hep-th0307058/1.tex
1: %Membrane topology and matrix regularization
2: %Short version
3: %For hep-th v2
4: %For hep-th v3
5: 
6: \documentclass[12pt]{article}
7: \setlength{\textwidth}{16.5cm}%A4~21cm
8: \setlength{\oddsidemargin}{-0.3cm}
9: 
10: %\textheight 23.5cm
11: %\topmargin -35pt
12: %\def\baselinestretch{1}
13: 
14: %\documentclass[a4j,twocolumn]{article}
15: %\usepackage[dviout]{graphicx}
16: \usepackage[dvips]{graphicx}
17: %\documentstyle[a4j,12pt]{article}
18: 
19: \def\Vec#1{{\mbox{\boldmath $#1$}}}
20: \def\beqa{\begin{eqnarray}}
21: \def\eeqa{\end{eqnarray}}
22: \def\beq{\begin{equation}}
23: \def\eeq{\end{equation}}
24: \def\fracp#1#2{\frac{\partial#1}{\partial#2}}
25: \def\non{\nonumber}
26: \def\L{\left}
27: \def\R{\right}
28: \def\ss{\Vec{\sigma}}
29: \def\Tr{\mbox{Tr}}
30: 
31: \def\NPB#1#2#3{Nucl. Phys. {\bf B#1}(#2)#3}
32: \def\PLB#1#2#3{Phys. Lett. {\bf B#1}(#2)#3}
33: \def\PRD#1#2#3{Phys. Rev. {\bf D#1}(#2)#3}
34: \def\PRL#1#2#3{Phys. Rev. Lett. {\bf #1}(#2)#3}
35: 
36: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
37: \makeatletter
38: \@addtoreset{equation}{section}
39: 
40: \begin{document}
41: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
42: \begin{titlepage}
43: \begin{flushright}
44: hep-th/0307058v3\\ %%%%%%%%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%
45: UT-KOMABA/03-11\\
46: April 2004 %%%%%%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%
47: \end{flushright}
48: \begin{center}
49: 
50: {\bf\Large Membrane topology and matrix regularization}
51: 
52: \vspace{1.5cm}
53: 
54: {\bf  Hidehiko~Shimada}\footnote
55: {
56: E-mail: shimada@hep1.c.u-tokyo.ac.jp
57: }	
58: 
59: \vspace{1.0cm}
60: 
61: {\it  Institute of Physics, Tokyo University,\\
62: Komaba, Megro-ku, Tokyo 153-8902, Japan}
63: 
64: \vspace{1.9cm}
65: \end{center}
66: \begin{abstract}
67: The problem of membrane topology in the matrix model of M-theory is considered. 
68: The matrix regularization
69: procedure, which makes a correspondence  between finite-sized matrices
70: and functions defined on a two-dimensional base space,
71: is reexamined.
72: %
73: It is found that the information of topology of the base space
74:  manifests itself in the eigenvalue distribution of a single matrix.
75: The precise manner of the manifestation is described.
76: The set of all eigenvalues can be decomposed into 
77: subsets whose members increase smoothly, provided that the 
78: fundamental approximations in matrix regularization hold well.
79: Those subsets are termed as eigenvalue sequences.
80: The eigenvalue sequences exhibit a branching phenomenon
81: which reflects Morse-theoretic information of topology.
82: 
83: Furthermore, exploiting the notion of eigenvalue sequences,
84: a new correspondence rule
85: between matrices and functions is constructed.
86: The new rule identifies the matrix elements directly with
87: Fourier components
88: of the corresponding function, evaluated along certain orbits.
89: The rule has semi-locality in the base space, so that it can be used
90: for all membrane topologies in a unified way.
91: %
92: A few numerical examples are studied, and  
93: consistency with previously known correspondence rules is discussed.
94: \end{abstract}
95: \vfill
96: \end{titlepage}
97: \setcounter{footnote}{0}
98: \renewcommand{\thefootnote}{\arabic{footnote}}
99: 
100: \section{Introduction}
101: Topological properties of a system are often important in 
102: investigating the dynamics of the system.
103: It seems certain that
104: M-theory\cite{11_10} 
105: has membranes as dynamical degrees of freedom.
106: Furthermore, the only existing proposal for formulation of M-theory, 
107: namely the matrix model of M-theory\cite{IMF}, can be considered
108: as an attempt to define quantum membrane
109: theory~\cite{Memb_Mr_mrG, Memb_Mr_mrH, S_Memb_Mr}.
110: More explicitly stated, it is a regularized version of 
111: membrane theory in lightcone gauge, dynamical
112: variables becoming $N \times N$ matrices instead of functions
113: defined on two-dimensional worldspace.
114: 
115: 
116: 
117: But, at present, the topological properties of membranes in M-theory are not 
118: known.
119: The concern of this paper is membrane topology
120: \footnote{
121: We use the word membrane topology 
122: to express the topology of a configuration of membranes in a time-slice.
123: The topology is not of a single membrane but of
124: a totality of membranes.
125: Thus, membrane topology is classified by the numbers of
126: membranes $n_i$ which has genus $i=0,1,\cdots$.
127: }
128: in the matrix model.
129: It is believed that the matrix model can describe
130: membranes of arbitrary topologies.
131: However, there has been a problem:
132: we do not know 
133: whether and how the information of the 
134: topology manifests itself in the matrix model.
135: The cause for this problem lies in the manner
136: in which the correspondence between matrices and functions has been given.
137: There has been no unique rule that can deal with all membrane topologies.
138: Instead, we have many different rules for different topologies
139: \cite{Memb_Mr_mrG, Memb_Mr_mrH, S_Memb_Mr, MrTor, MrGen},
140: interrelationships between those rules being unclear.
141: 
142: In this paper, we address this problem
143: by reexamining the regularization procedure,
144: the so-called matrix regularization. 
145: We shall show that the information indeed manifests itself 
146: in the eigenvalue distribution of a single matrix.
147: The precise manner of the manifestation will be described.
148: Moreover, we have constructed
149:  a new correspondence rule between functions and matrices
150: which can be applied to all membrane topologies in a unified way.
151: 
152: We start by discussing 
153: relevant aspects of matrix regularization, in section \ref{SMatReg}.
154: We take the simple view that, classically,
155: the matrix regularization is an approximation
156: of continuum theory by a discretized theory.
157: \footnote{
158: In quantum theory, at the same time, 
159: the matrix regularization is considered as
160:  a definition of continuum
161: theory by a non-trivial limit of discrete theories.
162: This is the reason why we should first treat the finite-$N$
163: theory carefully.
164: }
165: %
166: This approximation between two theories
167: is based solely on some fundamental large-$N$ approximation
168: formulae (\ref{AppNrm})-(\ref{AppBra}).
169: They play a vital role in this paper.
170: We also recall the
171: well-known mathematical analogy between
172: the matrix regularization and 
173: canonical quantization
174: of systems with one degree of freedom, which will be our main tool in 
175: subsequent discussions.
176: 
177: Then, in section \ref{STop}, 
178: we turn to the investigation of membrane topology 
179: in the matrix model.
180: Our basic observation is that, in order to study membrane topology,
181: it suffices to consider the two-dimensional base space, which we 
182: shall term as the $\ss$-space,
183: not the shape of the membranes in the target space. 
184: This observation greatly simplifies the analysis,
185: since it enables us to deal with only a single matrix, not many matrices.
186: 
187: We base the discussion on the analog of the Bohr-Sommerfeld quantization
188: condition.
189: We shall show that,
190: in the case where the fundamental approximations hold well,
191: the eigenvalue distribution of a matrix has a particular structure.
192: Namely, the set of all eigenvalues can be decomposed into subsets 
193: characterized by the following property:
194: the eigenvalues in one of the subsets, when sorted, increase smoothly.
195: We call these subsets as eigenvalue sequences.
196: The grouping of the eigenvalues into sequences
197: reveals a branching phenomenon of sequences.
198: We find that the branching phenomenon, in turn, reflects 
199: certain Morse-theoretic information of topology of the $\ss$-space. 
200: This is our answer to the above problem. Thus, the information of topology
201: manifests itself, in the  world of matrices, as a branching phenomenon of 
202: eigenvalue sequences.
203: 
204: 
205: Furthermore, the notion of eigenvalue sequences enables us to construct
206: a new correspondence rule between matrices 
207: and functions, which is the subject in section \ref{SMRCorr}. 
208: The matrix elements are approximately equal to 
209: Fourier components of the corresponding function,
210: calculated along appropriate orbits on the $\ss$-space.
211: The rule is analogous to the correspondence 
212: noticed by Heisenberg when he created Matrix Mechanics
213: pursuing Bohr's correspondence principle\cite{Heisenberg}.
214: There, the matrix elements of an observable, 
215: in the basis which makes
216: the Hamiltonian diagonal, are equal to the classical
217: Fourier components of the observable along the appropriate
218: classical orbits on the phase space. 
219: We shall show that the fundamental approximation formulae
220: hold well if the new correspondence rule
221: holds.
222: The correspondence rule contains the above-mentioned analog of the
223: Bohr-Sommerfeld condition.
224: This justifies the use of it 
225: in section \ref{STop}.
226: 
227: The new rule is  semi-local in the $\ss$-space, and consequently
228: can be applied for all membrane topologies uniformly, in
229: marked contrast with the previously known rules.
230: This, in particular, enables one to construct functions corresponding
231: to given matrices when the approximations are good. Using previous rules,
232: one could only do the reverse, namely, to construct matrices corresponding 
233: to given functions.
234: This is because one could not know the topology corresponding to
235: the given matrices, and therefore could not choose the rule to be used.
236: 
237: Apart from the unified treatment for all topologies, the new rule 
238: has the virtue that the identification of the matrix elements 
239: with Fourier components is direct, and
240: so that the geometrical meanings of the
241: matrix elements
242: are clear.
243: Our arguments are 
244: also relevant to the matrix model of type IIB string theory
245: \cite{TypeIIBMatrixModel},
246: since the same regularization is involved.
247: Further, the same kind of mathematics as that of matrix regularization appears
248: in such subjects as bound states of D-branes or 
249: non-commutative field theory. Ideas in this paper may find some applications
250: in those subjects.
251: 
252: A few illustrative numerical examples are given
253: in section \ref{SNumEx}.
254: The consistency between our new rule and previous rules
255: is checked by studying them. Finally, we conclude with some discussions
256: in section \ref{SConDis}.
257: 
258: 
259: \section{Matrix regularization} \label{SMatReg}
260: Let us briefly recall the matrix regularization procedure
261: from our viewpoint.
262: Although it is supermembrane theory in eleven dimension
263: \cite{S_Memb_Action}
264: that is relevant to M-theory,
265: we consider  bosonic membrane theory 
266: for simplicity of presentation.  
267: 
268: Firstly, we shall describe the continuum theory.
269: We parametrize the 
270: membranes by three parameters $(\tau,\sigma^1,\sigma^2) =
271: (\tau,\ss)$.
272: Then, the geometrical shape of membranes in spacetime is described by the coordinate
273: functions
274: $x^\mu(\tau,\ss)$. 
275: In lightcone gauge formalism,
276: $\tau$ is chosen to be equal to $x^+$, and $\ss$ is chosen so as to
277: make the area of a domain in the $\ss$-space proportional to  total 
278: $p^+$ contained in the domain. Here, we denote 
279: the momentum density vector of the membranes by $p^\mu$.
280: %
281: The canonical variables of the system are
282: transverse coordinates and momenta (which are functions 
283: defined on the $\ss$-space)
284: as well as zero modes, 
285: \beq
286: x^\alpha(\ss), p^\alpha(\ss); X^-, -P^+.\label{ContiCanVar}
287: \eeq
288: The Hamiltonian is given by 
289: \beq
290: H=[\ss]\int\frac{(p^\alpha)^2+
291: \frac{1}{2}(\{x^\alpha,x^\beta\})^2}{2 P^+} d^2\ss,
292: \label{ContiH}
293: \eeq
294: with Lie brackets 
295: \[
296: \{ f ,g \}= \fracp{f}{\sigma^1}\fracp{g}{\sigma^2}-
297: \fracp{f}{\sigma^2}\fracp{g}{\sigma^1}, \non
298: \]
299: where $f$ and $g$ are functions on the $\ss$-space. 
300: We have also introduced a conventional constant
301: $[\ss]$ which is the total area of the $\ss$-space,
302: $[\ss] =\int d^2\ss$.
303: The remaining ingredients of the theory are the phase space constraints
304: \beq
305: \{x^\alpha, p^\alpha\}(\ss)=0,\label{ContiCon}
306: \eeq
307: and its global version. They correspond to the local symmetry of the 
308: lightcone gauge theory
309: under reparametrization by area-preserving diffeomorphism (APD) on 
310: the $\ss$-space.
311: This is a local symmetry, because one can perform reparametrization 
312: by different 
313: APD for different $\tau$.
314: 
315: Secondly, we shall give the regularized theory.
316: The canonical variables are
317: $N\times N$ matrices as well as  zero modes,
318: \beq
319: \hat{x}^\alpha,\hat{p}^\alpha; X^-, -P^+.\label{RegCanVar}
320: \eeq
321: The Hamiltonian is given by
322: \beq
323: H= N\Tr \frac{(\hat{p}^\alpha)^2-\frac{1}{2}(2\pi
324: [\hat{x}^\alpha,\hat{x}^\beta])^2}
325: {2 P^+},\label{RegH}
326: \eeq
327: and the constraints are,
328: \beq
329: [\hat{x}^\alpha,\hat{p}^\alpha]=0,\label{RegCon}
330: \eeq
331: where $[~,~]$ is a commutator of matrices.
332: 
333: Now, we turn to the explanation of 
334: the matrix regularization.
335: The following fact is essential:
336: there exists a correspondence between appropriate
337: functions on the $\ss$-space
338: $f(\ss),g(\ss),\cdots$ 
339: and matrices $\hat{f}, \hat{g}, \cdots$ such that
340: the fundamental approximation 
341: formulae
342: \beqa
343: \frac{1}{[\ss]}\int f(\ss) d^2 \ss &\approx& \frac{1}{N} \Tr \hat{f} 
344: \label{AppNrm}\\
345: \widehat{f g} &\approx& \hat{f} \hat{g}
346: \label{AppMtp}\\
347: \widehat{\{f,g\}} &\approx& -i \frac{2\pi N}{[\ss]} [ \hat{f} ,\hat{g} ]
348: \label{AppBra}
349: \eeqa
350: hold.
351: Here, we denote by $\widehat{\{f,g\}}$ and $\widehat{f g}$
352: the matrices which correspond to the functions $\{f,g\}(\ss)$
353: and $f(\ss) g(\ss)$, respectively.
354: \footnote{
355: Maybe we should add the linearity of the correspondence,
356: $\widehat{f+g}=\hat{f}+\hat{g}$,
357: for the sake of completeness.
358: We have omitted it since it holds trivially in all our discussions.
359: }
360: The larger is $N$, the better is the approximation.
361: %
362: From these formulae it follows that 
363: the continuum theory, defined by (\ref{ContiCanVar})-(\ref{ContiCon})
364: can 
365: be approximated by a regularized theory defined by
366: (\ref{RegCanVar})-(\ref{RegCon}).
367: We stress the importance of above formulae.
368: They are almost the definition of the matrix regularization.
369: 
370: Since Lie brackets and matrix commutators 
371: both obey the Jacobi identity and antisymmetry, the important 
372: advantage of matrix regularization follows.
373: Namely, the regularized theory has local symmetry under the transformation
374: \beq
375: x^{\alpha\prime}(\tau)=U(\tau)x^\alpha(\tau) U(\tau)^{-1}, \ \ 
376: p^{\alpha\prime}(\tau)=U(\tau)p^\alpha(\tau) U(\tau)^{-1}, \label{LocSymMat}
377: \eeq
378: where $U(\tau)$ is an arbitrary matrix which is a function of $\tau$,
379: corresponding to
380: the APD symmetry in continuum theory.
381: 
382: The matrix regularization procedure is analogous to the
383: quantization of a system which has one degree of freedom,
384: as is well known. The analogy can be summarized as,
385: \beq
386: \begin{array}{c|c}
387: \mbox{Canonical quantization} 
388: & \mbox{Matrix regularization} \\
389: \hline
390:  (x,p)& (\ss^1,\ss^2)\\
391:  \mbox{Canonical transformation}& \mbox{Area-preserving diffeomorphism}\\
392: \{~,~\}_{P.B.} \rightarrow -i\frac{1}{\hbar} [~,~]&
393: \{~,~\} \rightarrow -i\frac{2\pi N }{[\ss] } [~,~]\\
394: \hbar & \frac{[\ss]}{2\pi N} \\
395: \end{array} \label{AnalogyTable}
396: \eeq
397: where $\{~,~\}_{P.B.}$ is the usual Poisson brackets.
398: We shall motivate our discussion by this analogy in section \ref{STop}.
399: 
400: We conclude this section with discussions on 
401: the previously known correspondence rules. 
402: We first recall the general manner the rules are formulated. 
403: We must, 
404: first of all, fix topology of the $\ss$-space.
405: After that, we consider a basis in the vector space of all functions 
406: defined on the $\ss$-space.
407: Then, we define an appropriate basis in the vector space of all 
408: $N \times N$ matrices, and postulate a correspondence
409: between it and
410: the basis in the space of the functions appropriately truncated.
411: The rules are, finally,
412: justified by checking
413: that the fundamental approximations (\ref{AppNrm})-(\ref{AppBra}) 
414: hold well for large $N$ by them.
415: 
416: This manner has made difficult
417: to consider whether and how membrane topology
418: manifests itself in the matrix model.
419: In particular,   
420: one can expand an arbitrary matrix 
421: by basis referring to any particular topology.
422: This fact, at first sight, seems to suggest 
423: that a configuration of matrix model 
424: could  be interpreted as membranes of arbitrary topology,
425: and there would be, therefore, no information of topology
426: in the matrix model.
427: 
428: This is not necessarily true. 
429: Even if one can formally expand some matrices
430: by a basis referring to a particular topology,
431: the fundamental approximations  may not work at all.
432: \footnote
433: {
434: This may be expected from the previously known rules.
435: Let us, for example, imagine a smooth function defined 
436: on a torus. One can construct the corresponding matrix
437: using the basis for torus topology.
438: One can then expand the matrix by the basis (in the space of matrices)
439: corresponding to the topology of a sphere, and construct a function
440: defined on a sphere. We expect that the resulting function
441: would have discontinuity or, in any case, some singularity
442: (see subsection II. C of \cite{ReviewByTaylor}).
443: This implies that 
444: the function varies considerably in a very small length scale.
445: Therefore the approximations may well be no good,
446: since, in general, 
447: the smaller the length scale of the variation of functions,
448: the larger must be $N$ in order that the approximations are good.
449: 
450: However, it is difficult to characterize precisely, 
451: using only previously known correspondence rules, when 
452: the approximations break down. 
453: Hence,
454: it has not been clear if this picture is indeed right.
455: }
456: In our perspective, that the matrix regularization is an approximation
457: scheme, we cannot, then, interpret the matrices as membranes of the
458: particular topology.
459: Information of topology 
460: may be hidden in the matrices in this way.
461: Through sections \ref{STop} and \ref{SMRCorr},
462: we shall see indeed that, 
463: provided that the approximations are good,
464: the information reflects in the
465: eigenvalue distribution.
466: 
467: 
468: 
469: 
470: \section{Membrane topology and matrix regularization}
471: \label{STop}
472: In this section, we show that the information of membrane topology
473: manifests itself in the matrix model.
474: Before explicit description of the manner of the manifestation,
475: let us give some basic observations.
476: 
477: If one wishes to specify the complete shape of membranes in the
478:  target space,
479: one needs information of many (that is, roughly speaking, 
480: as many as the dimension of the target space)
481: functions.
482: This would imply that one should study many matrices in the matrix model.
483: %
484: However, the information of membrane topology,
485: or at least the information of 
486: topology of the $\ss$-space,
487: is contained in one generic 
488: function defined on the $\ss$-space,
489: as is strongly suggested by Morse theory.
490: We choose, as our basic strategy, to consider the latter information.
491: Then, we shall seek  in a single 
492:  matrix the information of topology of the $\ss$-space.
493: 
494: There is another point we would like to discuss.
495: It is most natural to identify functions which are transformed into each
496: other by APD transformations.
497: We shall identify those matrices which are transformed into
498: each other by similarity transformations, 
499: since (\ref{AppBra}) tells us that
500: the counterpart of the APD transformation
501: is the similarity transformation.
502: This has some non-triviality, since 
503: it may happen that the identification is only allowed
504: approximately. Nonetheless, we shall carry out the identification,
505: because that the APD symmetry survives as (\ref{LocSymMat}) is the most
506: important advantage of the matrix regularization.
507: %
508: This identification and our strategy, to consider 
509: the topology of the $\ss$-space, act together to
510: greatly simplify the analysis.
511: Since one can always diagonalize a single matrix,
512: we can concentrate on the eigenvalue 
513: distribution of the matrix.
514: 
515: Having explained our basic strategy,
516: we shall now proceed to investigate the manner of the manifestation
517: of membrane topology in the eigenvalue distribution.
518: 
519: First, let us consider how one can read off the information of topology
520: from a function in an APD invariant way.
521: We choose an arbitrary generic function $f(\ss)$.
522: \footnote
523: {
524: We use the word generic in the sense of Morse theory:
525: we avoid  degenerate functions, constant functions for instance, 
526: which can be changed into  generic functions by arbitrarily small perturbations.
527: }
528: It could be one of the transverse coordinates, for instance.
529: The function is fixed, throughout our discussion, 
530: as a kind of reference.
531: Thus, we shall use $f$, shortly below, as both
532: an analog of the Hamiltonian in canonical 
533: quantization and a Morse function.
534: %
535: As a natural APD-invariant concept with a given function $f(\ss)$,
536: we introduce
537: an ordinary differential equation (ODE)
538: \beq
539: \frac{d}{dt}\ss =\{ \ss,f\} \label{EomOnSs}, 
540: \eeq
541: drawing analogy to the Hamiltonian equation 
542: of motion with a given Hamiltonian function $H(x,p)$, 
543: which is invariant under canonical transformations.
544: This ODE governs the motion of points of the $\ss$-space.
545: \footnote
546: {
547: We note that the independent variable of the ODE, $t$, is
548: just a mathematical tool to substantiate the analogy to canonical
549: quantization. It has nothing to do with physical time coordinates
550: of membrane theory.
551: }
552: Thus, we envisage an auxiliary Hamiltonian-like
553: dynamical system with the $\ss$-space as its phase space
554: and with $f$ as its Hamiltonian.
555: Since $f$ is conserved along the motion by the identity $\{f, f\}=0$,
556: an orbit of this equation is a part of an equal-$f$ line in the $\ss$-space.
557: It will form a closed loop because of the compactness of the $\ss$-space.
558: %%%%%%%%%%%%%%%%%%%%%%Ftorus%%%%%%%%%%%%%%%%%%%
559: \begin{figure}[t]
560: \begin{center}
561: \includegraphics[width=.4\linewidth]{trslpf}
562: \caption {The $\ss$-space which has topology of a torus.
563: The height in the figure is the reference function.
564: Some orbits of the ODE $d \ss/dt=\{ \ss, f\}$ are drawn,
565: which form closed loops that are, in turn, parts of
566: the equal-$f$ lines.
567: If one gradually increases the value of $f$,
568: one observes a branching phenomenon of the orbits:
569: appearing, branching, merging, and disappearing processes at the points
570: A, B, C, and D, respectively.
571: Depicted is essentially the $\ss$-space, 
572: so that the horizontal directions of the figure
573: have rather arbitrary meanings.
574: If one wishes, one can also give definite meanings to the
575: horizontal directions by interpreting 
576: this figure as the geometrical shape
577: of a membrane in the target space, and the reference function
578: as one of the coordinate functions. 
579: } 
580: \label{Ftorus}
581: \end{center}
582: \end{figure}
583: %%%%%%%Ftorus%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
584: 
585: Then, 
586: if we scan the $\ss$-space by gradually increasing the value of $f$,
587: we will observe 
588: branching processes of these orbits.
589: There are four types of these branching processes:
590: appearing, disappearing, branching  and merging.
591: %
592: Let us consider, for a typical example, the situation depicted 
593: in Fig.~\ref{Ftorus}.
594: The membrane topology is that of a torus.
595: The reference function $f$
596: is chosen to be the height in the figure, and 
597: some orbits of (\ref{EomOnSs}) are drawn.
598: In this example, at the points A, B, C, D, 
599: the orbits appear, branch, merge, disappear, respectively.
600: %
601: %
602: We can read off the information of topology from these processes.
603: This is just the well-known idea of 
604: Morse theory.
605: In particular, we obtain the Euler number of the $\ss$-space,
606: by subtracting the total number of the 
607: branching and  merging processes from
608: the total number of the appearing and disappearing processes.
609: 
610: Now, we shall show that this analysis of topology in the world of 
611: functions has a counterpart in the world of matrices. 
612: The analogy of the matrix regularization
613: to  canonical quantization is useful here.
614: In the latter, the Bohr-Sommerfeld quantization condition
615: determines the eigenvalues of the Hamiltonian operator $\hat{H}$ from the classical Hamiltonian function $H(x, p)$ defined on the phase space $(x, p)$.
616: Namely, we draw classical orbits in $(x, p)$ space, 
617: that are parts of equal-$H$ lines,
618: so that the areas of the domains between two adjacent orbits are equal to
619: $2\pi \hbar$. Then eigenvalues of $\hat{H}$ are given by 
620: the values of $H$ at these orbits.
621: %
622: Here, we shall
623: exploit the analogy, which is summarized in (\ref{AnalogyTable}),
624: and state the analog of the Bohr-Sommerfeld condition.
625: Namely, we draw orbits of (\ref{EomOnSs}) in the $\ss$-space
626: so that the areas of the domains
627: \footnote
628: {
629: The area of a domain means here area in the $\ss$-space
630: not in the target space.
631: Its physical meaning is the total $p^+$ contained in the domain,
632: apart from a conventional factor,
633: by the gauge choice made in the lightcone gauge formalism.
634: }
635:  between two adjacent orbits are equal to
636: $[\ss]/N$.
637: Since $[\ss]$ is the total area of the $\ss$-space,
638: this simply means that we divide the $\ss$-space
639: into $N$ parts of equal area.
640: Eigenvalues of $\hat{f}$ are then given by 
641: the values of $f$ at these orbits. 
642: We assume this rule to hold. We shall justify the assumption
643: in section \ref{SMRCorr}.
644: 
645: If we apply this rule to the case in Fig.~\ref{Ftorus}, then 
646: the eigenvalues of $\hat{f}$ can be grouped into 
647: four subsets each of which corresponds to the family of
648: the orbits belonging to 
649: (i) the region from the point A to 
650: the point B, (ii) the left branch of the torus from the point B 
651: to the point C,
652: (iii) the right branch of the torus from the point B to the point C,
653: (iv) the region  from the point C to the point D, respectively.
654: We call these subsets as 
655: eigenvalue sequences.
656: %
657: For large enough N, eigenvalues belonging to each sequence
658:  have the following property.
659: If we sort the eigenvalues contained in a sequence 
660: in increasing order of their values, and make a graph 
661: plotting the values of them versus 
662: their order,
663: then the plotted points can be linearly approximated locally.
664: To put it short, 
665: the eigenvalues in a sequence
666: increase smoothly.
667: It should be clear that, in general, if we do not group the eigenvalues 
668: properly,
669: then the graph become zigzag-shaped and the above property is  lost.
670: In section \ref{SMRCorr}, we see that
671: this linear approximation is  essential in order 
672:  the fundamental approximations
673: (\ref{AppNrm})-(\ref{AppBra}) to hold.
674: 
675: The eigenvalue sequences should exhibit the same branching phenomenon
676: as that of the orbits.
677: For the example of Fig.~\ref{Ftorus}, the sequence
678: (i) appears and then branches into the sequences (ii) and (iii).
679: They merge into the sequence (iv), and finally (iv) disappears.
680: %
681: It is clear that all these considerations work the same in general cases
682: other than that of Fig.~\ref{Ftorus}.
683: Thus, the information of membrane topology manifests itself
684: in the branching phenomenon (which consists of appearing, branching, merging,
685: disappearing processes) of eigenvalue sequences.
686: A few examples, including the case similar to the situation in 
687: Fig.~\ref{Ftorus}, are given in section \ref{SNumEx}.
688: 
689: 
690: \section{The new correspondence rule} \label{SMRCorr}
691: In this section, we present a new 
692: correspondence rule between matrices and functions,
693: and then show that the fundamental approximations
694: (\ref{AppNrm})-(\ref{AppBra})
695: stem from the rule.
696: 
697: 
698: We choose an arbitrary
699:  generic function $f$ and fix it as a reference, as in section 
700: \ref{STop}.
701: The rule is formulated in such a way that
702: the representation of matrices is so chosen that
703:  the matrix $\hat{f}$, corresponding to 
704: the function $f$, is diagonal.
705: For simplicity of notation,
706: we shall consider the case where only one
707: eigenvalue sequence is present. We explain 
708: the generalization later in this section. 
709: 
710: We first give the rule to determine the diagonal
711: matrix $\hat{f}$.
712: To this end, we set up some notations. 
713: We again consider 
714: ODE (\ref{EomOnSs})
715: \[
716: \frac{d\ss}{d t}= \{ \ss, f \} .
717: \]
718: A solution of this ODE is periodic, the point of the $\ss$-space 
719: circulating on a loop which is part of an equal-$f$ line. 
720: We shall denote its period, as a function of $f$, by $T(f)$.
721: %
722: We sort the eigenvalues of $\hat{f}$ 
723: in increasing order, and call them $\hat{f}_n$,
724: \beq
725: \cdots \le \hat{f}_{n-1} \le \hat{f}_n \le \hat{f}_{n+1} \le \cdots. 
726: \eeq
727: To be specific, we choose the representation such that
728: \beq
729: \hat{f}=\mbox{diag}(
730: \cdots, \hat{f}_{n-1}, \hat{f}_n, \hat{f}_{n+1}, \cdots 
731: ).
732: \eeq
733: %
734: The relation between the function $f$ and the 
735: matrix elements $\hat{f}_n$
736: is the analog of the Bohr-Sommerfeld quantization condition stated in 
737: section \ref{STop}. If $N$ is sufficiently large, the rule can be 
738: formulated as,
739: \beq 
740: \hat{f}_m-\hat{f}_n \approx (m-n) \frac{[\ss]}{ N}
741:  \frac{1}{T\bigl(\!
742:  \frac{\hat{f}_m+\hat{f}_n}{2}\!\bigr)}, \label{QtzOfArea}
743: \eeq 
744: when $|m-n|$ is small. 
745: We have used that for two nearby loops, one at 
746: $f$ and the other at $f+\delta f$,
747: the area $\delta S$ between them can be approximated by
748: \beq
749: \delta S=\oint \frac{\delta f}{|
750: \mbox{grad} f|}ds=T\!\biggl(\!f+\frac{\delta f}{2}\biggr)\ \delta f .
751: \eeq
752: We can construct $\hat{f}_m$ satisfying (\ref{QtzOfArea}) directly 
753: by the following method. We first define $S(f) = \int^f (1/T(f)) df$.
754: The value of $S(f)$ runs from $0$ to $[\ss]$ in this case
755: where there is
756: only one eigenvalue sequence.
757: We then consider the inverse function $f(S)$, and set
758: $\hat{f}_m=f(S_m)$, where $S_m$ are determined 
759: by $S_{m+1}-S_m= [\ss]/N$ up to a constant shift.
760: The shift should be of order $1/N$ for consistency.
761: \footnote 
762: {
763: We can determine the shift  by setting $S_1=[\ss]/(2N)$
764: for an eigenvalue sequence beginning
765: with an appearing process. 
766: %This means 
767: %the area of
768: %the part of the $\ss$-space satisfying
769: %$f(\ss)\leq \hat{f}_1$ is equal to $[\ss]/(2N)$. 
770: This is analogous to 
771: the $1/2$ in the Bohr-Sommerfeld condition 
772: $\oint p dq = (n + (1/2) ) 2 \pi \hbar$.
773: %, representing the effect of 
774: %zero point energy.
775: The justification for the above rule comes from
776: the fact that
777: (\ref{DerivedAppNrm}) holds at one more higher order
778: in $1/N$ by this rule.
779: In other words, the rule is just 
780: the midpoint rule for numerical integration.
781: Similar rule exists for an eigenvalue sequence
782: ending with a disappearing process.
783: \label{FootnoteOnShiftZeroPoint}}
784: 
785: 
786: %%%%%%%%%%%%%%%%%%%%% a loop integral along these loops
787: 
788: Having stated the correspondence rule for the reference function,
789: we next turn to the correspondence rule for an arbitrary function $g$.
790: We denote the matrix elements of the corresponding matrix $\hat{g}$
791: by $\hat{g}_{mn}$.
792: %
793: When $|m-n|$ is small, $\hat{g}_{mn}$ is equal to the 
794: Fourier component of
795: order $m-n$ of the function $g(\ss(t))$.
796: Here, $\ss(t)$ denotes the solution of (\ref{EomOnSs}) along which
797: the function $f(\ss)$ takes the (constant) value $(\hat{f}_m+\hat{f}_n)/2$. 
798: To obtain explicit formulae,
799: we define the Fourier components $g_s(f)$ by
800: \beq
801: g(\ss(t))=\sum_{s=-\infty}^{+\infty}
802: g_s(f)\ e^{i \left(\frac{2\pi}{T(f)} s \right) t}, \label{DefOfgOfs}
803: \eeq
804: where the parameter $f$ denotes the value of the function $f(\ss)$
805: along the solution $\ss(t)$.
806: We then set,
807: \footnote{
808: This relation is the direct
809: analog of the correspondence, in semi-classical region,
810: between quantum matrix elements
811: and Fourier components
812: along classical orbits,
813: first introduced in \cite{Heisenberg}.
814: Also, formulae which bear some resemblance
815: to ours appear in \cite{Memb_Matstr}, where a correspondence
816: between membrane theory and Matrix String Theory is considered.
817: %%%%%%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
818: See also \cite{Taylor}.
819: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
820: }
821: \beq
822: \hat{g}_{mn}=g_{m-n}\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr).
823: \label{MatrixElement}
824: \eeq
825: We also require that when $|m-n|$ gets larger, the value of 
826: $g_{mn}$ falls off rapidly. This condition naturally conforms 
827: with (\ref{MatrixElement}), provided that the function $g$ is 
828: sufficiently smooth and
829: $N$ is sufficiently large.
830: 
831: A comment to the rule (\ref{MatrixElement}) is in order.
832: We have freedom to change the orbit $\ss(t)$ by translation of $t$.
833: The amount of translation is a function of $f$, which we denote by 
834: $\Delta t(f)$. By this transformation, $g_s(f)$ becomes
835: \beq
836: e^{i s \frac{2\pi}{T}\ \!\!\!\Delta\! t(\!f\!)} \ g_s(f).
837: \eeq
838: Therefore, $\hat{g}_{mn}$ changes into
839: \beq
840: e^{i (m-n) \frac{2\pi}{T}\ \!\!\Delta\! t
841: \bigl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\bigr)} 
842: \ \hat{g}_{mn}. \label{TimeTranslation}
843: \eeq
844: This freedom has a counterpart in the world of matrices.  %v3
845: Namely, we can change $m$-th eigenvector by a phase factor $e^{i\delta_m}$.
846: By this transformation, $\hat{g}_{mn}$ becomes
847: \beq
848: e^{i (\delta_n -\delta_m)} \hat{g}_{mn}. \label{PhaseChose}
849: \eeq
850: Comparing (\ref{TimeTranslation}) and (\ref{PhaseChose}), we find that if
851: \beq
852: \delta_n -\delta_m \approx (m-n)\ \frac{2\pi}{T}\ 
853: \Delta t\ \!\!\bigg(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\bigg)
854: \label{RelationBetweenTimetranslationAndPhase}
855: \eeq
856: holds, then the two transformations are approximately identical.
857: We can construct $\delta_m$ satisfying 
858: (\ref{RelationBetweenTimetranslationAndPhase}) from given
859: $\Delta t(f)$, provided that $\Delta t(f)$ is sufficiently smooth
860: and $N$ is sufficiently large.
861: 
862: Equations (\ref{QtzOfArea}) and (\ref{MatrixElement})
863: constitute, then, 
864: our new correspondence rule. We shall now deduce
865: the fundamental approximations (\ref{AppNrm})-(\ref{AppBra})
866: from the new rule.
867: 
868: By (\ref{QtzOfArea}), we have divided the $\ss$-space into $N$
869: domains around orbits along which $f(\ss)$ takes the values 
870: $\hat{f}_1, \cdots, \hat{f}_N$.
871: We can evaluate the integral of an arbitrary function $\int g(\ss) d^2\ss$
872: approximately, 
873: by summing up the average values of $g(\ss)$ on these loops
874: multiplied by the areas of each domains.
875: Since (\ref{DefOfgOfs}) and (\ref{MatrixElement}) 
876: tell us that the average value of $g$
877: on the orbit along which $f(\ss(t))=\hat{f}_n$ is $\hat{g}_{nn}$, 
878: and since each area is equal to
879: $[\ss]/N$, we obtain
880: \beq
881: \int g(\ss) \ d^2\!\ss \approx \sum_{n} g_{nn} \frac{[\ss]}{N},
882: \label{DerivedAppNrm}
883: \eeq
884: which is nothing but (\ref{AppNrm}).
885: 
886: We next consider multiplication of matrices constructed by 
887: (\ref{MatrixElement})
888: \beq
889: (\hat{g}\hat{h})_{mn}
890: =\sum_l \hat{g}_{ml}\ \hat{h}_{ln}
891: =\sum_l g_{m-l}\ \!\!\biggl(\!\frac{\hat{f}_m+\hat{f}_l}{2}\!\biggr)\ 
892: h_{l-n}\ \!\!\biggl(\!\frac{\hat{f}_l+\hat{f}_n}{2}\!\biggr).
893: \label{(gh)mn_basic}
894: \eeq
895: Since $g_{m-l}$ and $g_{l-n}$ fall off rapidly when $|m-l|$ and $|l-n|$ 
896: are large,
897: respectively,
898: the terms in which  $l$ is not far away from 
899: $m$ or $n$ dominate the summation.
900: Then, by (\ref{QtzOfArea}), neglecting
901: higher order terms in $1/N$, we can replace both
902: $(\hat{f}_m+\hat{f}_l)/2$ and $(\hat{f}_l+\hat{f}_n)/2$ by 
903: $(\hat{f}_m+\hat{f}_n)/2$. We have, therefore,
904: \beq
905: (\hat{g}\hat{h})_{mn}\approx \sum_l g_{m-l}\ \!\!
906: \biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)\ 
907: h_{l-n}\ \!\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)=
908: (gh)_{m-n}\ \!\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)
909: =\widehat{gh}_{mn},
910: \eeq
911: where the second equality is the convolution law of Fourier series.
912: Thus, the matrix corresponding to the multiplication
913: of the two functions approximately coincides with
914: the multiplication of matrices corresponding to the functions.
915: We have derived (\ref{AppMtp}).
916: 
917: We have just seen that, to the leading order, the multiplication of 
918: the matrices
919: is commutative, since $\widehat{gh}=\widehat{hg}$, as a matter of course.
920: Incorporating one more higher order terms in $1/N$,
921: we shall evaluate the non-commutativity of the matrices.
922: Thus, from (\ref{(gh)mn_basic}), we have
923: \beq
924: (\hat{g}\hat{h})_{mn} \approx 
925: \sum_l \L(g_{m-l}\ \!\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)\ 
926: + \frac{\hat{f}_l-\hat{f}_n}{2} g^\prime_{m-l}\R)
927: \L(h_{l-n}\ \!\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)
928: + \frac{\hat{f}_l-\hat{f}_m}{2} h^\prime_{l-n}\R).
929: \eeq
930: Here, we set $g_s^\prime(f)=dg_s/df$.
931: \footnote
932: {
933: We choose the orbits in (\ref{DefOfgOfs}) smoothly,
934: so that
935: $d g_s / d f$ is well-defined.
936: }
937: We have omitted the value of $f$ at which $g^\prime$ or $h^\prime$
938: is evaluated,
939: since that does not affect the results to the order we are working.
940: By (\ref{QtzOfArea}), it follows that
941: \[
942: (\hat{g}\hat{h})_{mn} \approx \sum_{u+v=m-n}
943: \L(g_{u}\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)
944: +  \frac{[\ss]}{N}\frac{1}{T}\biggl(\frac{v}{2}\biggr)g^\prime_{u}\R)
945: \L(h_{v}\!\biggl(\!\frac{\hat{f}_m+\hat{f}_n}{2}\!\biggr)
946: +  \frac{[\ss]}{N}\frac{1}{T}\biggl(-\frac{u}{2}\biggr)h^\prime_{v}\R),
947: \]
948: where we have introduced new dummy indices $u=m-l, v=l-n$.
949: Then, finally, we have
950: \beqa
951: ([\hat{g}, \hat{h}])_{mn}
952: &\approx& \frac{[\ss]}{N}\frac{1}{T}\frac{1}{2}\L(
953: \sum_{u+v=m-n} \bigl(-g_u (u h^\prime_v) +  (v g^\prime_u) h_v\bigr)
954: - \big(g\leftrightarrow h\big) \R)  \non\\
955: &=&
956:  \frac{[\ss]}{N} \frac{1}{T} \sum
957:  \big( -\!(u g_u) h^\prime _v + g^\prime_u(v h_v)\ \big). 
958: \label{ghCommHighFinal}
959: \eeqa
960: 
961: In order to understand the relation of the last expression to the function 
962: $\{ g, h\}(\ss)$, it is instructive to consider
963: the special case $h=f$. Namely, we consider 
964: the case in which one of the functions is the reference function.
965: In that case we have by (\ref{QtzOfArea}),
966: \beq
967: [\hat{g},\hat{f}]_{mn}
968: =\hat{g}_{mn} (\hat{f}_n -\hat{f}_m)
969: \approx \frac{[\ss]}{N} \frac{1}{T} (n-m) g_{m-n}, \label{gfCommFin}
970: \eeq
971: which is the special case of (\ref{ghCommHighFinal}).
972: On the other hands, 
973: the Lie brackets between $g$ and $f$ can be  
974: expressed by a solution of  (\ref{EomOnSs}) 
975: as,
976: \beq
977: \{g, f\}(\ss) = \left( \frac{d}{dt} g\left(\ss\!\left(t
978: \right)\right) \right) 
979: \Bigg|_{\ss(t)=\ss} \label{HeisenbergEOMinSS},
980: \eeq
981: where the total derivative with respect to
982:  $t$ is taken at the point where the Lie bracket is calculated.
983: Then, from the definition of the Fourier component $g_s$,
984: (\ref{DefOfgOfs}),
985: we get
986: \beq
987: \widehat{ \{g, f\} }_{mn} = i (m-n) \frac{2\pi}{T} g_{m-n} .
988: \label{BragfFin}
989: \eeq
990: Comparing with (\ref{gfCommFin}), we obtain
991: \beq
992: \widehat{ \{g, f\} }_{mn} \approx  -i \frac{2 \pi N}{[\ss]} 
993: [\hat{g},\hat{f}]_{mn}, \label{CommIsLieBraSpecial}
994: \eeq
995: the special case of (\ref{AppBra}).
996: 
997: The last expression in (\ref{ghCommHighFinal}) and the
998: above derivation
999: of (\ref{CommIsLieBraSpecial}) suggest the natural generalization.
1000: We reinterpret the independent
1001: parameter of the ODE, $t$,
1002: as a function defined locally on the $\ss$-space. Then, from 
1003: (\ref{HeisenbergEOMinSS}), we find 
1004: \beq
1005: \{ t , f\} = \frac{dt}{dt} =1,
1006: \eeq
1007: which means that we can consider that $(t, f)$ as canonically conjugate
1008: variables in terms of the analogous canonical formalism.
1009: It follows that,
1010: \beq
1011: \{ g, h \} = \L(\fracp{g}{t}\R)_f \L(\fracp{h}{f}\R)_t - 
1012: \L(\fracp{g}{f}\R)_t \L(\fracp{h}{t}\R)_f. \label{ghBraAsLieBraBytf}
1013: \eeq
1014: The definition of the Fourier components $g_s$, (\ref{DefOfgOfs}), 
1015: is now interpreted 
1016: as the representation of $g$ as a function of $(t ,f)$
1017: \beq
1018: g(t,f)=\sum_{s=-\infty}^{+\infty}
1019: g_s(f) e^{i \left(\frac{2\pi}{T(f)} s \right) t}, \label{gAsAFunctionOftf}
1020: \eeq
1021: Substituting (\ref{gAsAFunctionOftf}) and the similar formula for $h$ 
1022: into (\ref{ghBraAsLieBraBytf})
1023: we get,
1024: \footnote
1025: {
1026: Technically, that $t$ is defined only locally poses a  problem.
1027: However, we can cope with it easily by introducing
1028: patches on each of which $t$ is well-defined, and considering 
1029: the relation between the patches.
1030: }
1031: \beq
1032: (\{g, h\})_s= \sum_{u+v=s} (i \frac{2\pi}{T} u g_u) h^\prime_v
1033: -  g^\prime_u( i \frac{2\pi}{T} v h_v).
1034: \eeq
1035: (Terms in which $(\partial/\partial f)_t $ acts on $1/T(f)$ cancel out.)
1036: By comparing this expression with (\ref{ghCommHighFinal}),
1037: we finally prove (\ref{AppBra}),
1038: \beq
1039: \widehat{ \{g, h\} }_{mn} \approx -i \frac{2\pi N }{[\ss]}
1040: [\hat{g} ,\hat{h} ]_{mn}.
1041: \eeq
1042: 
1043: Up to this point, our derivation has been confined to the case
1044: where there is only one eigenvalue sequence. 
1045: The extension to the general 
1046: case where there are several eigenvalue sequences
1047: is easy. Namely,
1048: we apply (\ref{QtzOfArea}) and (\ref{MatrixElement}) within
1049: each sequences
1050: separately. 
1051: They determine the
1052: matrix elements  between eigenvectors belonging to 
1053: the same sequence.
1054: We then set remaining matrix elements, that is, 
1055: matrix elements between eigenvectors which belong to 
1056: different sequences, to zero.
1057: Above derivations of the fundamental approximations work just the same.
1058: 
1059: This argument means that we can concentrate on the
1060: behaviour of functions on one branch of the $\ss$-space, 
1061: ignoring the behaviour on other branches.
1062: Also, since in our arguments
1063: the matrix element $\hat{g}_{mn}$ falls off rapidly
1064: when $|m-n|$ gets larger, we can ignore the behaviour of the functions
1065: at the place differing much in the value of $f$.
1066: These properties render our new rule a semi-local nature.
1067: That is, both the rule and the approximations work 
1068: locally in the direction $f$ changes.
1069: This situation is somewhat reminiscent of the uncertainty principle
1070: in the analogous quantum mechanical case. We have chosen
1071: the representation to make $\hat{f}$ diagonal. This choice 
1072: achieves minimum uncertainty in $f$, 
1073: and, at the same time, makes the conjugate variable $t$
1074: maximally uncertain.
1075: That our rule can be applied to any topology may be 
1076: considered as a direct consequence of this semi-locality.
1077: 
1078: The linear approximation (\ref{QtzOfArea})
1079: has been essential in the machinery of the derivations
1080: of (\ref{AppNrm})-(\ref{AppBra}).
1081: Therefore, it seems that the linear approximation, hence
1082: the existence of the eigenvalue sequences
1083: is necessary in order that the approximations are good.
1084: Also the use of the analog of the
1085: Bohr-Sommerfeld condition in section \ref{STop} 
1086: is justified, since the condition is nothing but (\ref{QtzOfArea}).
1087: 
1088: Unfortunately, it seems that our new rule does not 
1089: apply to the following exceptional quantities: matrix elements 
1090: near branching and merging processes.
1091: Our rule is essentially a WKB approximation.
1092: In the immediate 
1093: vicinity of the branching and merging processes, 
1094: there should be 
1095: tunneling effects which make the WKB approximation unreliable.
1096: Consider an analog problem in quantum mechanics, that is,
1097: the motion of a particle in the double-well potential.
1098: It is possible to deal with each well
1099: separately semi-classically,
1100: for sufficiently small $\hbar$, and for generic energy levels.
1101: Indeed, tunneling amplitudes between the wells in general 
1102: are negligibly small,
1103: behaving like $\exp{(-O(1)/\hbar)}$.
1104: However, for those rare energy levels which have energy close to 
1105: the value of the potential at the local maximum,
1106: the tunneling amplitudes are 
1107: not negligible. The break down of our rule could also be expected 
1108: from a more direct argument.
1109: The solution of (\ref{EomOnSs}), $\ss(t)$, in the vicinity of
1110: the branching and merging processes,
1111: spends most of the time near the branching point, moving very slowly.
1112: Then, even if $g(\ss)$ is a smooth function, $g(\ss(t))$ might develop
1113: singularity. Then, the validity of the condition used in our argument, 
1114: that $\hat{g}_{mn}$ is negligible for large $|m-n|$, might be
1115: questioned.
1116: 
1117: 
1118: \section{Examples}
1119: \label{SNumEx}
1120: In this section we shall present three examples.
1121: In the first example, by an analytical calculation,
1122: we show the equivalence between our new rule
1123: and the previously known rules. Both
1124: diagonal and 
1125: off-diagonal matrix elements are compared.
1126: %
1127: In the remaining two examples,
1128: our purpose is mainly to 
1129: illustrate the notion of eigenvalue sequences.
1130: We calculate numerically eigenvalues
1131: of matrices constructed by the previously known rules.
1132: We represent the resulting eigenvalue distribution
1133: in a method such that the structure discussed in section 3,
1134: namely the eigenvalue sequences and their
1135: branching phenomenon, can be easily seen.
1136: We confirm that the branching phenomenon of the
1137: eigenvalue sequences coincides with
1138: that of the orbits of the ODE (\ref{EomOnSs}).
1139: We further numerically compute the eigenvalues
1140: by our rule (\ref{QtzOfArea}), and compare them
1141:  with those calculated by the previously known rules.
1142: \paragraph{Example 1}  \label{ExSphA}
1143: \begin{figure}
1144: \parbox{.02\linewidth}{\ }
1145: \begin{minipage}{.45\linewidth}
1146: \begin{center}
1147: \includegraphics[width=.8\linewidth]{sphalpf}
1148: \caption{The $\ss$-space of spherical topology.
1149: The height is the reference function. The orbits
1150: of (\protect\ref{EomOnSs}) appear at A and disappear at B.
1151: } \label{FSphereA}
1152: \end{center}
1153: \end{minipage}
1154: \parbox{.05\linewidth}{\ }
1155: \begin{minipage}{.45\linewidth}
1156: \begin{center}
1157: \includegraphics[width=.8\linewidth]{sphaevf}
1158: \caption{Plot of the eigenvalues and the difference
1159: of the eigenvalues of the matrix corresponding
1160: to the height in Fig.~\protect\ref{FSphereA}. The eigenvalue
1161: sequence appears at A and disappears at B.} \label{FSphereAEv}
1162: \end{center}
1163: \end{minipage}
1164: \parbox{.07\linewidth}{\ }
1165: \end{figure} 
1166: We consider the $\ss$-space which has topology of a sphere.
1167: We represent the $\ss$-space as an unit sphere in $\xi, \eta, \zeta$-space,
1168: \beq
1169: \xi^2+\eta^2+\zeta^2=1
1170: \eeq
1171: with the area element given by
1172: \beq
1173: dS = \sin{\theta} d\theta d\phi,
1174: \eeq
1175: where $\theta$ and $\phi$ are polar coordinates defined by 
1176: $\zeta=\cos{\theta}, \xi=\sin{\theta} \cos{\phi},
1177: \eta=\sin{\theta} \sin{\phi}$.
1178: Then the Lie brackets are $\{\xi, \eta\} = \zeta, \cdots$.
1179: %
1180: We choose the simple reference function $f=\zeta$.
1181: Fig.~\ref{FSphereA} represents the $\ss$-space and the 
1182: reference function.
1183: The orbits of (\ref{EomOnSs})
1184: appear at the point A and disappear at the
1185: point B.
1186: 
1187: We first construct the matrix $\hat{\zeta}$ corresponding
1188: to the function $\zeta$, by our new rule.
1189: The area of the domain $\zeta \leq \zeta^\prime$ is given by
1190: \beq
1191: \frac{\zeta^\prime+1}{2} 4 \pi.
1192: \eeq
1193: Then, by the analog of the Bohr-Sommerfeld quantization condition,
1194: or (\ref{QtzOfArea}), we obtain
1195: \footnote{
1196: See also footnote \ref{FootnoteOnShiftZeroPoint}.
1197: }
1198: \beq
1199: (\hat{\zeta}_1, \cdots, \hat{\zeta}_N) =
1200: (-1+\frac{1}{N}, -1 + \frac{3}{N}, \cdots, 1-\frac{1}{N}). 
1201: \label{ZetaMatElementsNewRule}
1202: \eeq
1203: We further construct the matrices $\hat{\xi}$ and $\hat{\eta}$,
1204: corresponding to the functions $\xi$ and $\eta$.
1205: The solutions to the ODE (\ref{EomOnSs}) can be explicitly written as, 
1206: \beqa
1207: (\xi+ i \eta)(t) &=&  \sqrt{1-\zeta^2} e^{it}\\
1208: (\xi- i \eta)(t) &=&  \sqrt{1-\zeta^2} e^{-it}.\non
1209: \eeqa
1210: Then, from (\ref{MatrixElement}), the only non-zero
1211: matrix elements is,
1212: \beq
1213: (\hat{\xi} +i \hat{\eta})_{m+1,m}=
1214: \sqrt{1- \L(\frac{\hat{\zeta}_m+\hat{\zeta}_n}{2}\R)^2}
1215: =\sqrt{1- \frac{4}{N^2-1} \L(m-\frac{N}{2}\R)^2}
1216: = (\hat{\xi} -i \hat{\eta})_{m,m+1}. \label{XiEtaMatElmntsNewRule}
1217: \eeq
1218: 
1219: We shall now compare these results 
1220: with those obtained from the previously known rules.
1221: The rule for the spherical topology reads~\cite{%
1222: Memb_Mr_mrG, Memb_Mr_mrH, S_Memb_Mr},
1223: \beq
1224: \hat{\xi}= \sqrt{\frac{4}{N^2-1}}\hat{l}_x ,\ \ 
1225: \hat{\eta}= \sqrt{\frac{4}{N^2-1}}\hat{l}_y,\ \ 
1226: \hat{\zeta}= \sqrt{\frac{4}{N^2-1}}\hat{l}_z, \label{BasisToBasisSphere}
1227: \eeq
1228: where $\hat{l}_x, \hat{l}_y, \hat{l}_z$ are generators 
1229: of the representation of $SU(2)$ with spin $l=(N-1)/2$.
1230: Since eigenvalues of $\hat{l}_z$ are
1231: $\{-l, -l+1, \cdots, l \}$, we have,
1232: \beq
1233: \L(\hat{\zeta}_1, \cdots, \hat{\zeta}_N\R)=
1234: \L(-\sqrt{\frac{4}{N^2-1}}\frac{N-1}{2}, 
1235: -\sqrt{\frac{4}{N^2-1}}\frac{N+1}{2}, 
1236: \cdots,\sqrt{\frac{4}{N^2-1}}\frac{N-1}{2}\R),
1237: \eeq
1238: which coincides, for large $N$, with the result of our new rule,
1239: (\ref{ZetaMatElementsNewRule}).
1240: %
1241: Further, it is well known that
1242: in the basis where $\hat{l}_z$ is 
1243: diagonalized, $\hat{l}_x$ and $\hat{l}_y$ have matrix elements only 
1244: between the eigenvectors corresponding to adjacent eigenvalues.
1245: The expression for the non-zero matrix elements are,
1246: \beq
1247: <l_z^\prime+1|(\hat{l}_x+ i\hat{l}_y)|l_z^\prime> 
1248: = \sqrt{l(l+1)-l_z^\prime(l_z^\prime+1)}
1249: = <l_z^\prime|(\hat{l}_x- i\hat{l}_y)|l_z^\prime+1>
1250: \eeq
1251: where we have denoted by $|l_z^\prime>$ the eigenvectors of $\hat{l}_z$ 
1252: belonging to the eigenvalue $l_z^\prime$.
1253: Thus, the result of the previously known rule is,
1254: \beq
1255: (\hat{\xi} +i \hat{\eta})_{m+1,m}=
1256: \sqrt{1- \frac{4}{N^2-1} \L(m-\frac{N-1}{2}\R)\L(m-\frac{N+1}{2}\R)}
1257: =(\hat{\xi} -i \hat{\eta})_{m,m+1}.
1258: \label{SphMatElmntsOldRule}
1259: \eeq
1260: These matrix elements are also approximately equal to 
1261: (\ref{XiEtaMatElmntsNewRule}).
1262: %
1263: The agreements of our rule with the
1264: previously known rule for the simple functions
1265: $\xi, \eta, \zeta$
1266: imply agreements for more general functions
1267: which can be constructed by multiplying
1268: $\xi, \eta, \zeta$ finite (much less than $N$) times.
1269: The  reason for this is that  
1270: the approximate equality between
1271: multiplication of functions and that of matrices,
1272: (\ref{AppMtp}), is valid for both rules.
1273: \footnote
1274: {
1275: For torus topology, similar argument as
1276: in this example,
1277: using simple functions such as $\cos{\sigma^1}$ or $\sin{\sigma^2}$
1278: (see example 3 for definitions),
1279: has a tricky aspect
1280: since these simple functions are degenerate functions in the sense of 
1281: Morse theory.
1282: }
1283: 
1284: 
1285: 
1286: The eigenvalues $\hat{\zeta}_i$ and the difference 
1287: $\hat{\zeta}_{i+1}-\hat{\zeta}_i$
1288: of the eigenvalues are given in Fig.~\ref{FSphereAEv}.
1289: We see that the eigenvalues consist of one eigenvalue sequence.
1290: The sequence appears at the point A and disappears at the point B
1291: in Fig.~\ref{FSphereAEv}.
1292: They correspond to the branching points of
1293: orbits A, B in Fig.~\ref{FSphereA}.
1294: 
1295: \paragraph{Example 2} \label{ExSphB}
1296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1297: \begin{figure}
1298: \begin{center}
1299: \includegraphics[width=.4\linewidth]{sphblpf}
1300: \end{center}
1301: \caption{
1302: The $\ss$-space of spherical topology 
1303: with a more
1304: interesting branching phenomenon of orbits than
1305: Fig.~\protect\ref{FSphereA}.
1306: Appearing, branching, (first) disappearing and (final) disappearing
1307: processes occur at the points A, B, C and D, respectively. 
1308: } \label{FSphereB}
1309: \end{figure} 
1310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1311: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1312: \begin{figure}
1313: \begin{center}
1314: \includegraphics[width=.6\linewidth]{sphb4evf}\label{FSphereBEv}
1315: \end{center}
1316: \caption{The plot of the eigenvalues and their difference,
1317: of the matrix corresponding to the reference function
1318: (\protect\ref{RefFuncSphB})
1319: given in
1320: Fig.~\ref{FSphereB}. The eigenvalue distribution
1321: is calculated both by the previously known rule (open squares),
1322: and our new rule, namely, the analog of the Bohr-Sommerfeld condition
1323: (crosses). They agree almost completely.
1324: The branching phenomenon for 
1325: eigenvalue sequences is the same as that of the orbits
1326: in Fig.~\ref{FSphereB}.
1327: Horizontal lines signify critical values of $f$ at which the 
1328: processes in the branching phenomenon take place,
1329: calculated directly from $f$.
1330: }
1331: \end{figure} 
1332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1333: We treat another case of spherical topology,
1334: which exhibits a more
1335: interesting branching phenomenon of eigenvalue sequences
1336: than the previous example.
1337: Perturbing the reference function considered there,
1338: we here consider the reference function of the form
1339: \beq
1340: f(\ss)=a\zeta +b\xi + c \xi^2. \label{RefFuncSphB}
1341: \eeq
1342: The reference function
1343: and the $\ss$-space are schematically depicted in Fig.~\ref{FSphereB}.
1344: Orbits of (\ref{EomOnSs})
1345: appear at the point A and then branch into two
1346: families at the point B. Then, the orbits belonging to the right branch
1347: disappear at the point C, and finally the orbits belonging to the
1348: left branch disappear at the point D.
1349: 
1350: 
1351: The corresponding matrix $\hat{f}$ is given by 
1352: \beq
1353: \hat{f}=a \sqrt{\frac{4}{N^2-1}} \hat{l}_z + b \sqrt{\frac{4}{N^2-1}} \hat{l}_x+
1354: c\left(\sqrt{\frac{4}{N^2-1}} \hat{l}_x\right)^2, 
1355: \eeq
1356: if one uses the previously known correspondence
1357: rule (\ref{BasisToBasisSphere}). 
1358: We have computed numerically its eigenvalues, in the case
1359: $a=1, b=2, c=6$, with $N=40$.
1360: We have also obtained the eigenvalue distribution
1361: from our new rule.
1362: To this end, we have  computed numerically 
1363: the area of the $\ss$-space as a function of the
1364: height $f$ for each branches of the $\ss$-space.
1365: Then, by the analog of the
1366: Bohr-Sommerfeld condition (\ref{QtzOfArea}),
1367: \footnote
1368: {
1369: See also footnote \ref{FootnoteOnShiftZeroPoint}.
1370: }
1371: we have calculated the eigenvalues of $\hat{f}$.
1372: 
1373: We represent the eigenvalues by the following method
1374: to see the information of membrane topology.
1375: Firstly, we sort the eigenvalues in increasing order,
1376: \beq
1377: \lambda_1 \le \lambda_2 \le \cdots \le \lambda_N.
1378: \eeq
1379: They are given in Fig.~\ref{FSphereBEv}.
1380: In order to see the branching phenomenon clearly,
1381: it is useful to plot also the difference of the eigenvalues 
1382: $\lambda_{i+1}-\lambda_i$. 
1383: By the plot
1384: %, or just by 
1385: %scrutinising the plot of $\lambda_i$ very closely,
1386: one  finds that
1387: from the point B to the point C,
1388: the plot of $\lambda_i$ is zig-zag shaped.
1389: Thus, the plot of $\lambda_i$ gives a juxtaposition of
1390: four eigenvalue sequences.
1391: We see the same branching phenomenon
1392: of the sequences 
1393: as that of the orbits in Fig.~\ref{FSphereB}.
1394: %We thus have a confirmation of the 
1395: %discussion in section \ref{STop}.
1396: The agreement between our new rule and the previously known
1397: rule is remarkable.
1398: 
1399: \paragraph{Example 3} \label{ExTorus}
1400: We consider the $\ss$-space which has  topology of a torus. 
1401: The $\ss$-space can be represented by $[0, 2\pi)\times[0, 2\pi)$,
1402: where periodic boundary conditions are understood.
1403: We choose the reference function to be
1404: \beq
1405: f(\ss)=a \cos{\sigma^1} + b \cos{\sigma^2}.
1406: \label{RefFuncTorus}
1407: \eeq
1408: We assume that $a\neq b$, $a\neq 0$, $b\neq 0$, in order to avoid
1409: degenerate reference functions. To be specific we choose $0<a<b$.
1410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1411: \begin{figure}
1412: \parbox{0.075\textwidth}{\ }
1413: \parbox[t]{0.4\textwidth}{
1414: (a)
1415: \begin{center}
1416: \includegraphics[height=4.5cm]{trsctrf}
1417: \end{center}
1418: }
1419: \parbox{0.05\textwidth}{\ }
1420: \parbox[t]{0.4\textwidth}{
1421: (b)
1422: \begin{center}
1423: \includegraphics[height=5cm]{trsblpf}
1424: \end{center}
1425: }
1426: \caption{
1427: (a) Contour plot of the reference function 
1428: $f=a \cos{\sigma^1}+b \cos{\sigma^2}$, with $0<a<b$.
1429: (b) Schematic picture of the $\ss$-space and the reference
1430: function.
1431: The contours, {\it i.e.},
1432: the orbits of (\protect\ref{EomOnSs}) appear, branch, merge and disappear
1433: at the points A, B, C, and D, respectively. 
1434: In (a) the orbits are so written that the areas of the domains
1435: between two adjacent orbits are [\ss]/N.
1436: }
1437: \label{FTorusContour}
1438: \end{figure}
1439: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1440: %
1441: The reference function is represented in Fig.~\ref{FTorusContour}.
1442: It has essentially the same feature
1443: as the reference function in Fig.~\ref{Ftorus}.
1444: At the points A, B, C, D the function $f$ takes the critical values
1445: $-a-b, a-b, -a+b, a+b$, respectively.
1446: 
1447: In the previously known correspondence rule for torus topology,
1448: one postulates\cite{MrTor}, 
1449: \beq
1450: \widehat{e^{i\sigma^1}} = h_1, \ \ \widehat{e^{i\sigma^2}} = h_2, \label{BasisToBasisTorus}
1451: \eeq
1452: where $h_1$ and $h_2$ are the well-known
1453: $N \times N$ matrices which satisfy the relation
1454: $h_1 h_2 =h_2 h_1 \exp{( i 2\pi /N)}$.
1455: Then, it follows that
1456: \beqa
1457: \hat{f}= \frac{a}{2}(h_1+h_1^\dagger)+ \frac{b}{2} (h_2+h_2^\dagger).
1458: \eeqa
1459: We have computed the eigenvalue distribution of this matrix numerically, 
1460: in the case $a=1, b=3,$  with $N=30$.
1461: We can also calculate them by the new rule
1462: as we have done in the previous example. %\ref{ExSphB}.
1463: \footnote{
1464: Due to the poor knowledge in the vicinity of 
1465: the merging and branching processes discussed at
1466: the end of section \ref{SMRCorr},
1467: we have two (or rather one due to the symmetry of the present example)
1468: undetermined parameters of order $1/N$ mentioned in footnote
1469: \ref{FootnoteOnShiftZeroPoint}. 
1470: %Complete fixing of the order $1/N$ parameters requires
1471: %incorporation of the tunneling effect discussed at the end of
1472: %section \ref{SMRCorr}, and beyond the scope of the present paper.
1473: We have fixed the  order $1/N$ parameter by comparison to the result of
1474: the previously known rule.
1475: }
1476: The results by the two methods are given in  Fig.~\ref{FTorusEigenvalues}.
1477: %
1478: They
1479:  agree well, except at the vicinity of the
1480: branching process at the point B. 
1481: The reason for the discrepancy is 
1482:  noted at the end of the previous section:
1483: our new rule should not be trusted in the vicinity of branching processes.
1484: We can trust the previously known rule, on the other hand,
1485: since the fundamental approximations (\ref{AppNrm})-(\ref{AppBra}) are 
1486: guaranteed by the rule (\ref{BasisToBasisTorus}),
1487: irrespectively of the branching phenomenon.
1488: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1489: \begin{figure}
1490: \begin{center}
1491: \includegraphics[width=.6\linewidth]{trs2evf}
1492: \end{center}
1493: \caption{
1494: Same as Fig.~\protect\ref{FSphereBEv}
1495: but for the reference function
1496: (\protect\ref{RefFuncTorus}) given in
1497: Fig.~\protect\ref{FTorusContour}.
1498: The topology of the $\ss$-space is that of a torus.
1499: The eigenvalues calculated by the new rule and the previously known
1500: rule agree well except at the immediate
1501: vicinity of the point B or the point C.
1502: The branching phenomenon of  the 
1503: eigenvalue sequences is the same as that of the orbits in 
1504: Fig.~\protect\ref{FTorusContour}.
1505: The (approximate) degeneracy of eigenvalues 
1506: from the point B to the point C is only accidental, 
1507: being result of the symmetry of the $f$.
1508: }
1509: \label{FTorusEigenvalues}
1510: \end{figure}
1511: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1512: 
1513: In Fig.~\ref{FTorusEigenvalues},
1514: we see a sequence, which appears at the point A and branches
1515: into two sequences at the point B.
1516: Then, the two sequences merge at
1517: the point C,
1518: and finally the last sequence disappears at the point D.
1519: These branching processes of sequences 
1520: directly correspond to the branching processes
1521: of the orbits A, B, C, D in  Fig.~\ref{FTorusContour}.
1522: 
1523: 
1524: 
1525: \section{Conclusions and discussions} \label{SConDis}
1526: In this paper, we have clarified some elementary
1527: but unknown features of the matrix regularization procedure.
1528: We have worked under the simple view that it is an approximation of 
1529: a continuum theory by a discrete theory.
1530: The approximation between two theories is
1531: based solely
1532: on the fundamental approximation formulae (\ref{AppNrm})-(\ref{AppBra}).
1533: %
1534: We have constructed a new geometrical correspondence rule
1535: between matrices and functions. 
1536: We have shown the validity of the rule directly
1537: by deriving the fundamental approximations from it.
1538: %In the process of the derivation,
1539: %we have also seen what conditions are necessary for 
1540: %the approximations to be good.
1541: %
1542: The new rule is semi-local in the $\ss$-space,
1543: and, as a consequence, can be applied 
1544: to all membrane topologies in a unified way, in marked contrast 
1545: with previously known rules.
1546: Using our rule, 
1547: one can construct functions corresponding to given matrices 
1548: such that the fundamental approximations hold well,
1549: provided that these functions exist.
1550: Whether these functions exist for given matrices can be also determined.
1551: As a physical application, for given matrices
1552: $\hat{x}^\alpha, \hat{p}^\alpha$ of the
1553: matrix model, 
1554: one can construct the geometrical shape and the momentum densities
1555: of the membranes.
1556: 
1557: The new rule includes
1558: the linear approximation (\ref{QtzOfArea}), 
1559: which is the analog of the Bohr-Sommerfeld condition.
1560: The linear approximation
1561: has lead us to the particular structure
1562: of the eigenvalue distribution,
1563: namely the branching phenomenon of the eigenvalue sequences.
1564: The eigenvalue sequences, which we have introduced in this paper,
1565: are subsets
1566: of the all eigenvalues whose members
1567: can be linearly approximated locally.
1568: From the analog of the Bohr-Sommerfeld condition,
1569: we have shown that
1570: the branching phenomenon reflects the information of topology. 
1571: 
1572: Thus, we have clarified the manner
1573: the information of topology
1574: manifests itself
1575: in the eigenvalue distribution.
1576: It is natural to further ask  the question:
1577: ``How completely 
1578: can we read off the information of
1579: topology from given matrices?''.
1580: We shall give here some
1581: observations which are essential to this question.
1582: %
1583: In the first place,
1584: our argument implies that
1585: there is no information of topology in
1586: such ill-behaved matrices for which the fundamental approximation
1587: formulae (\ref{AppNrm})-(\ref{AppBra}) 
1588: do not hold well. 
1589: Indeed, 
1590: it is only for the case (\ref{AppNrm})-(\ref{AppBra}) work,
1591: that the linear approximation
1592: (\ref{QtzOfArea}) should hold.
1593: Hence, even the existence of the eigenvalue sequences
1594: is not guaranteed for those ill-behaved matrices.
1595: %
1596: Secondly, 
1597: there occurs overlapping 
1598: of topologies when we consider the interaction of membranes. 
1599: For a typical example, let us consider process
1600: shown in
1601: Fig.~\ref{FTopologyChangeConti}.
1602: At first there are two spheres.
1603: Then, these spheres approach each other
1604: and 
1605: the distance $\Delta$ (in the target space)
1606: between two spheres reduces to zero gradually,
1607: and finally the two spheres merge into a sphere.
1608: \begin{figure}
1609: \begin{minipage}{.45\linewidth}
1610: \includegraphics[width=.8\linewidth]{tpcngcn2}
1611: \caption{Typical case of topology changing of membrane}
1612: \label{FTopologyChangeConti}
1613: \end{minipage}
1614: \parbox{0.1\textwidth}{\ }
1615: \begin{minipage}{.45\linewidth}
1616: \includegraphics[width=.8\linewidth]{tpcngev2}
1617: \caption{Typical case of topology changing from the viewpoint
1618: of eigenvalue distribution}
1619: \label{FTplCngEvals}
1620: \end{minipage}
1621: \end{figure}
1622: This overlapping is also present in the eigenvalue distribution.
1623: The process from the viewpoint of eigenvalue distribution
1624: is as follows.
1625: In Fig.~\ref{FTplCngEvals} the 
1626: eigenvalue distribution of the matrix
1627: corresponding to
1628: the height in Fig.~\ref{FTopologyChangeConti} is shown.
1629: The eigenvalue distribution consists of two
1630: eigenvalue sequences. 
1631: If the distance $\Delta$ between the two eigenvalue sequences
1632: gradually reduces to zero, 
1633: then we cannot distinguish the eigenvalue distribution from 
1634: that of a matrix corresponding to 
1635: one sphere.
1636: It is interesting 
1637: to treat topology changing processes of 
1638: membranes by the matrix model in this way.
1639: 
1640: Our discussion in this paper has been of purely kinematical nature.
1641: To explore the dynamical implication of our rule is
1642: also clearly important.
1643: For example, 
1644: our consideration has made clear the distinction
1645: between the configurations of matrices
1646: which approximate membranes well and
1647: which do not.
1648: It is interesting to consider whether and how the former 
1649: configurations dominate in the path integral of the
1650: matrix model.
1651: 
1652: We would like to comment on the issue of
1653: the membrane instability\cite{Memb_Instability}.
1654: Let us consider a configuration of membranes 
1655: which has a spike-like portion whose area is less than $1/N$.
1656: If we simply apply the analog of the Bohr-Sommerfeld
1657: condition, we should fail to include the information of the spike
1658: into the matrices. Stated more appropriately,
1659: our argument tells us that
1660: the configuration
1661: cannot be well approximated by $N\times N$ matrices.
1662: We want to stress that this spike has an essential difference 
1663: to the spike which is considered in  the  membrane instability.
1664: One uses the word spike for a portion of a surface
1665:  when its linear dimension is large,
1666: and at the same time its area is small. The difference between the spike
1667: in our context and the spike in the instability context lies 
1668: in the meaning of the area.
1669: In the former, the area means area in the $\ss$-space, that is 
1670: essentially $p^+$.
1671: In the latter, the area means area in the target space, or
1672: the energy of the spike.
1673: This difference is meaningful. Indeed,
1674: there are membranes which have portions that have 
1675: small energy but large $p^+$ or vice versa.
1676: In particular, one can construct configurations of the matrix model
1677: which approximate well the membranes with spikes in the sense of the
1678: membrane instability. 
1679: 
1680: As a direction of further investigation of the matrix regularization
1681: procedure itself,
1682: we recall the discussion at the end of section
1683: \ref{SMRCorr}.
1684: Our rule in the present form does not include tunneling effects between
1685: sequences. Although our rule
1686: gives correct overall behaviour of the matrix elements,
1687: we could not trust the rule in the present form
1688: to investigate the matrix elements in the immediate vicinity of 
1689: the merging and branching processes. 
1690: Concrete
1691: examples of the processes are the points B, C in example 3 in section 5.
1692: %the point B in example 2 is also branching
1693: \footnote{
1694: We can trust our rule 
1695: near appearing or disappearing processes,
1696: such as the point C
1697: in example 2 of section \ref{SNumEx}.
1698: The jump
1699: in the plot of the difference of eigenvalues 
1700: is a natural consequence of our rule.
1701: }
1702: It is an important task
1703: to extend our rule to incorporate the tunneling effects\cite{Tunneling}.
1704: One possible strategy would be  to revisit the analog problem,
1705: namely the quantum mechanics of a particle in a double-well potential.
1706: We can construct a formula to relate
1707: the semi-classical wave functions in both wells, which is valid
1708: even for the energy level
1709: near the local maximum, extending the ordinary argument
1710: in the WKB approximation
1711: using Airy functions.
1712: %
1713: Another interesting question
1714: is the uniqueness
1715: of the correspondence which gives the fundamental approximations.
1716: Although we cannot, at present, provide the proof,
1717: we suspect that the correspondence rule
1718: from which (\ref{AppNrm})-(\ref{AppBra}) can be derived
1719: is unique up to similarity transformation.
1720: Indeed, examples studied in section 5 suggest
1721: that the previously known rules and our new rule are
1722: the same up to similarity transformation.
1723: An immediate consequence of the uniqueness is that
1724: a change of the reference function
1725: should amount to a similarity transformation.
1726: \footnote{
1727: This property implies that
1728: the fundamental approximations are no good
1729: for a configuration of matrices which 
1730: consists of matrices corresponding to different topologies.
1731: }
1732: 
1733: 
1734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%v3 %%%%%%%%%%%%%%%%%%%%%%%
1735: It is believed that
1736: the matrix regularization can be
1737: extended for general even dimensional base spaces on which the
1738: Lie brackets can be defined.
1739: The correspondence rule between matrices and functions
1740: can be easily constructed by using tensor product of matrices,
1741: when the topology of the base space are given by direct
1742: product of some two dimensional spaces.
1743: These extensions are important in the matrix model
1744: of M-theory, in order to incorporate longitudinal 5-branes.
1745: To extend our analysis to 
1746: study topological properties of these higher dimensional objects
1747: is also an interesting problem.
1748: We believe that the analysis analogous to the WKB approximation,
1749: used throughout in this paper, will also be useful
1750: for the higher dimensional case.
1751: However, it would be a challenging task,
1752: since the WKB approximation itself is not fully understood
1753: for generic non-integrable Hamiltonian systems on four 
1754: or more dimensional phase spaces,
1755: compared to that for the necessarily integrable systems
1756: on two dimensional phase spaces considered in this paper.
1757: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1758: 
1759: We conclude 
1760: with three possible applications of the new correspondence rule.
1761: 
1762: (1) One can construct various 
1763: interesting configurations of the matrix model by our rule. 
1764: A particular merit of our rule
1765: in this respect is that it can be applied to a membrane
1766: which has genus higher than two, with no more difficulty 
1767: than to a membrane having topology of a sphere or a torus.
1768: 
1769: (2) Our rule may be useful when investigating
1770: the Lorentz symmetry of 
1771: the matrix model.
1772: It has been tried to regularize the Lorentz generators of 
1773: the continuum theory in order to construct those of the matrix model
1774: \cite{MembLorentzSymm}.
1775: However, since some of the Lorentz generators are not built up by
1776: simple multiplications or integrations or Lie brackets,
1777: one has been inclined to use the basis expansion of the
1778: previously known
1779: correspondence rules.
1780: Since there are many different expansions for different topologies,
1781: it has been difficult to define the
1782: Lorentz generators in a unique way.
1783: Since our rule can be applied uniformly to all topologies,
1784: it is a promising tool in 
1785: constructing
1786: definitions of the Lorentz generators of the matrix model in a unique way.
1787: 
1788: (3) The geometrical interpretation
1789: of matrix elements in our rule may make 
1790: the problem of the large $N$ limit of the matrix model accessible.
1791: The problem can be interpreted as
1792: a renormalization of the membrane theory.
1793: We hope that our rule,
1794: by determining the short-distance (or rather, the small-area)
1795: degrees of freedom,
1796: enables us to construct a block-spin transformation 
1797: of the matrix model.
1798: 
1799: \paragraph{}
1800: I would like to thank, first of all,
1801: Prof. T. Yoneya for discussions, encouragements
1802:  and careful reading of the manuscript.
1803: I would like to thank, for discussions and encouragements, 
1804: other members and
1805: former members of Komaba particle theory group,
1806: especially Y. Aisaka, S. Dobashi
1807: and T. Sato.
1808: Also, I would like to thank W. Taylor, S. Iso and T. Shimada
1809: for discussions and encouragements, %%%%%%%%%%v3%%%%%
1810: and J. Hoppe for helpful and encouraging correspondence.
1811: %%%%%%%%%%%%%%%%%%%%%%%%%%%% v3%%%%%%%%%%%%%%%%%%%%%%%%%
1812: 
1813: 
1814: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% v3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1815: \paragraph{Note Added:}
1816: This paper is an extended version of
1817: the author's master thesis \cite{MasterThesis}
1818: submitted to University of Tokyo on April 2002,
1819: where the basic results %in this paper
1820: were preliminarily reported.
1821: %
1822: The author has recently noticed that Hyakutake
1823: has constructed matrices corresponding to
1824: axial symmetric membrane configurations \cite{Hyakutake},
1825: which are special cases of the general prescription given in 
1826: section \ref{SMRCorr} of this paper.
1827: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% v3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1828: 
1829: \begin{thebibliography}{99}
1830: \bibitem{11_10}
1831: E. Witten, Nucl. Phys. {\bf B443}(1995)85.
1832: \bibitem{IMF}
1833: T. Banks, W. Fischler, S. H. Shenker,
1834: L. Susskind Phys. Rev. {\bf D55}(1997)5112.
1835: \bibitem{Memb_Mr_mrG}
1836: J. Goldstone unpublished. %(1982).
1837: \bibitem{Memb_Mr_mrH}
1838: J. Hoppe MIT Ph. D. thesis (1982)\\
1839: (available at http://www.aei-potsdam.mpg.de/\~{}hoppe).
1840: \bibitem{S_Memb_Mr}
1841: B. de Wit, J. Hoppe, H. Nicolai Nucl. Phys. {\bf B305}[FS23](1988)545. 
1842: \bibitem{MrTor}
1843: D. Fairlie, P. Fletcher, C. Zachos, \PLB{218}{1989}{203}.
1844: 
1845: D. Fairlie, C. Zachos, \PLB{224}{1989}{101}.
1846: 
1847: E. Floratos, \PLB{228}{1989}{335}.
1848: \bibitem{MrGen}
1849: M. Bordemann, E. Meinreken, M. Schlichenmaier,
1850: Comm. Math. Phys.{\bf 165}(1994)281.
1851: \bibitem{Heisenberg}
1852: W. Heisenberg, Z. Phys.
1853: {\bf 33}(1925)879.
1854: \bibitem{TypeIIBMatrixModel}
1855: N. Ishibashi, H. Kawai, Y. Kitazawa, A. Tsuchiya,
1856: Nucl. Phys. {\bf B498}(1997)467.
1857: \bibitem{S_Memb_Action}
1858: E. Bergshoeff, E. Sezgin, P. K.Townsend,
1859: \PLB{189}{1987}{75},\\Ann. Phys. (NY){\bf 185}(1988)330.
1860: \bibitem{ReviewByTaylor}
1861: W. Taylor, Rev. Mod. Phys. {\bf 73}(2001)419.
1862: \bibitem{Memb_Matstr}
1863: Y. Sekino, T. Yoneya, \NPB{619}{2001}{22}.
1864: %%%%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%%%%
1865: \bibitem{Taylor}
1866: W. Taylor, \PLB{394}{1997}{283}.
1867: %%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%%%%%%%
1868: \bibitem{Memb_Instability}
1869: B.de Wit, M.L\"{u}scher, H.Nicolai, \NPB{320}{1989}{135}.
1870: \bibitem{Tunneling}
1871: H. Shimada, work in progress.
1872: \bibitem{MembLorentzSymm}
1873: B. de Wit, U. Marquard, H. Nicolai, Comm. Math. Phys. {\bf 128}(1990)39.
1874: 
1875: K. Ezawa, Y. Matsuo, K. Murakami, Phys. Rev. {\bf D57}(1998)5118.
1876: %%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%%%%%%%
1877: \bibitem{MasterThesis}
1878: H. Shimada, Master's Thesis, University of Tokyo (2002).
1879: \bibitem{Hyakutake}
1880: Y. Hyakutake, \NPB{675}{2003}{241}.
1881: %%%%%%%%%%%%%%%%%%%%%%%v3%%%%%%%%%%%%%%%%%%%%%%
1882: \end{thebibliography}
1883: \end{document}
1884: 
1885: