cite{DS-sym} under the name of the
182: symmetric quadrilateral lattice. In \cite{DS-sym} the symmetric Darboux
183: equations have been solved by the non-local
184: $\bar\partial$ dressing method. The corresponding reduction of the fundamental
185: transformation \cite{TQL} of QL was given on the algebraic level
186: in \cite{MM}.
187:
188: Within the context of the so
189: called Darboux maps such a reduction of the discrete Darboux equations was
190: studied in \cite{Schief-JNMP}, where the system was reformulated in
191: a convenient scalar form (see also \cite{DJM})
192: and related with the superposition formulas for the Darboux
193: transformations of the CKP hierarchy. In \cite{KingSchief} such special Darboux
194: maps were characterized geometrically using the classical Pascal's theorem for
195: hexagons inscribed in conics. Also the B\"{a}cklund transformation for such
196: maps was investigated geometrically in \cite{KingSchief}.
197: In fact,
198: the attempt to understand results of \cite{Schief-JNMP,KingSchief} in the
199: quadrilateral lattice approach was my
200: motivation to undertake again, after \cite{DS-sym},
201: studies of the symmetric reduction of the Darboux equations.
202:
203: An important reduction of the quadrilateral lattice with the symmetric rotation
204: coefficients is the Egorov lattice \cite{Schief-Egorov}. It was solved in
205: \cite{DS-sym} using the inverse spectral transform. In \cite{AKV} the
206: algebro-geometric techniques were used to construct large class of such
207: lattices. In 1998 Peter Grinevich
208: isolated from \cite{AKV} these of the algebro-geometric conditions
209: which give rise the symmetric quadrilateral lattices \cite{PG-private}.
210:
211: The main goal of the paper is to give new pure geometric characterization of
212: the symmetric quadrilateral lattice and of the
213: corresponding reduction of the fundamental transformation.
214: Because of the intimate relation of the lattice with
215: the discrete CKP equation we will call it also the C-quadrilateral lattice.
216: We study the geometric and algebraic properties of the CQL reduction of the
217: fundamental transformation. In particular, we show that the
218: reduced transformation satisfies the Bianchi permutability principle. To
219: make the theory of such lattices complete we apply the algebro-geometric
220: techniques to construct large classes of the lattices together with
221: corresponding solutions of the Darboux equations. In doing that we introduce
222: the
223: backward and the dual Baker--Akhiezer functions of the quadrilateral lattice.
224:
225: The paper is organized as follows. In Section \ref{sec:MQL-affine} we summarize
226: relevant facts from the quadrilateral lattice theory and we introduce the notation.
227: In Section \ref{sec:CQL} we introduce new geometric definition of the
228: C-quadrilateral lattice and we study its integrability (multidimensional
229: consistency) using geometric means. We
230: also present its algebraic description in terms of symmetric Darboux
231: equations. Section \ref{sec:AG} is devoted to presentation of the
232: algebro-geometric construction of the C-quadrilateral (symmetric) lattices.
233: In Section \ref{sec:C-fund} we introduce geometrically the C-reduction
234: of the fundamental transformation of the quadrilateral lattice \cite{TQL} and we
235: link it with the earlier algebraic results of \cite{MM}. We also prove the
236: corresponding permutability theorem for this transformation. Finally in the
237: Appendix we present briefly
238: the relation between the Darboux maps \cite{Schief-JNMP} and the
239: quadrilateral lattices.
240:
241: The main geometric results of the paper were presented on the
242: LMS Research Symposium \emph{Methods of Integrable Systems in Geometry},
243: Durham, UK (August, 2006). It is my pleasure to thank the organizers of the
244: Meeting for invitation and support.
245:
246: \section{The multidimensional quadrilateral lattice (affine description)}
247: \label{sec:MQL-affine}
248: \subsection{The discrete Laplace and Darboux equations}
249: Consider a multidimensional quadrilateral lattice (MQL); i.~e., a mapping
250: $x :\ZZ^N \rightarrow \PP^M(\FF)$, $3\leq N\leq M$,
251: with all the elementary quadrilaterals planar \cite{MQL}; here $ZZ^N$ is the
252: $N$-dimensional integer lattice, and $\PP^M(\FF)$ is $M$-dimensional projective
253: space over the field $\FF$.
254: In the affine gauge (which will be extremely useful in the paper) the
255: lattice is represented by a mapping
256: $\bx: \ZZ^N \rightarrow \FF^M$ and the planarity
257: condition can be formulated in terms of the Laplace equations
258: \begin{equation} \label{eq:Laplace}
259: \D_i\D_j\bx=(T_{i} A_{ij})\D_i\bx+
260: (T_j A_{ji})\D_j\bx, \qquad i\not= j, \qquad i,j=1 ,\dots, N,
261: \end{equation}
262: where $T_i$ is the translation operator in the $i$-th direction, and
263: $\D_i = T_i - 1$ is the corresponding partial difference operator.
264:
265: Due to compatibility of the system \eqref{eq:Laplace}
266: the coefficients $A_{ij}$ satisfy the MQL (or discrete Darboux) system of
267: equations
268: \begin{equation} \label{eq:MQL-A}
269: \D_k A_{ij} + (T_kA_{ij})A_{ik} =
270: (T_jA_{jk})A_{ij} +(T_k A_{kj})A_{ik} ,
271: \qquad i, j, k \quad \text{distinct}.
272: \end{equation}
273: The $j\leftrightarrow k$ symmetry of RHS of \eqref{eq:MQL-A} implies existence
274: of the potentials $H_{i}$, $i=1, \dots , N$,
275: (called the Lam\'e coefficients) such that
276: \begin{equation} \label{def:A-H}
277: A_{ij}= \frac{\D_j H_i}{H_i} \; , \qquad i\ne j \; .
278: \end{equation}
279: If we introduce the suitably scaled tangent
280: vectors $\bX_i$, $i=1,...,N$,
281: \begin{equation} \label{def:HX}
282: \D_i\bx = (T_iH_i) \bX_i,
283: \end{equation}
284: then equations \eqref{eq:Laplace} can be rewritten as a
285: first order system expressing the fact that $j$th variation of $\bX_i$ is
286: proportional to $\bX_j$ only
287: \begin{equation} \label{eq:lin-X}
288: \D_j\bX_i = (T_j Q_{ij})\bX_j, \qquad i\ne j \; .
289: \end{equation}
290: The proportionality factors $Q_{ij}$, called the rotation coefficients,
291: can be found from the linear equations
292: \begin{equation} \label{eq:lin-H}
293: \D_iH_j = (T_iH_i) Q_{ij}, \qquad i\ne j \; ,
294: \end{equation}
295: adjoint to \eqref{eq:lin-X}.
296: The compatibility condition for the system~\eqref{eq:lin-X} (or its
297: adjoint)
298: gives the following new form of the discrete Darboux equations
299: \begin{equation} \label{eq:MQL-Q}
300: \D_kQ_{ij} = (T_kQ_{ik})Q_{kj}, \qquad i\neq j\neq k\neq i.
301: \end{equation}
302:
303: \subsection{The backward data, the connection factors and the $\tau$-function}
304: The backward tangent vectors $\tbX_i$, $i=1,\dots,N$, are defined similarly to the
305: forward tangent vectors $\bX_i$ but with the help of the backward
306: shifts $T_i^{-1}$ and of the backward
307: difference operator $\tD_i := 1- T_i^{-1}$:
308: \begin{equation} \label{eq:lin-bX}
309: \tD_i\tbX_j = (T_i^{-1} \tQ_{ij})\tbX_i \; , \qquad \text{or}
310: \qquad \D_i\tbX_j = (T_i\tbX_i)\tQ_{ij}, \quad i\ne j ;
311: \end{equation}
312: we define also the backward rotation coefficients $\tQ_{ij}$ as the
313: corresponding proportionality factors. Notice that the backward tangent
314: vectors $\tbX_i$ satisfy the adjoint linear system \eqref{eq:lin-H}.
315: \begin{figure}
316: \begin{center}
317: \includegraphics[width=8cm]{back-ij.eps}
318: \end{center}
319: \caption{Definition of the backward data}
320: \label{fig:back}
321: \end{figure}
322:
323: The backward Lam\'e coefficients $\tH_i$, , $i=1,\dots,N$, are defined by
324: \begin{equation} \label{eq:b-H-X}
325: \tD_i\bx = (T_i^{-1}\tH_i ) \tbX_i \; , \qquad \text{or}
326: \qquad \D_i\bx = \tH_i (T_i\tbX_i ) \, ,
327: \end{equation}
328: and satisfy the system
329: \begin{equation} \label{eq:lin-bH}
330: \tD_j\tH_i = (T_j^{-1}\tH_{j})\tQ_{ij}, \qquad \text{or}
331: \qquad \D_j\tH_i = (T_j\tQ_{ij})\tH_{j}, \quad i\ne j \; .
332: \end{equation}
333: As a consequence of equations \eqref{eq:lin-bX} or equations \eqref{eq:lin-bH}
334: the functions $\tQ_{ij}$ satisfy the MQL
335: equations~\eqref{eq:MQL-Q}.
336: Define backward Laplace coefficients
337: \begin{equation}
338: \tA_{ij} = \frac{\tD_j \tH_i}{\tH_i}\; ,
339: \end{equation}
340: then $\bx$ satisfies the backward Laplace equation
341: \begin{equation} \label{eq:Laplace-b}
342: \tD_i\tD_j\bx = (T_i^{-1}\tA_{ij})\tD_i\bx +
343: (T_j^{-1}\tA_{ji})\tD_j\bx
344: \; , \qquad i\ne j \; .
345: \end{equation}
346:
347:
348: The connection factors $\rho_i:\ZZ^N\to \FF$ are the
349: proportionality coefficients between
350: $\bX_i$ and $T_i\tbX_i$ (both vectors are proportional to $\D_i\bx$):
351: \begin{equation} \label{eq:def-rho}
352: \bX_i = \rho_i ( T_i\tbX_i) \; , \qquad
353: i=1,\dots ,N \; .
354: \end{equation}
355: Other forward and backward data of the lattice $\bx$ are related
356: through the following formulas
357: \begin{align} \label{eq:Q-Qt}
358: \tH_i & = \rho_i T_iH_i , \\
359: \rho_j T_j\tQ_{ij} & = \rho_i T_iQ_{ji} \; .
360: \end{align}
361: Moreover, the factors $\rho_i$ are satisfy
362: equations
363: \begin{equation} \label{eq:rho-constr}
364: \frac{T_j\rho_i}{\rho_i} = 1 - (T_iQ_{ji})(T_jQ_{ij}), \qquad i\ne j \; ,
365: \end{equation}
366: which imply existence of yet another potential
367: (the $\tau$-function of the quadrilateral lattice) such that
368: \begin{equation} \label{eq:tau}
369: \rho_i = \frac{T_i \tau}{\tau} .
370: \end{equation}
371:
372: Given a quadrilateral lattice $\bx$, the forward data $\{\bX_i, H_i, Q_{ij}\}$
373: and the backward data $\{\tilde\bX_i, \tilde{H}_i, \tilde{Q}_{ij}\}$
374: are defined up to rescaling by functions $a_i(m_i)$, $b_i(m_i)$ of single
375: variables
376: \begin{gather} \label{eq:forward-scaling}
377: \bX_i \to a_i\bX_i, \qquad T_i H_i \to \frac{1}{a_i} T_i H_i , \qquad
378: T_j Q_{ij} \to \frac{a_i}{a_j} T_j Q_{ij}, \\
379: \label{eq:backward-scaling}
380: T_i\tilde\bX_i \to \frac{1}{b_i}T_i\tilde\bX_i,
381: \qquad \tilde{H}_i \to b_i\tilde{H}_i, \qquad
382: T_j \tilde{Q}_{ij} \to \frac{b_i}{b_j} T_j \tilde{Q}_{ij},
383: \end{gather}
384: then also
385: \begin{equation}
386: \rho_i \to a_i b_i \rho_i.
387: \end{equation}
388:
389: \begin{Rem}
390: Since $Q_{ij}$ and $\tQ_{ij}$ are both solutions of the discrete Darboux
391: equations \eqref{eq:MQL-Q}, then equations
392: \eqref{eq:def-rho}-\eqref{eq:rho-constr} describe a special symmetry
393: transformations of \eqref{eq:MQL-Q}, first found in \cite{KoSchief2} without any
394: associated geometric meaning.
395: \end{Rem}
396:
397: \section{The C-quadrilateral lattice}
398: \label{sec:CQL}
399: \subsection{Geometric definition of the C-quadrilateral lattice}
400: The geometric arena, where the reduction of the quadrilateral lattice studied in
401: this paper lives, is the affine space. Recall that
402: the affine transformations are these
403: projective transformations which leave invariant a fixed hyperplane
404: $H_\infty\subset\PP^M$, called the hyperplane at infinity (see, for example
405: \cite{Coxeter-PG}). Two lines of $\AAf^M = \PP^M \setminus H_\infty$
406: (identified with $FF^M$, up to
407: fixing the origin) are
408: called parallel if they intersect in a point of $H_\infty$
409:
410: \begin{Def} \label{def:C-hexahedron}
411: A hexahedron with planar faces in the affine space
412: $\AAf^M$ is called a \emph{C-hexahedron} if the three points obtained by
413: intersection of the common lines of the pairs of planes of its opposite
414: faces with the hyperplane $H_\infty$ at infinity
415: are collinear (see Figure~\ref{fig:CQL-constr}).
416: \end{Def}
417: \begin{figure}
418: \begin{center}
419: \includegraphics[width=12cm]{CQL-constr4.eps}
420: \end{center}
421: \caption{The C-hexahedron}
422: \label{fig:CQL-constr}
423: \end{figure}
424: \begin{Rem}
425: We exclude for a time being
426: from our considerations the degenerate case when
427: a pair of opposite faces of the hexahedron is parallel, i.~e. their common
428: line belongs to $H_\infty$.
429: \end{Rem}
430: \begin{Cor}
431: Equivalently the C-hexahedron is characterized by coplanarity of
432: suitable parallel shifts of
433: the intersection lines of the opposite face planes.
434: \end{Cor}
435: \begin{Def} \label{def:C-hex}
436: A quadrilateral lattice $\bx:\ZZ^N\to\AAf^M$ is called a
437: \emph{C-quadrilateral lattice} (CQL) if all its hexahedra are C-hexahedra.
438: \end{Def}
439: The following Proposition gives an analytic characterization of C-quadrilateral
440: lattices in terms of their rotation coefficients.
441: \begin{Prop} \label{prop:Q3-CQL}
442: A quadrilateral lattice is subject to C-reduction if and only if its
443: rotation coefficients satisfy the constraint
444: \begin{equation} \label{eq:CQL-constr}
445: (T_jQ_{ij}) (T_kQ_{jk}) (T_iQ_{ki}) = (T_jQ_{kj}) (T_kQ_{ik}) (T_iQ_{ji}),
446: \qquad i, j, k \quad \text{distinct}.
447: \end{equation}
448: \end{Prop}
449: \begin{proof}
450: Denote by $\bt^k_{ij}$ ($i,j,k$ are
451: distinct) the direction vector of the common line of the plane
452: $\langle \bx, T_i \bx, T_j\bx \rangle$ and its $k$-opposite
453: $\langle T_k\bx, T_i T_k \bx, T_j T_k \bx \rangle$.
454: It must be therefore decomposed in
455: the basis $\{ \bX_i,\bX_j \}$ and in the basis $\{ T_k\bX_i,T_k \bX_j \}$.
456: Assuming its decomposition in the second basis we get
457: \begin{equation*}
458: \bt^k_{ij} = a T_k\bX_i + b T_k \bX_j = a \bX_i + b \bX_j +
459: (a T_k Q_{ik} + b T_k Q_{jk} ) \bX_k,
460: \end{equation*}
461: where we have used the linear problem \eqref{eq:lin-X}. Because the coefficient
462: in front of $\bX_k$ must vanish, the vector can be therefore
463: chosen as
464: \begin{equation}
465: \bt^k_{ij} = T_k Q_{jk} \bX_i - T_k Q_{ik} \bX_j .
466: \end{equation}
467: Notice that the condition in Definition~\ref{def:C-hex} is equivalent to
468: to coplanarity of the vectors $\bt^k_{ij}$, $\bt^j_{ik}$ and
469: $\bt^i_{jk}$, and the statement follows from equation
470: \begin{equation*}
471: \bt^k_{ij}\wedge\bt^j_{ik}\wedge\bt^i_{jk} =
472: \left( (T_jQ_{ij}) (T_kQ_{jk}) (T_iQ_{ki}) -
473: (T_jQ_{kj}) (T_kQ_{ik}) (T_iQ_{ji})
474: \right) \bX_i \wedge \bX_j \wedge \bX_k, \qquad i, j, k
475: \quad \text{distinct}.
476: \end{equation*}
477: \end{proof}
478: \begin{Rem}
479: In the degenerate case, when
480: a pair of opposite faces of the hexahedron is parallel, the corresponding
481: rotation coefficients vanish. Because they appear on different sides of equation
482: \eqref{eq:CQL-constr}, then the constraint is automatically satisfied.
483: \end{Rem}
484: As it was shown in \cite{DS-sym}, the constraint \eqref{eq:CQL-constr}
485: allows to rescale the forward and backward
486: data, using possibility given by equations
487: \eqref{eq:forward-scaling}-\eqref{eq:backward-scaling},
488: to the form such that
489: \begin{equation} \label{eq:symm-constr}
490: Q_{ij} = \tilde{Q}_{ij}, \qquad \text{or}
491: \qquad \rho_i T_i Q_{ji} = \rho_j T_j Q_{ij}, \qquad i\ne j,
492: \end{equation}
493: i.e., the rotation coefficients of CQL
494: are, in a sense, \emph{symmetric} with respect to
495: interchanging of their indices.
496: Under such a constraint equations \eqref{eq:rho-constr} and \eqref{eq:tau}
497: allow to express the rotation coefficients in terms of the $\tau$-function
498: as follows~\cite{Schief-JNMP}
499: \begin{equation}
500: (T_jQ_{ij})^2 = \frac{T_i \tau}{T_j \tau}
501: \left( 1 - \frac{(T_i T_j \tau) \tau}{(T_i \tau)( T_j \tau)} \right).
502: \end{equation}
503: Then the discrete
504: Darboux equations equations \eqref{eq:MQL-Q} can be
505: rewritten in the following quartic form
506: \begin{equation} \begin{split}
507: (\tau \, T_i T_j T_k \tau - T_i \tau \, T_j T_k \tau -
508: T_j \tau \, T_i T_k \tau -T_k \tau \, T_i T_j & \tau )^2 = \\
509: 4( T_i \tau \, T_j \tau \, T_i T_k \tau \, T_j T_k \tau +
510: T_i \tau \, T_k \tau \, T_i T_j \tau & \, T_j T_k \tau +
511: T_j \tau \, T_k \tau \, T_i T_k \tau \, T_i T_j \tau - \\
512: & T_i \tau \, T_j \tau \, T_k \tau \, T_i T_j T_k \tau -
513: \tau \, T_i T_j \tau \, T_j T_k \tau \, T_i T_k \tau ),
514: \end{split} \end{equation}
515: called in~\cite{Schief-JNMP} the discrete CKP equation.
516:
517: Finally, we present a result which we will use in Section
518: \ref{sec:C-fund}. Its formal algebraic proof can be found in \cite{MM} but,
519: essentially, it uses the
520: facts that (i) the functions $\rho_i$ connect solutions of the forward and
521: backward linear problems, (ii) a solution of the adjoint linear problem
522: \eqref{eq:lin-bH} is connected in this correspondence with solution of the
523: linear problem \eqref{eq:lin-X} but with backward rotation coefficients, and
524: (iii) backward and forward rotation coefficients in the CQL reduction coincide.
525: \begin{Lem}[\cite{MM}] \label{lem-ry*-Y}
526: Given solution
527: $\bY_i^*:\ZZ^N\to(\FF^K)^*$ of the adjoint linear problem \eqref{eq:lin-H}
528: for "symmetric" rotation coefficients \eqref{eq:symm-constr} then
529: $\rho_i (T_i\bY_i^*)^t:\ZZ^N\to\FF^K$ satisfies the corresponding linear
530: problem \eqref{eq:lin-X}.
531: \end{Lem}
532:
533: \subsection{Multidimensional consistency of the C-constraint}
534: \label{sec:M-cons-CQL}
535: As it was shown in \cite{MQL} the planarity condition, which allows to
536: construct
537: the point $x_{123}$ as in Lemma \ref{lem:gen-hex}, does not lead to any further
538: restrictions if we increase dimension of the lattice. Because in the CQL case
539: the constraint is imposed on the 3D (elementary hexahedra)
540: level, then to assure its multidimensional consistency we have to check the
541: four dimensional consistency. The multidimensional
542: consistency of the C-reduction with the geometric integrability scheme
543: would be the immediate consequence of its 4D consistency.
544:
545: In fact, the consistency of the constraint has
546: been proved algebraically
547: in \cite{DS-sym} starting from its algebraic form \eqref{eq:symm-constr}.
548: However, in that proof the main difficulty was shifted to the proof of existence
549: (in multidimensions) of the special choice of the $\tau$-function. In this paper
550: we present pure geometric proof of the 4D consistency of the CQL. We first
551: recall the relevant result on four dimensional consistency of the QL, which
552: is the consequence of the of the following geometric observation.
553: \begin{Lem}[The 4D consistency of the geometric integrability scheme]
554: \label{lem:4D-consist-QL}
555: Consider points $x_0$, $x_1$, $x_2$, $x_3$ and $x_4$
556: in general position in $\PP^M$, $M\geq 4$. Choose generic points
557: $x_{ij}\in\langle x_0, x_i, x_j \rangle$, $1\leq i < j \leq 4$,
558: on the corresponding planes, and using
559: the planarity condition construct the points
560: $x_{ijk}\in\langle x_0, x_i, x_j , x_k\rangle$, $1\leq i < j < k \leq 4$ -- the
561: remaining vertices of the four (combinatorial) cubes.
562: Then the intersection point $x_{1234}$ of the three planes
563: \[\langle x_{12}, x_{123}, x_{124} \rangle, \;
564: \langle x_{13}, x_{123}, x_{134} \rangle, \;
565: \langle x_{14}, x_{124}, x_{134} \rangle \quad \text{in} \quad
566: \langle x_{1}, x_{12}, x_{13}, x_{14} \rangle,
567: \]
568: coincides with
569: the intersection point of the three planes
570: \[\langle x_{12}, x_{123}, x_{124} \rangle, \;
571: \langle x_{23}, x_{123}, x_{234} \rangle, \;
572: \langle x_{24}, x_{124}, x_{234} \rangle, \quad \text{in} \quad
573: \langle x_{2}, x_{12}, x_{23}, x_{24} \rangle,
574: \]
575: which is the same as
576: the intersection point of the three planes
577: \[\langle x_{13}, x_{123}, x_{134} \rangle, \;
578: \langle x_{23}, x_{123}, x_{234} \rangle, \;
579: \langle x_{34}, x_{134}, x_{234} \rangle, \quad \text{in} \quad
580: \langle x_{3}, x_{13}, x_{23}, x_{34} \rangle,
581: \]
582: and
583: the intersection point of the three planes
584: \[\langle x_{14}, x_{124}, x_{134} \rangle, \;
585: \langle x_{24}, x_{124}, x_{234} \rangle, \;
586: \langle x_{34}, x_{134}, x_{234} \rangle, \quad \text{in} \quad
587: \langle x_{4}, x_{14}, x_{24}, x_{34} \rangle.
588: \]
589: \end{Lem}
590:
591: \begin{Rem}
592: In fact, the point $x_{1234}$ is the unique
593: intersection point of the four three dimensional subspaces
594: $\langle x_{1}, x_{12}, x_{13}, x_{14} \rangle$,
595: $\langle x_{2}, x_{12}, x_{23}, x_{24} \rangle$,
596: $\langle x_{3}, x_{13}, x_{23}, x_{34} \rangle$,
597: and
598: $\langle x_{4}, x_{14}, x_{24}, x_{34} \rangle$ of the four dimensional subspace
599: $\langle x_{0}, x_{1}, x_{2}, x_{3} , x_{4} \rangle$. This observation
600: generalizes naturally to the case of more dimensional hypercube with the
601: planar facets.
602: \end{Rem}
603:
604:
605: It turns out
606: that the geometric core of the integrability of CQL is provided by
607: the Gallucci's theorem on eight skew lines (see, for example \cite{Coxeter-PG}).
608: \begin{Th}[The Gallucci theorem]
609: If three skew lines all meet three other skew lines, any transversal to the
610: first set of three meets any transversal to the second set.
611: \end{Th}
612: Motivated by the corresponding results
613: of \cite{KingSchief} we first prove a useful Lemma,
614: where the Gallucci theorem is used.
615: \begin{Lem} \label{lem:C-consistency}
616: Consider four adjacent hexahedra with planar faces which share a vertex of a
617: 4D hypercube. If three of them are C-hexahedra then
618: the same holds also for the forth one.
619: \end{Lem}
620: \begin{proof}
621: We start from generic 4D hexahedron with planar faces. Denote by
622: $\pi_{1234}$
623: the three dimensional subspace of $H_\infty$ being its intersection with the 4D
624: subspace $\langle x_0, x_1, x_2, x_3, x_4 \rangle$ of the hexahedron. By
625: $\pi_{ijk}$ denote the intersection line of the 3D subspace $\langle x_0, x_i,
626: x_j, x_k \rangle$ of one of the four 3D hexahedra with $H_\infty$.
627:
628: Denote by $h^k_{ij}$ the intersection points of the
629: opposite planes $\langle x_0, x_i, x_j \rangle$,
630: $\langle x_k, x_{ik}, x_{jk} \rangle$ of the four
631: 3D hexahedra with $H_\infty$
632: (all indices are distinct and range from $1$ to $4$, eventually
633: they should be reordered); these are the points entering into the definition of the
634: C-hexahedra.
635: By $h^\ell_{ijk}$ denote the intersection lines of the
636: three dimensional subspaces $\langle x_0, x_i, x_j, x_k \rangle$,
637: $\langle x_\ell, x_{i\ell}, x_{j\ell}, x_{k\ell} \rangle$ and
638: $H_\infty$. There are four lines of this type (the $h$-family).
639: Notice that point $h^k_{ij}$ is the intersection point of $h^k_{ij\ell}$ with
640: the plane $\pi_{ijk}$. On the plane $\pi_{ijk}$ we have therefore apart from
641: three collinear point $h^\ell_{ij}$, $h^\ell_{ik}$ and $h^\ell_{jk}$ (they
642: belong to $h^\ell_{ijk}$)
643: also three
644: other points $h^k_{ij}$, $h^j_{ik}$ and $h^i_{jk}$.
645:
646:
647: The C-reduction condition of the
648: hexahedron with with vertices $x_0$, $x_i$, $x_j$ and $x_k$ means collinearity
649: of the points points $h^k_{ij}$, $h^j_{ik}$ and $h^i_{jk}$. If such a line
650: exists we denote it by $g_{ijk}$.
651: It intersects not
652: only lines $h^k_{ij\ell}$, $h^j_{ik\ell}$ and $h^i_{jk\ell}$ (in points
653: $h^k_{ij}$, $h^j_{ik}$ and $h^i_{jk}$, correspondingly), but also the fourth
654: line $h^\ell_{ijk}$ of the $h$-family, as both lines belong to the plane
655: $\pi_{ijk}$.
656:
657: Let us assume that the C-reduction condition is satisfied for three hexahedra,
658: i.e., three such $g$-lines exist, all transversal to four $h$-lines.
659: Assume that $g_{ijk}$, $g_{ij\ell}$ and $g_{ik\ell}$ exist,
660: define the line $G_{jk\ell}$ as the unique line passing through the points
661: $h^j_{k\ell}$ and $h^k_{j\ell}$, thus transversal to the lines
662: $h^j_{ik\ell}$ and $h^k_{ij\ell}$.
663: Because $G_{jk\ell}$ is contained in $\pi_{jk\ell}$ then it must intersect
664: also the line $h^i_{jk\ell}$ --- the third line of the $h$-family.
665: By Gallucci's theorem it
666: intersects therefore the forth line $h^\ell_{ijk}$.
667:
668: That intersection point belongs to the plane $\pi_{jk\ell}$ (containing
669: the points $h^j_{k\ell}$ and $h^k_{j\ell}$ which define $G_{jk\ell}$)
670: therefore it must be $h^\ell_{jk} = h^\ell_{ijk} \cap \pi_{jk\ell}$.
671: We have therefore shown that also the three points
672: $h^j_{k\ell}$, $h^k_{j\ell}$ and $h^\ell_{jk}$ are collinear, i.e.,
673: $G_{jk\ell}= g_{jk\ell}$.
674: \end{proof}
675: \begin{Prop}
676: Under hypotheses of Lemma~\ref{lem:gen-hex} assume that the C-reduction
677: condition is satisfied for 3D hexahedra meeting in one vertex of a 4D
678: hypercube.
679: Then the condition holds for all hexahedra of the hypercube.
680: \end{Prop}
681: \begin{proof}
682: Without loss of generality assume that the common vertex of the four hexahedra
683: (we know that it is enough to assume the C-reduction condition for three of
684: them) is $x_0$. Then three of them share the vertex $x_i$, which implies the
685: constraint for the forth one. All remaining
686: four hexahedra of the 4D hypercube which do
687: not share the point $x_0$ are of that type.
688: \end{proof}
689:
690:
691: On the level of the
692: multidimensional quadrilateral lattice there exists simple
693: alternative algebraic proof of an analogue of Lemma~\ref{lem:C-consistency},
694: which combined with ideas behind the proof of Proposition~\ref{prop:Q3-CQL}
695: would give the algebraic proof of the Lemma.
696: \begin{Cor}
697: Consider a 4D hexahedron withe vertices
698: $x$, $T_i x$, $T_j x$, $T_k x$, $T_\ell x$ of the quadrilateral lattice
699: $x:\ZZ^N\to\AAf^M$, $4\leq N \leq M$. If three of the four
700: 3D hexahedra meeting in the vertex $x$ are the C-hexahedra then the
701: same holds also for the forth one.
702: \end{Cor}
703: \begin{proof}
704: Assume that the fourth one is the hexahedron with basic vertices
705: $x$, $T_j x$, $T_k x$ and $T_\ell x$.
706: By Proposition~\ref{prop:Q3-CQL} the C-reduction
707: condition for the remaining three reads
708: \begin{align}
709: (T_jQ_{ij}) (T_kQ_{jk}) (T_iQ_{ki}) &= (T_jQ_{kj}) (T_kQ_{ik}) (T_iQ_{ji}),\\
710: (T_i Q_{ji}) (T_j Q_{\ell j}) (T_\ell Q_{i\ell}) &=
711: (T_jQ_{ij}) (T_\ell Q_{j\ell}) (T_iQ_{\ell i}),\\
712: (T_kQ_{ik}) (T_\ell Q_{k\ell}) (T_iQ_{\ell i}) &=
713: (T_iQ_{ki}) (T_kQ_{\ell k}) (T_\ell Q_{i \ell}).
714: \end{align}
715: Multiplying the equations we obtain (assuming that the rotation coefficients do
716: not vanish)
717: \begin{equation}
718: (T_kQ_{jk})(T_jQ_{\ell j})(T_\ell Q_{k\ell}) =
719: (T_jQ_{kj})(T_\ell Q_{j\ell})(T_kQ_{\ell k}).
720: \end{equation}
721: \end{proof}
722:
723: \section{Algebro-geometric construction of the C-quadrilateral lattice}
724: \label{sec:AG}
725: In this Section we apply the algebro-geometric approach, well known in the
726: theory of integrable systems \cite{BBEIM}, to the symmetric Darboux equations
727: and to the C-quadrilateral lattice.
728: Similar
729: restrictions on the algebro-geometric data appeared in \cite{DJKM-CKP} in
730: construction of quasi periodic solutions of the CKP hierarchy.
731: For discrete Laplace equations the algebro-geometric techniques were first
732: applied in \cite{Krichever-4p}. In \cite{AKV} the Egorov reduction of the
733: symmetric lattice \cite{Schief-Egorov}
734: was studied by means of the algebro-geometric methods. In 1998 Peter Grinevich
735: isolated from \cite{AKV} these of the algebro-geometric conditions which
736: give rise the symmetric quadrilateral lattices \cite{PG-private}.
737: Below we present that result on algebro-geometric
738: description of the C-(symmetric) quadrilateral lattice within broader context of
739: various discrete Baker--Akhiezer functions.
740:
741: In this Section we work over the field of complex numbers, i.e., $\FF=\CC$, and
742: we discuss the corresponding reality condintions. We mention that
743: many of the results gien below
744: can be transferred to the finite-field case, as in
745: \cite{BD,DBK}.
746:
747:
748: \subsection{The Baker--Akhiezer function of the quadrilateral lattice}
749: Let us consider a compact non-degenerate
750: Riemann surface $\mathcal{R}$ of genus $g$, a
751: non-special divisor $D=P_1 + \dots + P_g$ on $\mathcal{R}$, the
752: $N$ pairs of points $Q_i^\pm\in\mathcal{R}$, and the normalization
753: point $Q_\infty\in\mathcal{R}$. The system $\{ \mathcal{R}, D, Q_i^\pm, Q_\infty
754: \}$ is called \emph{the algebro-geometric data} used in the construction of the
755: quadrilateral lattice.
756: For simplicity we assume that all the
757: points above (including the points of $D$) are distinct.
758:
759: For $m\in\ZZ^N$ we define the meromorphic function
760: $\psi(m):\mathcal{R}\to\CC\PP$ by prescribing
761: its analytical properties:\\
762: (i) as a function on $\mathcal{R}\setminus
763: \cup_{i=1}^N \{ Q_i^\pm \}$ it may
764: may have as singularities only simple poles in points of the divisor
765: $D$;\\
766: (ii) in points $Q_i^+$ (in points $Q_i^-$) it has poles (correspondingly, zeros)
767: of the order $m_i$, where by the pole of the negative
768: order we mean zero of the corresponding order;\\
769: (iii) in the point $Q_\infty$ the function $\psi$ is normalized to $1$.
770:
771: By the standard (see \cite{AKV,BBEIM}) application of the Riemann--Roch theorem,
772: such a function exists and is unique.
773: When $z_i^\pm(P)$ is a local coordinate centered at $Q_i^\pm$ then
774: $\psi$ in a neighbourhood of the point $Q_i^\pm$ is of the form
775: \begin{equation}
776: \psi(m|P) = \left(z_i^\pm(P)\right)^{\mp m_i}\left( \sum_{s=0}^\infty
777: \xi_{s}^{i,\pm}(m)\left(z_i^\pm(P)\right)^{s} \right).
778: \end{equation}
779: %
780: By the standard reasoning in the finite-gap theory one can prove
781: the following result \cite{AKV} which connects the function $\psi$ with the
782: quadrilateral lattice theory.
783: \begin{Th}[\cite{AKV}]
784: For arbitrary point $P\in\mathcal{R}$ the Baker--Akhiezer
785: function $\psi$ satisfies
786: in the variable
787: $m\in\ZZ^N$ the following system of linear equations
788: \begin{equation} \label{eq:Laplace-al-g}
789: \Delta_i\Delta_j\psi(m|P) =
790: \left( T_i \frac{\Delta_j \xi_{0}^{i,+}(m)}{\xi_{0}^{i,+}(m)}\right)
791: \Delta_i\psi(m|P) +
792: \left( T_j \frac{\Delta_i \xi_{0}^{j,+}(m)}{\xi_{0}^{j,+}(m)} \right)
793: \Delta_j\psi(m|P) ,\quad i\ne j.
794: \end{equation}
795: \end{Th}
796: \begin{Cor} \label{cor:measure}
797: Let $d {\boldsymbol{\mu}}(P)=(d\mu_1(P),\dots\, d\mu_M(P))^t$
798: be a vector valued measure on
799: $\mathcal{R}$, then the function $\bx:\ZZ^N\to\CC^M$ given by
800: \begin{equation} \label{eq:x-measure}
801: \bx(m) = \int_{\mathcal{R}} \psi(m|P) d {\boldsymbol{\mu}}(P),
802: \end{equation}
803: defines a quadrilateral lattice (in the complex affine space)
804: with the functions $\xi_0^{i,+}$ as the Lam\'{e}
805: coefficients $H_i$.
806: \end{Cor}
807: \begin{Rem}
808: Similarly one shows hat $\bx$ satisfies the backward Laplace equations
809: \eqref{eq:Laplace-b} with the backward Lam\'e coefficients (proportional to)
810: $\xi_0^{i,-}$; see also Proposition \ref{prop:t-psi}.
811: \end{Rem}
812: In order to have real quadrilateral
813: lattices we must impose on the algebro-geometric data
814: certain reality restrictions.
815: \begin{Cor} \label{cor:reality}
816: Assume that the Riemann surface $\mathcal{R}$
817: allows for an anti-holomorphic involution $\imath$. If
818: \begin{equation}
819: \imath(D) = D, \qquad \imath(Q_i^\pm) = Q_i^\pm, \qquad
820: \imath(Q_\infty) = Q_\infty, \qquad d {\boldsymbol{\mu}}(\imath(P)) =
821: \overline{d {\boldsymbol{\mu}}P)},
822: \end{equation}
823: then the lattice $\bx$, given by equation \eqref{eq:x-measure} is real.
824: Moreover, if the local coordinate systems $z_i^\pm$ are compatible with
825: $\imath$, i.e.,
826: \begin{equation}
827: z_i^\pm(\imath(P)) = \overline{z_i^\pm(P)},
828: \end{equation}
829: then the Lam\'{e} coefficients, as defined above, are real functions.
830: \end{Cor}
831:
832:
833: \subsection{Other Baker--Akhiezer functions}
834: In this Section we define
835: auxiliary Baker-Akhiezer functions which play the role of the forward
836: and backward normalized tangent vectors. Then we define the dual (adjoint)
837: Baker-Akhiezer function of the quadrilateral lattice.
838:
839: \subsubsection{The forward Baker--Akhiezer functions $\psi_i$}
840: Below we consider new functions $\psi_i$
841: whose relation to $\psi$ is analogous to that between $\bX_i$ to $\bx$.
842: Given $m\in\ZZ^N$ define \cite{AKV} the functions
843: $\psi_i(m)$ as meromorphic functions on $\mathcal{R}$ having the
844: following analytic properties:\\
845: (i) as a function on $\mathcal{R}\setminus
846: \cup_{i=1}^N \{Q_i^\pm\}$ it
847: may have as singularities only simple poles in points of the divisor
848: $D$;\\
849: (ii) in points $Q_j^+$ (in points $Q_j^-$) it has poles
850: of the order $m_j+\delta_{ij}$ (correspondingly, zeros of the order $m_j$);\\
851: (iii) in the point $Q_\infty$ the function $\psi$ is equal to $0$.
852:
853: By the Riemann--Roch theorem
854: the space of such functions is one dimensional.
855: By choosing local coordinate $z_i^+$ near
856: $Q_i^+$ the function $\psi_i$ can be made unique by fixing its lowest order term at
857: $Q_i^+$ to one. Then near $Q_j^\pm$ we have the following local expansions
858: \begin{equation}
859: \psi_i(m|P) = \left(z_j^\pm(P)\right)^{\mp m_j}
860: \left( \frac{\delta_{ji}\delta_{\pm +}}{z_j^+(P)} + \sum_{s=0}^\infty
861: \zeta_{i,s}^{j,\pm}(m)\left(z_j^\pm(P)\right)^{s} \right).
862: \end{equation}
863: \begin{Th}[\cite{AKV}] \label{th:psi-i-AG}
864: The functions $\psi_i$ satisfy the equations
865: \begin{align}
866: \label{eq:lin-x-X-AG}
867: \Delta_i\psi(m|P)&= (T_i\xi_{i,0}^{+}(m))\psi_i(m|P),\\
868: \label{eq:lin-X-AG}
869: \Delta_j\psi_i(m|P) &= (T_j\zeta_{i,0}^{j,+}(m))\psi_i(m|P),\qquad j\ne i,
870: \end{align}
871: whose expansion at $Q_k^+$, gives
872: \begin{align}
873: \Delta_i\xi_{k,0}^{+} & = (T_i\xi_{i,0}^{+}(m))\zeta_{i,0}^{k,+},
874: \qquad k\ne i,\\
875: \Delta_j\zeta_{i,0}^{k,+} & = (T_j\zeta_{i,0}^{j,+})
876: \zeta_{j,0}^{k,+} , \qquad k\ne j,
877: \end{align}
878: and allows for the identification
879: \begin{equation}
880: Q_{ij}(m) = \zeta_{i,0}^{j,+}(m).
881: \end{equation}
882: \end{Th}
883: \begin{Cor}
884: In notation of Corollary \ref{cor:measure} we have
885: \begin{equation} \label{eq:X-measure}
886: \bX_i(m) = \int_{\mathcal{R}} \psi_i(m|P) d {\boldsymbol{\mu}}(P).
887: \end{equation}
888: \end{Cor}
889: \begin{Rem}
890: Notice that different choices of local coordinates $z_i^+$ correspond to
891: rescaling of the forward data in agreement with equation \eqref{eq:forward-scaling}.
892: \end{Rem}
893: \subsubsection{The backward Baker--Akhiezer functions $\tilde\psi_i$}
894: We define the corresponding algebro-geometric
895: analog of the backward normalized tangent vectors
896: $\tbX_i$.
897: Given $m\in\ZZ^N$ define the functions
898: $\tilde\psi_i(m)$ as meromorphic functions on $\mathcal{R}$ having the
899: following analytic properties:\\
900: (i) as a function on $\mathcal{R}\setminus
901: \cup_{i=1}^N \{Q_i^\pm\}$ it may
902: may have as singularities only simple poles in points of the divisor
903: $D$;\\
904: (ii) in points $Q_j^+$ (in points $Q_j^-$) it has poles
905: of the order $m_j$ (correspondingly, zeros of the order $m_j-\delta_{ij}$);\\
906: (iii) in the point $Q_\infty$ the function $\psi$ is equal to $0$.
907:
908: By the Riemann--Roch theorem the space of such functions is one dimensional.
909: By choosing local coordinates $z_i^-$ near
910: $Q_i^-$ the function $\tilde\psi_i$ can be made unique by fixing its lowest order
911: term at
912: $Q_i^-$ to one. Then near $Q_j^\pm$ we have the following local expansions
913: \begin{equation} \label{eq:exp-t-psi_i}
914: \tilde\psi_i(m|P) = \left(z_j^\pm(P)\right)^{\mp m_j}
915: \left( \frac{\delta_{ji}\delta_{\pm -}}{z_j^-(P)} + \sum_{s=0}^\infty
916: \tilde\zeta_{i,s}^{j,\pm}(m)\left(z_j^\pm(P)\right)^{s} \right).
917: \end{equation}
918: By the standard methods \cite{BBEIM} one can prove the following
919: analog of Theorem \ref{th:psi-i-AG}
920: \begin{Prop} \label{prop:t-psi}
921: The functions $\psi$ and $\tilde\psi_i$ are connected by the formulas
922: \begin{align}
923: \D_i\psi(m|P) & = -\left(T_i\tilde\psi_i(m|P)\right)\xi_{0}^{i,-}(m),
924: \label{eq:Ti-tpsi} \\
925: \D_j\tilde\psi_i(m|P) & = -\left(T_j\tilde\psi_j(m|P)\right)
926: \tilde\zeta_{i,0}^{j,-}(m) , \label{eq:Tj-tpsi-i}
927: \qquad j\ne i,
928: \end{align}
929: whose expansion at $Q_k^-$, gives
930: \begin{align}
931: \D_i\xi_{0}^{k,-} & = -\left(T_i\tilde\zeta_{i,0}^{k,-}\right)\xi_{0}^{i,-},
932: \qquad k\ne i,\\
933: \D_j \tilde\zeta_{i,0}^{k,-} &= - (T_j \tilde\zeta_{j,0}^{k,-})
934: \tilde\zeta_{i,0}^{j,-}, \qquad k\ne j,
935: \end{align}
936: and allows for the identification
937: \begin{equation}
938: \tH_i(m) = - \xi_{0}^{i,-}(m), \qquad \tQ_{ij}(m) = -\tilde\zeta_{j,0}^{i,-}(m).
939: \end{equation}
940: \end{Prop}
941: \begin{Cor}
942: In notation of Corollary \ref{cor:measure} we have
943: \begin{equation} \label{eq:tX-measure}
944: \tbX_i(m) = \int_{\mathcal{R}} \tilde\psi_i(m|P) d {\boldsymbol{\mu}}(P).
945: \end{equation}
946: Moreover, by comparing the analytical properties the functions $\psi_i$ and
947: $\tilde\psi_i$ we obtain
948: \begin{align}
949: \psi_i(m|P) & = \left(T_i\tilde\psi_i(m|P)\right)\zeta_{i,0}^{i,-}(m) ,\\
950: T_i\tilde\psi_i(m|P) & = \left(T_i\tilde\zeta_{i,0}^{i,+}(m)\right) \psi_i(m|P),
951: \end{align}
952: which allows for the identification
953: \begin{equation}
954: \rho_i(m) = \zeta_{i,0}^{i,-}(m) = \frac{1}{T_i\tilde\zeta_{i,0}^{i,+}}.
955: \end{equation}
956: \end{Cor}
957: \begin{Rem}
958: Notice that different choices of local coordinates $z_i^-$ correspond to
959: rescaling of the backward data given by equation \eqref{eq:backward-scaling}.
960: \end{Rem}
961: \begin{Rem}
962: We are not concerned here about explicit theta-function formulas for the
963: Baker--Akhiezer functions and related potentials, see however \cite{AKV}. In
964: particular, the $\tau$-function of the quadrilateral lattice is, essentially
965: \cite{AD-Chicago},
966: the Riemann theta function.
967: \end{Rem}
968:
969: \subsubsection{The dual Baker--Akhiezer functions}
970: In definition of the dual (adjoint) Baker--Akhiezer function of the
971: quadrilateral lattice we use the idea applied in \cite{DKJM} to construction of
972: the adjoint Baker--Akhiezer function of the KP hierarchy.
973:
974: Denote by $\omega_\infty$ the meromorphic differential
975: with the only singularity being the second order pole at
976: $Q_\infty$, and whose holomorphic part is normalized
977: by vanishing of $\omega_\infty$ at
978: points of the divisor $D$
979: \begin{equation}
980: \omega_\infty(P_i)=0, \quad i=1,\dots,g.
981: \end{equation}
982: \begin{Rem}
983: By choosing a coordinate
984: system $z_\infty(P)$ centered at $Q_\infty$
985: the differential $\omega_\infty$ can be made unique by
986: fixing its singular part in $z_\infty(P)$ as
987: \begin{equation}
988: \omega_\infty(P) = \left( \frac{1}{z_\infty(P)^2} + O(1)\right)dz_\infty(P),
989: \end{equation}
990: but we will not use that in the sequel.
991: \end{Rem}
992: Denote by $D^*$ the divisor of other $g$ zeros of $\omega_\infty$, and use it
993: to define the dual Baker--Akhiezer function $\psi^*$
994: exchanging also the role of the points $Q_i^+$ and $Q_i^-$:\\
995: (i) as a function on $\mathcal{R}\setminus
996: \cup_{i=1}^N \{ Q_i^\pm\}$ it may
997: may have as singularities only simple poles in points of the divisor
998: $D^*$;\\
999: (ii) in points $Q_i^+$ (in points $Q_i^-$) it has zeros (correspondingly, poles)
1000: of the order $m_i$;\\
1001: (iii) in the point $Q_\infty$ the function $\psi^*$ is normalized to $1$.
1002:
1003: Using the Riemann--Roch theorem one can show that such function
1004: $\psi^*(m|P)$ exists and is unique.
1005: In a neighbourhood of the point $Q_i^\pm$ it is of the form
1006: \begin{equation}
1007: \psi^*(m|P) = \left(z_i^\pm(P)\right)^{\pm m_i}\left( \sum_{s=0}^\infty
1008: \xi_{s}^{*i,\pm}(m)\left(z_i^\pm(P)\right)^{s} \right).
1009: \end{equation}
1010: Using the similar procedure like in the previous section it can be shown that the
1011: dual function $\psi^*$ satisfies the Laplace equations with Lam\'{e} coefficients
1012: $\xi_{0}^{*i,-}$, and it satisfies the backward Laplace equations with the backward
1013: Lam\'{e} coefficients
1014: $\xi_{0}^{*i,+}$.
1015: \begin{Rem}
1016: The meromorphic differential form
1017: \begin{equation}
1018: \omega = \psi \psi^* \omega_\infty
1019: \end{equation}
1020: is singular only at $Q_\infty$ with the
1021: singularity being the second order pole.
1022: By the residue theorem the integral of $\omega$ around a closed contour around
1023: $Q_\infty$ vanishes, which is the quadrilateral lattice counterpart of the
1024: celebrated bilinear identity \cite{DKJM} on the
1025: algebro-geometric level.
1026: \end{Rem}
1027:
1028:
1029: In analogy to the Baker--Akhiezer functions $\psi_i$ and $\tilde\psi_i$
1030: we may define the corresponding
1031: dual Baker--Akhiezer functions. In the sequel we will need the analog of
1032: $\tilde\psi_i$, which is defined as follows.
1033: Given $m\in\ZZ^N$ define the functions
1034: $\psi_i^*(m)$ as meromorphic functions on $\mathcal{R}$ having the
1035: following analytic properties:\\
1036: (i) as a function on $\mathcal{R}\setminus
1037: \cup_{i=1}^N Q_i^\pm$ it may
1038: may have as singularities only simple poles in points of the divisor
1039: $D^*$;\\
1040: (ii) in points $Q_j^+$ (in points $Q_j^-$) it has zeros
1041: of the order $m_j-\delta_{ij}$ (correspondingly, poles of the order $m_j$);\\
1042: (iii) in the point $Q_\infty$ the function $\psi^*_i$ is equal to $0$.
1043:
1044: By the Riemann--Roch the space of such functions is one dimensional.
1045: By choosing local coordinates $z_i^+$ near
1046: $Q_i^+$ the function $\psi_i^*$ can be made unique by fixing its lowest order term at
1047: $Q_i^+$ to one. Then near $Q_j^\pm$ we have the following local expansions
1048: \begin{equation} \label{eq:exp-psi_i^*}
1049: \psi_i^*(m|P) = \left(z_j^\pm(P)\right)^{\pm m_j}
1050: \left( \frac{\delta_{ji}\delta_{\pm +}}{z_j^+(P)} + \sum_{s=0}^\infty
1051: \zeta_{i,s}^{*j,\pm}(m)\left(z_j^\pm(P)\right)^{s} \right).
1052: \end{equation}
1053: As before one can study the relation between $\psi$ and $\psi_i^*$.
1054: However, we will be interested in the following connection formulas
1055: between the
1056: $\zeta$-coefficients of both functions.
1057: \begin{Prop} \label{prop:omega-z-z}
1058: Denote by $a_i^\pm$ the first coefficients of the expansion of
1059: $\omega_\infty$ near points $Q_i^\pm$
1060: \begin{equation}
1061: \omega_\infty(P) = [a_i^\pm + O(z_i^\pm(P))]dz_i^\pm(P),
1062: \end{equation}
1063: then by vanishing of the sum of residues of the differential
1064: $\psi_i(m|P)\psi_j^*(m|P)\omega_\infty$ we have
1065: \begin{equation} \label{eq:omega-z-z}
1066: a_i^\pm\zeta_{j,0}^{*i,\pm}(m) + a_j^\pm\zeta_{i,0}^{j,\pm}(m) = 0, \qquad
1067: i \neq j.
1068: \end{equation}
1069: \end{Prop}
1070:
1071: \subsection{The algebro-geometric C-quadrilateral lattices}
1072: Finally, we show that under certain restrictions on the algebro-geometric data
1073: the finite-gap construction gives C-reduced quadrilateral lattice. This type of
1074: restrictions appeared in \cite{DJKM-CKP} in construction of quasi-periodic
1075: solutions of the CKP hierarchy.
1076: \begin{Prop} \label{prop:CQL-alg-geom}
1077: Assume that $\mathcal{R}$ is equipped with the holomorphic involution
1078: $\sigma:\mathcal{R}\to\mathcal{R}$ such that
1079: \begin{equation}
1080: \sigma(D^*) = D, \qquad \sigma(Q_i^\pm) = Q_i^\mp, \qquad
1081: \sigma(Q_\infty) = Q_\infty,
1082: \end{equation}
1083: then
1084: \begin{align}
1085: \label{eq:psi-sigma}
1086: \psi\circ\sigma & = \psi^*,\\
1087: \label{eq:psi_i-sigma}
1088: \tilde\psi_i\circ\sigma & = c_i \psi^*_i, \qquad c_i\in\CC.
1089: \end{align}
1090: \end{Prop}
1091: \begin{proof}
1092: In the standard way we compare analytic properties of both sides of each
1093: equation. The
1094: function $\psi\circ\sigma$ has the following analytic properties:\\
1095: (i) as a function on $\mathcal{R}\setminus
1096: \cup_{i=1}^N \{Q_i^\pm\}$ it
1097: may have as singularities only simple poles in points of the divisor
1098: $D^*$;\\
1099: (ii) in points $Q_i^+$ (in points $Q_i^-$) it has zeros (correspondingly, poles)
1100: of the order $m_i$;\\
1101: (iii) in the point $Q_\infty$ the function $\psi\circ\sigma$ is normalized to
1102: $1$.\\
1103: Comparison with the analytic properties of $\psi^*$ and the Riemann--Roch
1104: theorem gives equation \eqref{eq:psi-sigma}.
1105:
1106: Let us describe the analytic properties of the superposition
1107: $\tilde\psi_i\circ\sigma$:\\
1108: (i) as a function on $\mathcal{R}\setminus
1109: \cup_{i=1}^N \{Q_i^\pm\}$ it may
1110: may have as singularities only simple poles in points of the divisor
1111: $D^*$;\\
1112: (ii) in points $Q_j^+$ (in points $Q_j^-$) it has zeros
1113: of the order $m_j-\delta_{ij}$ (correspondingly, poles of the order $m_j$);\\
1114: (iii) in the point $Q_\infty$ the function
1115: $\tilde\psi_i\circ\sigma$ is equal to $0$.\\
1116: Therefore the function $\tilde\psi_i\circ\sigma$ must be proportional to
1117: $\psi^*_i$.
1118: \end{proof}
1119: \begin{Cor}
1120: Notice that
1121: when the local coordinates, which fix normalization of the functions,
1122: are chosen in agreement with the involution $\sigma$
1123: \begin{equation} \label{eq:zi-zi+red}
1124: z_i^-(\sigma(P)) = z_i^+(P),
1125: \end{equation}
1126: then the proportionality in equation
1127: \eqref{eq:psi_i-sigma} becomes equality (i.e., $c_i=1$). Moreover, under such
1128: conditions the
1129: expansions \eqref{eq:exp-t-psi_i} and \eqref{eq:exp-psi_i^*} give
1130: \begin{equation} \label{eq:tz-*z}
1131: \tilde\zeta_{i,s}^{j,\pm}(m) = \zeta_{i,s}^{*j,\mp}(m).
1132: \end{equation}
1133: \end{Cor}
1134: \begin{Th}[\cite{PG-private}]
1135: Under assumptions of Proposition \ref{prop:CQL-alg-geom}, the quadrilateral
1136: lattice constructed according to Corollary \ref{cor:measure} is subject to
1137: the C-(symmetric) reduction.
1138: \end{Th}
1139: \begin{proof}
1140: Assume for a time being that the local coordinates $z_i^+$ are chosen in such a
1141: way that the first coefficients $a_i^+$ of the expansion of
1142: $\omega_\infty$ near points $Q_i^+$ are equal (see Proposition
1143: \ref{prop:omega-z-z}), and the local coordinates $z_i^-$ are chosen according to
1144: equation \eqref{eq:zi-zi+red}; we may think of this special choice as using the
1145: allowed freedom \eqref{eq:forward-scaling}-\eqref{eq:backward-scaling} in
1146: definition of the backward and forward data.
1147: Then equations \eqref{eq:omega-z-z} and
1148: \eqref{eq:tz-*z}
1149: imply that
1150: \begin{equation}
1151: \zeta_{i,0}^{j,+}(m) = -\tilde\zeta_{j,0}^{i,-}(m) , \qquad \text{or} \quad
1152: Q_{ij} = \tilde{Q}_{ij}.
1153: \end{equation}
1154: \end{proof}
1155:
1156: \section{Transformation of the C-quadrilateral lattice}
1157: \label{sec:C-fund}
1158: We introduce geometrically the C-reduction
1159: of the fundamental transformation of the quadrilateral lattice. We also
1160: connect this definition with earlier algebraic results of \cite{MM}.
1161: Then we prove the
1162: corresponding permutability theorem for this transformation.
1163: \subsection{The vectorial fundamental
1164: transformation of the quadrilateral lattice}
1165: Let us first recall some basic facts concerning the vectorial fundamental
1166: transformation of the quadrilateral lattice.
1167: Geometrically, the (scalar) fundamental transformation is the relation
1168: between two
1169: quadrilateral
1170: lattices $x$ and $x^\prime$ such that for each direction $i$ the
1171: points $x$, $x^\prime$, $x_{(i)}$ and $x^\prime_{(i)}$ are coplanar.
1172:
1173: We
1174: present below the algebraic description of its vectorial extension (see
1175: \cite{MDS,TQL,MM} for details) in the affine formalism.
1176: Given the solution $\boldsymbol{Y}_i:\mathbb{Z}^N\to\FF^K$,
1177: of the linear system \eqref{eq:lin-X}, and given the solution
1178: $\boldsymbol{Y}^*_i:\mathbb{Z}^N\to(\FF^K)^*$, of the linear system
1179: \eqref{eq:lin-H}. These allow to
1180: construct the linear operator valued potential
1181: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*):
1182: \mathbb{Z}^N\to M^K_K(\FF)$,
1183: defined by
1184: \begin{equation} \label{eq:Omega-Y-Y}
1185: \Delta_i \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1186: \boldsymbol{Y}_i \otimes T_i\boldsymbol{Y}^*_i,
1187: \qquad i = 1,\dots , N;
1188: \end{equation}
1189: similarly, one defines
1190: $\boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*):
1191: \mathbb{Z}^N\to M^M_K(\FF)$ and
1192: $\boldsymbol{\Omega}(\boldsymbol{Y},H):
1193: \mathbb{Z}^N\to \FF^K$ by
1194: \begin{align} \label{eq:Omega-X-Y}
1195: \Delta_i \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*) & =
1196: \boldsymbol{X}_i \otimes T_i \boldsymbol{Y}^*_i, \\
1197: \Delta_i\boldsymbol{\Omega}(\boldsymbol{Y},H) & =
1198: \boldsymbol{Y}_i \otimes T_i H_i.
1199: \end{align}
1200: \begin{Prop}
1201: If $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)$ is invertible then
1202: the vector function $\bx^\prime:\ZZ^N\to\FF^M$ given by
1203: \begin{equation} \label{eq:fund-vect}
1204: \bx^\prime = \bx -
1205: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*)
1206: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1207: \boldsymbol{\Omega}(\boldsymbol{Y},H),
1208: \end{equation}
1209: represents a quadrilateral lattice (the fundamental transform of $x$),
1210: whose Lam\'e coefficients $H_i^\prime$, normalized tangent vectors
1211: $\bX_i^\prime$ and rotation coefficients $Q_{ij}^\prime$ are given by
1212: \begin{align}
1213: \label{eq:fund-vect-H}
1214: H_i^\prime &= H_i -
1215: \boldsymbol{Y}^*_i
1216: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1217: \boldsymbol{\Omega}(\boldsymbol{Y},H),\\
1218: \label{eq:fund-vect-X}
1219: \bX^\prime_i & = \bX_i -
1220: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*)
1221: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1222: \boldsymbol{Y}_i,\\
1223: \label{eq:fund-vect-Q}
1224: Q_{ij}^\prime & = Q_{ij} -
1225: \boldsymbol{Y}^*_j
1226: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1227: \boldsymbol{Y}_i.
1228: \end{align}
1229: Moreover \cite{MM}, the connection coefficients $\rho_i$
1230: and the $\tau$-function
1231: transform according
1232: to
1233: \begin{align}
1234: \label{eq:fund-vect-rho}
1235: \rho^\prime_i & = \rho_i(1 + T_i \boldsymbol{Y}^*_i
1236: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^{-1}
1237: \boldsymbol{Y}_i),\\
1238: \tau^\prime & = \tau \det\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*).
1239: \end{align}
1240: \end{Prop}
1241: The vectorial
1242: fundamental transformation can be considered as superposition of
1243: $K$ (scalar) fundamental transformations; on intermediate stages
1244: the rest of the transformation data should be suitably transformed as well.
1245: Such a description contains already the principle of permutability of such
1246: transformations, which follows from the following observation~\cite{TQL}.
1247: \begin{Prop}
1248: Assume the following splitting of the data of the vectorial fundamental
1249: transformation
1250: \begin{equation}
1251: \boldsymbol{Y}_i = \left( \begin{array}{c}
1252: \boldsymbol{Y}_i^a \\ \boldsymbol{Y}_i^b \end{array} \right),\qquad
1253: \boldsymbol{Y}_i^* = \left( \begin{array}{cc}
1254: \boldsymbol{Y}_{ai}^{*} & \boldsymbol{Y}_{b i}^{*} \end{array} \right),
1255: \end{equation}
1256: associated with the partition $\FF^K = \FF^{K_a} \oplus \FF^{K_b}$,
1257: which implies the following splitting of the potentials
1258: \begin{equation} \label{eq:split-fund-1}
1259: \boldsymbol{\Omega}(\boldsymbol{Y},H) = \left( \begin{array}{c}
1260: \boldsymbol{\Omega}(\boldsymbol{Y}^a,H) \\
1261: \boldsymbol{\Omega}(\boldsymbol{Y}^b,H) \end{array} \right), \qquad
1262: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) = \left( \begin{array}{cc}
1263: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_a^*) &
1264: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_b^*) \\
1265: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}_a^*) &
1266: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}_b^*)\end{array} \right),
1267: \end{equation}
1268: \begin{equation} \label{eq:split-fund-2}
1269: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*) = \left( \begin{array}{cc}
1270: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_a^*) &
1271: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_b^*)\end{array} \right).
1272: \end{equation}
1273: Then the vectorial fundamental transformation is equivalent to the following
1274: superposition of vectorial fundamental transformations:\\
1275: 1) Transformation $\bx\to\bx^{\{a\}}$ with the data
1276: $\boldsymbol{Y}_i^a$, $\boldsymbol{Y}_{ai}^*$ and the corresponding
1277: potentials
1278: $\boldsymbol{\Omega}(\boldsymbol{Y}^a,H)$,
1279: $\boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}_a^*)$,
1280: $\boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}_a^*)$
1281: \begin{align}
1282: \label{eq:fund-vect-a}
1283: \bx^{\{a\}} & = \bx -
1284: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1285: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1286: \boldsymbol{\Omega}(\boldsymbol{Y}^a,H),\\
1287: \boldsymbol{X}_i^{\{a\}} & = \boldsymbol{X}_i -
1288: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1289: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1290: \boldsymbol{Y}^a_i,
1291: \\
1292: H_i^{\{a\}} & = H_i - \boldsymbol{Y}^*_{i a}
1293: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1294: \boldsymbol{\Omega}(\boldsymbol{Y}^a,H).
1295: \end{align}
1296: 2) Application on the result the vectorial fundamental transformation with the
1297: transformed data
1298: \begin{align}
1299: {\boldsymbol{Y}}_i^{b\{a\}} & = \boldsymbol{Y}_i^b -
1300: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1301: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1302: \boldsymbol{Y}^a_i,
1303: \\
1304: {\boldsymbol{Y}}_{i b}^{*\{a\}} & = \boldsymbol{Y}_{i b}^* -
1305: \boldsymbol{Y}^*_{i a}
1306: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1307: \boldsymbol{\Omega}(\boldsymbol{Y}^a, \boldsymbol{Y}_{b}^*),
1308: \end{align}
1309: and potentials
1310: \begin{align}
1311: {\boldsymbol{\Omega}}(\boldsymbol{Y}^b,H)^{\{a\}} & =
1312: \boldsymbol{\Omega}(\boldsymbol{Y}^b,H) -
1313: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1314: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1315: \boldsymbol{\Omega}(\boldsymbol{Y}^a,H)=
1316: \boldsymbol{\Omega}({\boldsymbol{Y}}^{b\{a\}},H^{\{a\}}),
1317: \\
1318: {\boldsymbol{\Omega}}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b)^{\{a\}} & =
1319: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b) -
1320: \boldsymbol{\Omega}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_a)
1321: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1322: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_b)=
1323: \boldsymbol{\Omega}({\boldsymbol{Y}}^{b\{a\}},{\boldsymbol{Y}}_b^{*\{a\}}),
1324: \\
1325: {\boldsymbol{\Omega}}(\boldsymbol{X},\boldsymbol{Y}^*_b)^{\{a\}} & =
1326: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_b) -
1327: \boldsymbol{\Omega}(\boldsymbol{X},\boldsymbol{Y}^*_a)
1328: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_a)^{-1}
1329: \boldsymbol{\Omega}(\boldsymbol{Y}^a,\boldsymbol{Y}^*_b)=
1330: \boldsymbol{\Omega}({\boldsymbol{X}}^{\{a\}},{\boldsymbol{Y}}_b^{*\{a\}}),
1331: \label{eq:fund-vect-potentials-slit}
1332: \end{align}
1333: i.e.,
1334: \begin{equation} \label{eq:fund-vect-a-b}
1335: \bx^\prime = \bx^{\{a,b\}} = \bx^{\{a\}} -
1336: {\boldsymbol{\Omega}}(\boldsymbol{X},\boldsymbol{Y}^*_b)^{\{a\}}
1337: [{\boldsymbol{\Omega}}(\boldsymbol{Y}^b,\boldsymbol{Y}^*_b)^{\{a\}}]^{-1}
1338: {\boldsymbol{\Omega}}(\boldsymbol{Y}^b,H)^{\{a\}}.
1339: \end{equation}
1340: \end{Prop}
1341: \begin{Rem}
1342: The same result $\bx^\prime = \bx^{\{a,b\}}={\bx}^{\{b,a\}}$
1343: is obtained exchanging the order of transformations, exchanging also the indices
1344: $a$ and $b$ in formulas
1345: \eqref{eq:fund-vect-a}-\eqref{eq:fund-vect-a-b}.
1346: \end{Rem}
1347:
1348:
1349: \subsection{The CQL (symmetric) reduction of the fundamental transformation}
1350: In this section we describe restrictions on the data of the fundamental
1351: transformation in order to preserve the reduction from QL to CQL.
1352: As usually (see, for example \cite{TQL,q-red,BQL}) a reduction of the
1353: fundamental transformation for a special quadrilateral lattice
1354: mimics the geometric properties of the lattice. Because the basic geometric
1355: property of the (scalar) fundamental transformation can be interpreted as
1356: construction of a "new level" of the quadrilateral lattice, then it is natural
1357: to define the reduced transformation in a similar spirit. Our definition of
1358: \emph{the CQL reduction of the fundamental transformation} is therefore based on
1359: the following observation.
1360: \begin{Lem} \label{lem:BQL-fund}
1361: Given quadrilateral lattice $x:\ZZ^N\to\AAf^M$ and its fundamental transform
1362: $x^\prime$ constructed under additional assumption that for any point
1363: $x$ of the lattice and any pair $i,j$ of different directions, the hexahedra
1364: with basic vertices
1365: $x$, $T_i x$, $T_j x$ and $x^\prime$ satisfy the C-reduction condition.
1366: Then both the starting lattice $x:\ZZ^N\to\AAf^M$ and its
1367: transform $x^\prime:\ZZ^N\to\AAf^M$ are C-quadrilateral lattices.
1368: \end{Lem}
1369: \begin{proof}
1370: As $N\geq 3$, by
1371: Lemma~\ref{lem:C-consistency} we have that also the hexahedra
1372: with basic vertices
1373: $x$, $T_i x$, $T_j x$ and $T_k x$, with $i,j,k$ distinct,
1374: satisfy the C-reduction condition. The similar statement for the transformed
1375: lattice is a consequence of
1376: the $4$-dimensional consistency of the CQL lattice.
1377: \end{proof}
1378: \begin{Def}
1379: The fundamental transform $x^\prime$ of a C-quadrilateral lattice
1380: $x:\ZZ^N\to\AAf^M$
1381: constructed under additional assumption that for any point $x$ of the lattice
1382: and any pair $i,j$ of different directions, the hexahedra
1383: with basic vertices
1384: $x$, $T_i x$, $T_j x$ and $x^\prime$ satisfy the C-reduction condition,
1385: is called
1386: \emph{the CQL reduction} of the fundamental transformation.
1387: \end{Def}
1388: The following result gives the corresponding
1389: restriction of the data of the (scalar) fundamental transformation.
1390: \begin{Prop} \label{prop:CQL-fund}
1391: Let $x$ be a C-quadrilateral lattice with rotation coefficients
1392: satisfying constraint \eqref{eq:symm-constr}, and $x^\prime$ its
1393: C-reduced fundamental transform. Then there exists a constant $c$ such that
1394: the data
1395: $Y_i:\ZZ^N\to \FF$ and $Y_i^*:\ZZ^N\to\FF$ of the transformation are connected
1396: by relation
1397: \begin{equation}
1398: Y_i = c \, \rho_i T_i Y_i^*, \qquad i = 1,\dots , N, \qquad c\in\FF.
1399: \end{equation}
1400: \end{Prop}
1401: \begin{proof}
1402: We start from considerations similar to that of proof of
1403: Proposition~\ref{prop:Q3-CQL}. The idea is to interpret
1404: the fundamental transformation as construction of a new
1405: level of the quadrilateral lattice. The potential
1406: $\boldsymbol{\Omega}(\boldsymbol{X},Y^*)$, called also the Combescure vector of
1407: the transformation, serves as the normalized tangent vector \cite{TQL},
1408: which we denote by
1409: $\bX_\prime$, in the transformation direction ``$\prime$".
1410:
1411: Denote by $\bt^\prime_{ij}$ ($i,j$ are
1412: distinct) the direction vector of the common line of the planes
1413: $\langle \bx, T_i \bx, T_j\bx \rangle$ and
1414: $\langle \bx^\prime, T_i \bx^\prime, T_j \bx^\prime \rangle$.
1415: It must be therefore decomposed in
1416: the basis $\{ \bX_i,\bX_j \}$ and in the basis
1417: $\{ \bX_i^\prime,\bX_j^\prime \}$.
1418: Assuming its decomposition in the second basis we get
1419: \begin{equation*}
1420: \bt^\prime_{ij} = a \bX^\prime_i + b \bX_j^\prime = a \bX_i + b \bX_j -
1421: \bX_\prime(a Y_i + b Y_j ) \frac{1}{\boldsymbol{\Omega}(Y,Y^*)},
1422: \end{equation*}
1423: where we have used the transformation equation \eqref{eq:fund-vect-X}.
1424: Because the coefficient
1425: in front of $\bX_\prime$ must vanish, the vector can be therefore
1426: chosen as
1427: \begin{equation}
1428: \bt^\prime_{ij} = Y_j \bX_i - Y_i \bX_j .
1429: \end{equation}
1430:
1431: Similarly, denote by $\bt^j_{\prime i}$
1432: the direction vector of the intersection line of the plane
1433: $\langle \bx, \bx^\prime, T_i\bx \rangle$ with
1434: $\langle T_j \bx, T_j \bx^\prime, T_i T_j \bx\rangle$.
1435: It must be therefore decomposed in
1436: the basis $\{ \bX_\prime, \bX_i \}$ and in the basis
1437: $\{ T_j \bX_\prime, T_j\bX_i \}$.
1438: By using equations \eqref{eq:Omega-X-Y} and \eqref{eq:lin-X} we can
1439: choose the vector as
1440: \begin{equation}
1441: \bt^\prime_{ij} = (T_j Q_{ij}) T_j \bX_\prime - (T_j Y_j^*) T_j\bX_i =
1442: (T_j Q_{ij}) \bX_\prime - (T_j Y_j^*) \bX_i.
1443: \end{equation}
1444: Because
1445: \begin{equation*}
1446: \bt^j_{\prime i}\wedge\bt^i_{\prime j}\wedge\bt^\prime_{ij} =
1447: \left( (T_jQ_{ij}) Y_j (T_iY^*_i) -
1448: (T_iQ_{ji}) Y_i (T_j Y^*_j)
1449: \right) \bX_\prime \wedge \bX_i \wedge \bX_j \qquad i\ne j,
1450: \end{equation*}
1451: then the C-reduction condition of the $i$, $j$, $\prime$ hexahedron takes the
1452: form of equation \eqref{eq:CQL-constr}
1453: \begin{equation}
1454: (T_i Q_{ji}) Y_j (T_iY^*_i)=
1455: (T_j Q_{ij})Y_i (T_j Y^*_j).
1456: \end{equation}
1457: Making use of condition \eqref{eq:symm-constr} we obtain that
1458: \begin{equation}
1459: \rho_j (T_j Y^*_j) Y_i = \rho_i (T_i Y^*_i) Y_j .
1460: \end{equation}
1461: Then we use Lemma \ref{lem-ry*-Y}, which states that
1462: $\rho_i T_i Y_i^*$ satisfies the same linear problem
1463: \eqref{eq:lin-X} as $Y_i$ does.
1464: Finally, application of
1465: the following Lemma concludes the proof.
1466: \end{proof}
1467: \begin{Lem}
1468: Any two scalar solutions $Y_i$ and $\tilde{Y}_i$ of the linear problem
1469: \eqref{eq:lin-X}, which satisfy the constraint
1470: \begin{equation} \label{eq:hY-Y}
1471: \hat{Y}_j Y_i = \hat{Y}_i Y_j,
1472: \end{equation}
1473: must be proportional.
1474: \end{Lem}
1475: \begin{proof}[Proof of the Lemma]
1476: Assume that none of the solutions is trivial (then the proportionality constant
1477: would be zero) and define
1478: \begin{equation}
1479: r_i = \frac{\hat{Y}_i}{Y_i}.
1480: \end{equation}
1481: By equation \eqref{eq:lin-X} we find
1482: \begin{equation}
1483: \Delta_j r_i = \frac{(T_j Q_{ij})(\hat{Y}_j Y_i - \hat{Y}_i Y_j)}
1484: {Y_i (T_j Y_i)}, \qquad j \ne i,
1485: \end{equation}
1486: which vanishes because of the assumption \eqref{eq:hY-Y}, therefore $r_i$ may
1487: depend on the variable $m_i$ only. Inserting then $\hat{Y}_i = r_i Y_i$ into
1488: equation \eqref{eq:hY-Y} we obtain $r_i = r_j$, which implies that all the $r$'s
1489: are equal to the same constant.
1490: \end{proof}
1491: \begin{Rem}
1492: In the non-degenerate situation, i.e. $c\ne 0$, which we assume in the sequel,
1493: we can put $c=1$, because (up to initial value)
1494: $\boldsymbol{\Omega}(cY,{Y}^*) = c\boldsymbol{\Omega}(Y,{Y}^*)$,
1495: and
1496: $\boldsymbol{\Omega}(cY,H)= c\boldsymbol{\Omega}(Y,H)$, and the final result
1497: \eqref{eq:fund-vect} is independent of $c$.
1498: \end{Rem}
1499: \subsection{Permutability theorem for
1500: the CQL reduction of the fundamental
1501: transformation}
1502: In this section we study restrictions of the data of the vectorial fundamental
1503: transformation, which are compatible with the CQL reduction. In that part we
1504: follow the corresponding results of \cite{MM} (see also Proposition 4.9 of
1505: \cite{DS-sym}).
1506: Then we show the
1507: corresponding permutability property of the transformation.
1508:
1509: \begin{Prop}[\cite{MM}]
1510: Given a solution $\bY_i^*$ of the adjoint linear problem \eqref{eq:lin-H} for
1511: the C-quad\-ri\-la\-te\-ral lattice whose rotation coefficients satisfy
1512: the CQL constraint \eqref{eq:symm-constr} then
1513: \begin{equation} \label{eq:C-ft-YY}
1514: \bY_i = \rho_i (T_i \bY_i^*)^t
1515: \end{equation}
1516: provides a vectorial solution of the linear problem
1517: \eqref{eq:lin-X}, and the corresponding potential
1518: $\boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) $ allows for the following
1519: constraint
1520: \begin{equation}\label{eq:C-ft-O}
1521: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^t=
1522: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*).
1523: \end{equation}
1524: With such a data the transformed lattice $\bx^\prime$ given by
1525: \eqref{eq:fund-vect} is C-quadrilateral lattice as well.
1526: \end{Prop}
1527:
1528: \begin{Rem}
1529: In \cite{MM}, instead of relation \eqref{eq:C-ft-YY} it was used more general
1530: relation
1531: \begin{equation}
1532: \hat\bY_i = A \rho_i (T_i \bY_i^*)^t,
1533: \end{equation}
1534: where $A$ is an arbitrary linear operator. Then also the constraint
1535: \eqref{eq:C-ft-O} had to be replaced by
1536: \begin{equation} \label{eq:C-ft-AO}
1537: A\boldsymbol{\Omega}(\hat{\boldsymbol{Y}},\boldsymbol{Y}^*)^t=
1538: \boldsymbol{\Omega}(\hat{\boldsymbol{Y}},\boldsymbol{Y}^*) A^t,
1539: \end{equation}
1540: which is however equivalent, up to initial data,
1541: to \eqref{eq:C-ft-O} due to
1542: \begin{equation}
1543: A\boldsymbol{\Omega}(A\boldsymbol{Y},\boldsymbol{Y}^*)^t -
1544: \boldsymbol{\Omega}(A\boldsymbol{Y},\boldsymbol{Y}^*) A^t =
1545: A\left( \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*)^t -
1546: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) \right) A^t .
1547: \end{equation}
1548: Moreover, because (up to initial value)
1549: $\boldsymbol{\Omega}(A\bY,{\bY}^*) = A\boldsymbol{\Omega}(\bY,{\bY}^*)$,
1550: and
1551: $\boldsymbol{\Omega}(A\bY,H)= A\boldsymbol{\Omega}(\bY,H)$ the final result
1552: \eqref{eq:fund-vect} is independent of (non-degenerate) $A$.
1553: \end{Rem}
1554:
1555: \begin{Prop}
1556: The fundamental vectorial transform given by
1557: \eqref{eq:fund-vect} with the data restricted by conditions
1558: \eqref{eq:C-ft-YY} and \eqref{eq:C-ft-O}
1559: can be considered as the superposition of $K$ (scalar)
1560: discrete CQL reduced fundamental transforms.
1561: \end{Prop}
1562: \begin{proof}
1563: For $K=1$ we obtain the
1564: CQL reduction of the fundamental transformation in the
1565: setting of Proposition~\ref{prop:CQL-fund} (with $c=1$). For $K>1$
1566: the statement follows from the standard reasoning applied to superposition of
1567: two reduced vectorial fundamental transformations (compare with
1568: \cite{TQL,q-red,BQL}).
1569:
1570: Assume the splitting
1571: $\FF^K=\FF^{K_a}\oplus\FF^{K_b}$
1572: and the induced splitting
1573: \begin{equation}
1574: \boldsymbol{Y}_i^* = \left( \begin{array}{cc}
1575: \boldsymbol{Y}_{ai}^{*}\; & \boldsymbol{Y}_{b i}^{*} \end{array} \right),
1576: \end{equation}
1577: of the basic data $\boldsymbol{Y}_i^*$ of the
1578: transformation. Then we have also
1579: \begin{equation}
1580: \boldsymbol{Y}_i = \left( \begin{array}{c}
1581: \boldsymbol{Y}_i^a \\ \boldsymbol{Y}_i^b \end{array} \right) =
1582: \left( \begin{array}{c}
1583: \rho_i(T_i\boldsymbol{Y}^*_{ai})^t \\ \rho_i(T_i\boldsymbol{Y}^*_{bi})^t
1584: \end{array} \right),
1585: \end{equation}
1586: and (in the shorthand notation, compare equations
1587: \eqref{eq:split-fund-1}-\eqref{eq:split-fund-2})
1588: \begin{equation}
1589: \boldsymbol{\Omega}(\boldsymbol{Y},\boldsymbol{Y}^*) =
1590: \left( \begin{array}{cc}
1591: \boldsymbol{\Omega}^a_a & \boldsymbol{\Omega}^a_b \\
1592: \boldsymbol{\Omega}^b_a & \boldsymbol{\Omega}^b_b
1593: \end{array} \right),
1594: \end{equation}
1595: while the constraint \eqref{eq:C-ft-O} reads
1596: \begin{equation} \label{eq:C-ft-Os-split}
1597: (\boldsymbol{\Omega}^a_a)^t = \boldsymbol{\Omega}^a_a, \qquad
1598: (\boldsymbol{\Omega}^a_b)^t = \boldsymbol{\Omega}^b_a, \qquad
1599: (\boldsymbol{\Omega}^b_b)^t = \boldsymbol{\Omega}^b_b .
1600: \end{equation}
1601: By straightforward algebra, using equations
1602: \eqref{eq:C-ft-Os-split}, one checks that the transformed data
1603: satisfy the CQL constraints \eqref{eq:C-ft-YY} and \eqref{eq:C-ft-O}
1604: as well, i.e.,
1605: \begin{equation}
1606: \boldsymbol{Y}_i^{b \{ a\}} = \rho_i^{\{ a\}}
1607: (T_i\boldsymbol{Y}^{*\{ a\}}_{b i})^t ,\qquad
1608: (\boldsymbol{\Omega}^{b \{ a\}}_b)^t = \boldsymbol{\Omega}^{b \{ a\}}_b,
1609: \end{equation}
1610: which concludes the proof.
1611: \end{proof}
1612: \begin{Rem}
1613: In the case with matrix $A$ as in previous Remark, the scalar components of the
1614: vectorial transformation do not satisfy (unless $A$ is diagonal) the CQL
1615: reduction condition of Proposition~\ref{prop:CQL-fund}.
1616: \end{Rem}
1617: \begin{Rem}
1618: Because the CQL-reduced fundamental transformation can be considered as
1619: construction of new levels of the C-quadrilateral lattice, then
1620: if we denote by $x^{\{1,2\}}$ the
1621: C-quadrilateral lattice obtained by superposition of two (scalar) such
1622: transforms from $x$ to $x^{\{1\}}$ and $x^{\{2\}}$, then
1623: for each direction $i$ of the lattice the hexahedra with basic vertices
1624: $x$, $T_i x$, $x^{\{1\}}$ and $x^{\{2\}}$ are C-hexahedra.
1625: Similarly, if we consider superpositions of three (scalar) transforms of the
1626: C-quadrilateral lattice $x$ then
1627: the hexahedra with basic vertices
1628: $x$, $x^{\{1\}}$, $x^{\{2\}}$ and $x^{\{3\}}$ are
1629: C-hexahedra.
1630: \end{Rem}
1631: \section{Conclusion and remarks}
1632: We presented new geometric interpretation of the discrete CKP equation within
1633: the theory of quadrilateral lattices. The paper should be considered as
1634: supplementary to \cite{DS-sym}. It has been also written in a format similar
1635: to \cite{BQL}, where we presented novel geometric interpretation of the discrete
1636: BKP equation.
1637: Results of the paper show once again the
1638: fundamental role of the incidence geometry structures
1639: in the integrable geometry. We remark that the integrability
1640: of the discrete BKP equations was a consequence of the M\"{o}bius theorem on
1641: mutually inscribed tetrahedrons, and the integrability
1642: of the discrete CKP equations was a consequence of the Gallucci theorem.
1643: However, it turns out \cite{Coxeter-PG} that
1644: both theorems are two diferent faces of a more fundamental
1645: result concerning the so called quadrangular
1646: sets of points.
1647:
1648: In the Appendix we present the theory of the Darboux maps
1649: within the quadrilateral lattice theory thus showing the fundamental role of the
1650: quadrilateral lattice in integrable discrete geometry.
1651:
1652:
1653: \appendix
1654: \section{The Darboux maps within the quadrilateral lattice theory}
1655: We would like to present an interpretation of the so called Darboux maps
1656: \cite{Schief-JNMP,KingSchief} within the quadrilateral lattice theory.
1657: Denote by $\EE =
1658: \EE(\ZZ^N)$ the set of edges of the $\ZZ^N$ lattice. Consider a map
1659: \begin{equation}
1660: \boldsymbol{v}:\EE\to\RR^M,
1661: \end{equation}
1662: regarded as a set of $N$ maps $\boldsymbol{v}^i:\ZZ^N\to\RR^M$ of edges in
1663: $i$th
1664: direction. It is termed a discrete Darboux map if the four images of the edges
1665: of any face of the $\ZZ^N$ lattice are collinear, i.e., there exist functions
1666: $\rho_{ij}$, $i\ne j$, such that
1667: \begin{equation} \label{eq:Darboux-map}
1668: \D_j\boldsymbol{v}^i = (T_i\rho^{ij})
1669: (T_i\boldsymbol{v}^j - T_j\boldsymbol{v}^i),
1670: \qquad i\ne j.
1671: \end{equation}
1672: Compatibility of equations \eqref{eq:Darboux-map} implies \cite{Schief-JNMP}
1673: that the functions $\rho^{ij}$ satisfy the discrete Darboux equations
1674: \eqref{eq:MQL-A}.
1675: \begin{Rem}
1676: In order to use the Darboux equations in the form \eqref{eq:MQL-A} the
1677: definition of $\rho^{ij}$ in this paper is shifted with respect to that used in
1678: \cite{Schief-JNMP,KingSchief}.
1679: \end{Rem}
1680:
1681: We will briefly demonstrate that the Darboux maps can be interpreted as suitably rescaled
1682: normalized backward tangent vectors $\tbX_i$;
1683: compare Figure
1684: \ref{fig:back} with Figure \ref{fig:Darboux-map},
1685: where also the geometric construction of the Darboux map is given.
1686: \begin{figure}
1687: \begin{center}
1688: \includegraphics{Darboux-map.eps}
1689: \end{center}
1690: \caption{The Darboux maps and the quadrilateral lattices}
1691: \label{fig:Darboux-map}
1692: \end{figure}
1693:
1694: \begin{Prop}
1695: Consider the quadrilateral lattice $\bx:\ZZ^N\rightarrow\RR^M$ together with
1696: its backward tangent vectors $\tbX_i$ and the corresponding backward rotation
1697: coefficients $\tQ_{ij}$.
1698: Let $\tv_i$ be a scalar solution of the backward linear problem
1699: \eqref{eq:lin-bX}
1700: \begin{equation} \label{eq:lin-tv}
1701: \tD_i\tv_j = (T_i^{-1} \tQ_{ij})\tv_i \; , \qquad \text{or}
1702: \quad \D_i\tv_j = (T_i\tv_i)\tQ_{ij}, \quad i\ne j \; ,
1703: \end{equation}
1704: define the maps $\by_i:\ZZ^N\to\RR^M$, $i=1,\dots , N$,
1705: \begin{equation} \label{eq:y-tX}
1706: \by_i = T_i\left( \frac{1}{\tv_i}\tbX_i \right).
1707: \end{equation}
1708: Then the maps $\by_i$ satisfy the Darboux map equations
1709: \begin{equation} \label{eq:Darboux-map-vby}
1710: \D_i \by_j = (T_j B_{ji})(T_j \by_i - T_i \by_j),
1711: \end{equation}
1712: with the coefficients
1713: \begin{equation} \label{eq:Darboux-map-Lame}
1714: B_{ij} = \frac{\D_j \tv_i}{\tv_i}, \qquad i\ne j, \quad i,j=1,\dots , N.
1715: \end{equation}
1716: \end{Prop}
1717: \begin{proof}
1718: By direct verification using the fact that
1719: both $\tbX_i$ and $\tv_i$ satisfy the same linear system \eqref{eq:lin-bX}.
1720: Geometrically, by results of Appendix A1 of \cite{DNS-Bianchi}, it means
1721: that $\by_i$ and
1722: $\by_j$ represent mutual Laplace transforms \cite{DCN} in the affine gauge, i.e.
1723: $\by_i$, $T_j\by_i$, $\by_j$ and $T_i\by_j$ are collinear and satisfy equation
1724: of the form of \eqref{eq:Darboux-map-vby}.
1725: \end{proof}
1726: \begin{Cor}
1727: The above result can be reversed, i.e., any Darboux map gives rise via equations
1728: \eqref{eq:y-tX} and \eqref{eq:Darboux-map-Lame} to a system of normalized
1729: backward tangent vectors of a quadrilateral lattice. Thus the correspondence
1730: between Darboux maps and quadrilateral lattices occurs on the geometric linear
1731: level.
1732: \end{Cor}
1733: \begin{Rem}
1734: Notice that the functions
1735: $\tv_i$ satisfy the forward adjoint linear problem \eqref{eq:lin-H} with the
1736: rotation coefficients $\tQ_{ij}$ which satisfy the MQL equations
1737: \eqref{eq:MQL-Q}. Then without any calculation we infer
1738: that the coefficients $B_{ij}$
1739: are solutions of the discrete Darboux (MQL) equations \eqref{eq:MQL-A}.
1740: \end{Rem}
1741: Finally, we mention that to the vectors $\by_i$ it can be given
1742: geometric meaning as non-homogeneous
1743: coordinates in $H_\infty$ of the intersections points
1744: $\langle x, T_i x \rangle \cap H_\infty$ of
1745: the tangent lines to the quadrilateral lattice $x:\ZZ^N\to\PP^M$ with the
1746: hyperplane at infinity $H_\infty$ (see Section \ref{sec:M-cons-CQL}).
1747: Using the Pascal hexagon theorem it can be shown that within this interpretation
1748: the "conic condition" of \cite{KingSchief} is equivalent to Definition
1749: \ref{def:C-hexahedron} of the C-hexahedron.
1750:
1751: \section*{Acknowledgements}
1752: The paper was supported by the Polish Ministry of
1753: Science and Higher Education research grant 1~P03B~017~28.
1754:
1755:
1756:
1757: \bibliographystyle{amsplain}
1758:
1759: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1760: \begin{thebibliography}{10}
1761:
1762: \bibitem{AKV}
1763: A. A. Akhmetshin, I. M. Krichever and Y. S. Volvovski, \emph{Discrete analogues
1764: of the Darboux--Egoroff metrics}, Proc. Steklov Inst. Math. \textbf{225} (1999)
1765: 16--39.
1766:
1767: \bibitem{AvL}
1768: H. Aratyn and J. van de Leur, \emph{The CKP hierarchy and the WDVV
1769: prepotential}, Bilinear integrable systems: from classical to quantum,
1770: continuous to discrete, 1--11, NATO Sci. Ser. II Math. Phys. Chem., 201,
1771: Springer, Dordrecht, 2006.
1772:
1773: \bibitem{BBEIM}
1774: E.~D. Belokolos, A.~I. Bobenko, V.~Z. Enol'skii, A.~R. Its, and V.~B. Matveev,
1775: \emph{Algebro-geometric approach to nonlinear integrable equations},
1776: Springer-Verlag, Berlin, 1994.
1777:
1778: \bibitem{BD}
1779: M.~Bia{\l}ecki, A.~Doliwa, \emph{Algebro-geometric solution of the discrete
1780: KP equation over a finite field out of a hyperelliptic curve},
1781: Comm. Math. Phys. \textbf{253} (2005), 157--170.
1782:
1783: \bibitem{Bianchi}
1784: L.~Bianchi, \emph{Lezioni di geometria differenziale}, Zanichelli, Bologna,
1785: 1924.
1786:
1787: \bibitem{BoKo}
1788: L. V. Bogdanov and B. G. Konopelchenko, \emph{Lattice and $q$-difference
1789: Darboux--Zakharov--Manakov systems via $\bar\partial$ method}, J. Phys. A: Math.
1790: Gen. \textbf{28} L173--L178.
1791:
1792: \bibitem{BoKo-N-KP}
1793: L. V. Bogdanov and B. G. Konopelchenko, \emph{Analytic-bilinear approach to
1794: integrable hiererchies II. Multicomponent KP and 2D Toda hiererchies},
1795: J. Math. Phys. \textbf{39} (1998) 4701--4728.
1796:
1797:
1798: \bibitem{Coxeter-PG}
1799: H. S. M. Coxeter, \emph{Introduction to geometry}, Wiley and Sons, New York,
1800: 1961.
1801:
1802: \bibitem{Darboux-OS}
1803: G. Darboux, \emph{Le\c{c}ons sur les syst\'{e}mes orthogonaux et les
1804: coordonn\'{e}es curvilignes}, Gauthier-Villars, Paris, 1910.
1805:
1806: \bibitem{DKJM}
1807: E.~Date, M.~Kashiwara, M. Jimbo and T.~Miwa, \emph{Transformation groups for
1808: soliton equations}, [in:] Proceedings of RIMS Symposium on Non-Linear
1809: Integrable Systems --- Classical Theory and Quantum Theory (M. Jimbo and T.
1810: Miwa, eds.) World Science Publishing Co., Singapore, 1983, pp. 39--119.
1811:
1812: \bibitem{DJKM-CKP}
1813: E.~Date, M. Jimbo, M.~Kashiwara and T.~Miwa, \emph{KP hierarchies of orthogonal
1814: and symplectic type. Transformation groups for soliton equations VI},
1815: J. Phys. Soc. Japan \textbf{50} (1981) 3813--3818.
1816:
1817:
1818: \bibitem{DJM}
1819: E. Date, M. Jimbo and T. Miwa, \emph{Method for generating discrete soliton
1820: equations. III}, J. Phys. Soc. Japan \textbf{52} (1983) 388-393.
1821:
1822: \bibitem{DCN}
1823: A. Doliwa, {\it Geometric discretisation of the Toda system},
1824: Phys. Lett. A {\bf 234} (1997) 187--192.
1825:
1826: \bibitem{q-red}
1827: A.~Doliwa, \emph{Quadratic reductions of quadrilateral lattices}, J. Geom.
1828: Phys. \textbf{30} (1999) 169--186.
1829:
1830: \bibitem{AD-Chicago}
1831: A. Doliwa, \emph{Integrable multidimensional
1832: discrete geometry:
1833: quadrilateral lattices, their transformations and reductions},
1834: [in:]
1835: Integrable Hierarchies and Modern Physical Theories, (H. Aratyn and A. S.
1836: Sorin, eds.) Kluwer, Dordrecht, 2001 pp. 355--389.
1837:
1838:
1839: \bibitem{BQL}
1840: A. Doliwa, \emph{The B-quadrilateral lattice, its transformations and the
1841: algebro-geometric construction}, J. Geom. Phys. \textbf{57} (2007) 1171--1192.
1842:
1843: \bibitem{MQL}
1844: A.~Doliwa and P.~M. Santini, \emph{Multidimensional quadrilateral lattices
1845: are integrable}, Phys. Lett. A \textbf{233} (1997), 365--372.
1846:
1847: \bibitem{DS-sym}
1848: A.~Doliwa and P.~M. Santini, \emph{The symmetric, {D}-invariant and {E}gorov
1849: reductions of the
1850: quadrilateral lattice}, J. Geom. Phys. \textbf{36} (2000) 60--102.
1851:
1852: \bibitem{DS-EMP}
1853: A.~Doliwa and P.~M.~Santini, \emph{Integrable systems and discrete
1854: geometry}, [in:] Encyclopedia of Mathematical Physics, J. P. Fran\c{c}ois,
1855: G. Naber and T. S. Tsun (eds.), Elsevier, 2006, Vol. III, pp. 78-87.
1856:
1857:
1858: \bibitem{TQL}
1859: A. Doliwa, P. M. Santini and M. Ma{\~n}as,
1860: \emph{Transformations of Quadrilateral Lattices}, J. Math. Phys. \textbf{41}
1861: (2000) 944--990.
1862:
1863: \bibitem{DNS-Bianchi}
1864: A. Doliwa, M. Nieszporski and P. M. Santini,
1865: \emph{Geometric discretization of the Bianchi system}, J. Geom Phys.
1866: \textbf{52}
1867: (2004) 217--240.
1868:
1869:
1870: \bibitem{DBK}
1871: A.~Doliwa, M.~Bia{\l}ecki, P.~Klimczewski, \emph{The Hirota equation over finite
1872: fields: algebro-geometric approach and multisoliton solutions} J. Phys. A
1873: \textbf{36} (2003) 4827--4839.
1874:
1875: %\bibitem{FarkasKra}
1876: %H.~M. Farkas and I.~Kra, \emph{{R}iemann surfaces}, Springer--Verlag, Berlin,
1877: % 1992.
1878:
1879: \bibitem{PG-private}
1880: P. Grinevich, private communication, Rome, 1998.
1881:
1882: \bibitem{GuHuZhou}
1883: C. H. Gu, H. S. Hu and Z. X. Zhou, \emph{Darboux transformations in
1884: integrable systems. Theory and their applications to geometry}, Springer,
1885: Dordrecht, 2005
1886:
1887: \bibitem{KvL}
1888: V. G. Kac and J. van de Leur, \emph{The $n$-component KP hierarchy and
1889: representation theory}, [in:] Important developments in soliton theory, (A. S.
1890: Fokas and V. E. Zakharov, eds.) Springer, Berlin, 1993, pp. 302--343.
1891:
1892:
1893: \bibitem{KingSchief}
1894: A. D. King and W. K. Schief, \emph{Application of an incidence theorem for
1895: conics: Cauchy problem and integrability of the dCKP equation}, J. Phys. A
1896: \textbf{39} (2006) 1899-1913.
1897:
1898: \bibitem{KoSchief2}
1899: B.~G. Konopelchenko and W.~K. Schief, \emph{Three-dimensional integrable
1900: lattices in {Euclidean} spaces: Conjugacy and orthogonality}, Proc. Roy. Soc.
1901: London A \textbf{454} (1998), 3075--3104.
1902:
1903:
1904: \bibitem{Krichever-4p} I. M. Krichever, \emph{Two-dimensional periodic
1905: difference operators and algebraic geometry}, Dokl. Akad.
1906: Nauk SSSR {\bf 285} (1985) 31--36.
1907:
1908: \bibitem{MM}
1909: M. Ma\~{n}as, \emph{Fundamental transformation for quadrilateral lattices:
1910: first potentials and $\tau$-functions, symmetric and pseudo-Egorov reductions},
1911: J.~Phys. A \textbf{34} (2001) 10413--10421.
1912:
1913: \bibitem{MDS}
1914: M. Ma\~{n}as, A. Doliwa and P.M. Santini, \emph{Darboux transformations for
1915: multidimensional quadrilateral lattices. I}, Phys. Lett. A \textbf{232}
1916: (1997) 99--105.
1917:
1918: \bibitem{RogersSchief}
1919: C. Rogers and W. K. Schief, \emph{B\"{a}cklund and Darboux transformations.
1920: Geometry and modern applications in soliton theory}, Cambridge University Press,
1921: Cambridge, 2002.
1922:
1923:
1924:
1925:
1926: \bibitem{Schief-Egorov}
1927: W. K. Schief, \emph{Three-dimensional integrable lattices in Euclidean spaces:
1928: conjugacy and orthogonality}, talk given at the Workshop: Nonlinear Systems,
1929: Solitons and Geometry, Oberwolfach, October 1997.
1930:
1931:
1932:
1933: \bibitem{Schief-JNMP}
1934: W.~K. Schief, \emph{Lattice geometry of the discrete Darboux, KP, BKP and CKP
1935: equations. Menelaus' and Carnot's theorems}, J. Nonl. Math. Phys. \textbf{10}
1936: Supplement 2 (2003) 194--208.
1937:
1938:
1939: \bibitem{Sym}
1940: A. Sym, \emph{Soliton surfaces and their applications}, [in:]
1941: Geometric aspects
1942: of the Einstein equations and integrable systems, Lecture Notes in Physics
1943: \textbf{239}, (R.~Martini, ed.), Springer, 1985, pp. 154--231.
1944:
1945:
1946: \end{thebibliography}
1947:
1948:
1949: \end{document}
1950:
1951: