hep-th0304251/dw.tex
1: \documentclass[12pt]{article}
2: 
3: \usepackage{amssymb}  
4: \usepackage{amscd} 
5: %\usepackage{xypic} 
6: %\usepackage{showkeys} 
7: %\usepackage[matrix,arrow,curve]{xy}             
8: \usepackage{wrapfig}
9: \usepackage{graphicx}
10: \usepackage{cite}
11: 
12: \topmargin -10 mm 
13: \oddsidemargin 2 mm 
14: \evensidemargin 0 mm 
15: \textwidth 155 mm 
16: \textheight 225 mm 
17:  
18: \renewcommand {\theequation}{\thesection.\arabic{equation}} 
19: \newcommand {\newsection}{\setcounter{equation}{0}\section} 
20:  
21: \def \pp {{\mathbb P}} 
22: \def \zz {{\mathbb Z}} 
23: \def \cc {{\mathbb C}} 
24: \def \rr {{\mathbb R}} 
25: \def \de {\partial} 
26: \def \Ka {K{\"a}hler} 
27: \def \nn {{\cal N}}
28: \def \PP {{\cal P}}
29: 
30: \begin{document}
31: \rightline{Bicocca-FT-03-10}
32: \rightline{IHES/P/03/27}
33: \rightline{RR 017.04.03}
34: \vskip .1in \hfill hep-th/0304251
35: 
36: \hfill
37: 
38: \vspace{20pt}
39: 
40: 
41: 
42: 
43: \begin{center}
44: {\Large \textbf{On the Geometry of Matrix Models for N=1*}}
45: \end{center}
46: 
47: \vspace{6pt}
48: 
49: \begin{center}
50: \textsl{M. Petrini $^{a}$, A. Tomasiello $^b$, A. Zaffaroni $^{c}$} 
51: \vspace{20pt}
52: 
53: \textit{$^a$ Centre de Physique Th{\'e}orique, Ecole Polytechnique, 48 Route de Saclay; F-91128 Palaiseau Cedex, France.}
54: 
55: \vspace{10pt} 
56: 
57: \textit{$^b$ I.H.E.S.,
58: Le Bois-Marie, Bures-sur-Yvette, 91440, France.}
59: 
60: \vspace{10pt} 
61: 
62: \textit{$^c$ Universit{\`a} di Milano-Bicocca and INFN\\ Piazza della 
63: Scienza 3, I-20126 Milano, Italy.}
64: 
65: \vspace{10pt} 
66: 
67: \end{center}
68: 
69: \vspace{12pt}
70: 
71: \begin{center}
72: \textbf{Abstract}
73: \end{center}
74: 
75: \vspace{4pt} {\small \noindent 
76: We investigate the geometry of the matrix model associated with 
77: an $\nn=1$ super Yang-Mills theory with three adjoint fields,
78: which is a massive deformation of $\nn=4$.
79: We study in particular the Riemann surface underlying solutions
80: with arbitrary number of cuts.
81: We show that
82: an interesting geometrical structure emerges where
83: the Riemann surface is related  on-shell to the Donagi-Witten 
84: spectral curve. We explicitly identify the quantum field theory 
85: resolvents in terms of geometrical data on the surface.
86: }
87: %\vfill\eject
88: %\noindent
89: \vfill
90: \vskip 5.mm
91:  \hrule width 5.cm
92: \vskip 2.mm
93: {\small
94: \noindent e-mail: Michela.Petrini@cpht.polytechnique.fr, 
95: Alessandro.Tomasiello@cpht.polytechnique.fr, 
96: alberto.zaffaroni@mib.infn.it}
97: 
98: 
99: 
100: \newpage
101: 
102: \newsection{Introduction}
103: 
104: Recently it has been proposed to use matrix models to extract information
105: about holomorphic quantities of a certain class of $\nn=1$ gauge
106: theories. More precisely,
107: %Some recent works have proposed a relation between 
108: %$\nn=1$ gauge theories and matrix models. 
109: the exact superpotential
110: and the condensates of chiral operators in the gauge theory can be
111: computed from the free energy of an associated matrix model \cite{dv3}.
112: This proposal has been extensively tested in the case 
113: of a pure $U(N)$ theory with an adjoint field with potential
114: $W(\Phi)$ (and possibly matter in the fundamental
115: representation), showing that the matrix
116: model results and the quantum field theory ones agree. 
117: %The case of a pure $U(N)$ theory with an adjoint field with potential
118: %$W(\Phi)$ (and possibly matter in the fundamental
119: %representation) is relatively well understood. This case 
120: %has been extensively studied, showing that the matrix
121: %model results and the quantum field theory ones agree. 
122: This model has 
123: an interesting
124: geometrical structure in terms of a Riemann surface \cite{int,cv,cdsw,csw3}.
125: %emerges by studying the Ward identity for the model. 
126: %From the analysis of the pure $U(N)$ theory
127: In particular, the resolvents in the matrix
128: model and in the field theory seem to be related to
129: geometrical quantities on this surface. This geometrical
130: structure is deeply related to the Seiberg-Witten curve of the 
131: $\nn=2$ theory of which the $\nn=1$ theory is a deformation.
132: In this paper we will focus on the geometry of the $\nn=1$ theories
133: that are obtained as deformations of the $\nn=4$ SYM. More precisely,
134: we consider the model with three adjoint fields, a mass term for two
135: of them and a generic potential for the third one. The associated
136: matrix model was solved in the case of a single cut in \cite{dv3,dhks},
137: showing remarkable agreement with the quantum field theory results.
138: In this paper, we are interested in the geometry of the 
139: matrix model for arbitrary
140: number of cuts. Since the model can be considered as an $\nn=1$ 
141: deformation of an $\nn=2$ theory with a massive hypermultiplet,
142: the corresponding geometry is expected to be related to
143: the Donagi-Witten curve \cite{dw}.   The emergence of an elliptic curve 
144: in the case of a single cut \cite{dv3} is a first example of this
145: connection. We will investigate the relation between the matrix model and 
146: the Donagi-Witten curve, 
147: and the geometrical structure that emerges in this way.
148: 
149: To be more concrete, 
150: we will study an $\nn=1$ supersymmetric gauge theory with
151: gauge group $U(N)$ and three chiral fields $\Phi$, $X$, $Y$ 
152: in the adjoint representation. The theory is characterized
153: by the superpotential
154: \begin{equation}
155:   \label{eq:supot}
156:   {\cal W}(\Phi, X, Y)= i\Phi [X, Y] + m X\,Y + W(\Phi)\ ,
157: \end{equation}
158: where $W=\sum_{k=1}^{n+1}g_k{\rm Tr}\Phi^k$ 
159: is a polynomial in $\Phi$ of degree $n+1$. 
160: This model can be seen as a deformation of the $\nn=2$
161: theory with a massive adjoint hypermultiplet considered by
162: Donagi and Witten \cite{dw}.
163: We are mainly interested in a
164: classical vacuum  where $X$ and $Y$ vanish and $W'(\Phi)=0$,
165: even though  our results can be applied also to other vacua. 
166: The $N$ classical 
167: eigenvalues of
168: $\Phi$ can be distributed over the $n$ points $\phi_i$ where $W'(x)=
169: \prod_{i=1}^n (x-\phi_i)=0$. 
170: The gauge group is correspondingly broken to $\prod_{i=1}^{n} U(N_i)$, 
171: where $N_i$ is the number of
172: eigenvalues corresponding to the $\phi_i$ critical point of $W$. 
173: At low energy, we can write an effective superpotential 
174: \begin{equation}
175: W_{\rm eff}(S_i,\phi_i,m,\tau)
176: \end{equation}
177: for the condensate superfields 
178: $S_i={\rm Tr} (W_{\alpha}W^{\alpha})_{(i)}$.
179: 
180: In quantum field theory, the vacuum expectation value of
181: $W_{\rm eff}$ can be computed (in principle) from the knowledge of
182: the underlying $\nn=2$ Seiberg-Witten curve. The addition of the
183: superpotential $W(\Phi)$ indeed selects particular points in the
184: moduli space of the $\nn=2$ theory, which can be found by
185: explicitly solving the equations of motion. Once these points
186: are found, $W_{\rm eff}$ can be determined via
187: \begin{equation} 
188: W_{\rm eff}=\langle W(\Phi)\rangle_{\nn=2}
189: =\sum_{k=1}^{n+1}g_k{\rm Tr}\langle\Phi^k\rangle_{\nn=2} \ .
190: \end{equation}
191: In our case the $\nn=2$ theory is described by the Donagi-Witten curve 
192: for $U(N)$, which is an $N$-sheeted covering of a torus.
193: The points selected by the equations of motion are typically
194: points of partial or maximal degeneracy of the curve. In the 
195: generic vacuum $\prod_{i=1}^{n} U(N_i)$ we indeed expect that
196: $N-n$ monopoles are massless. 
197: 
198: In the corresponding matrix model, the vacuum
199: $\prod_{i=1}^{n} U(N_i)$ is associated with an $n$-cut
200: solution of the equations of motion. We will show that, 
201: as in \cite{dv3}, the matrix model can be associated with a 
202: Riemann surface of genus equal to the number of cuts.
203: $W_{\rm eff}$ is then computed from geometrical data on this surface.
204: It is then interesting to investigate the relation of this
205: geometrical structure with the Donagi-Witten curve.
206: The single cut case was solved in \cite{dv3,dhks}. 
207: It corresponds to a field theory massive vacuum where 
208: the $\nn=2$ spectral curve has maximal degeneracy
209: and itself becomes a torus with modular parameter $\tau/N$. 
210: The emergence of such torus on the matrix model side was explicitly 
211: determined in  \cite{dv3,dhks}. 
212: The connection with the $\nn=2$ curve becomes more explicit when
213: one considers the opposite case of maximal number of cuts.
214: As in \cite{cv}, we can introduce a degree $N+1$
215: superpotential with 
216: $W'(x)=
217: \epsilon \prod_{i=1}^N (x-\phi_i)=0$. In the vacuum where
218: all the $N$ eigenvalues of $\Phi$ are distinct the gauge group
219: is broken to the maximal abelian subgroup $U(1)^N$.  
220: In the limit where the superpotential is turned off ($\epsilon\rightarrow 0$), 
221: we should recover the dynamic of
222: the $\nn=2$ theory. We will show that the Riemann surface of
223: the associated matrix model 
224: with $N$ cuts becomes in this limit the Donagi-Witten curve.
225: 
226: We will also analyze the geometry of the matrix model 
227: for an arbitrary number of cuts.
228: We will show that, upon minimization,
229: the Riemann surface always becomes a covering of a torus
230: and we will discuss the relation of this surface
231: with the Donagi-Witten construction. 
232: We will also identify the field theory
233: resolvents with geometrical quantities on the
234: Riemann surface. All the results have 
235: a direct analogue in the case of a pure $U(N)$ theory 
236: \cite{dv3,cv,cdsw,csw3}. 
237: %The only result that we will not be able to write
238: %explicitly is a complete set of loop equations, since these are
239: %considerably more complicated than in the case of pure gauge.
240: %We will nevertheless derive some suggestive relations. 
241: We will mainly consider 
242: the on-shell theory, where a minimization with respect to the moduli
243: has been performed. 
244: The organization of the paper is as follows. In Section 2, we discuss
245: the Riemann surface associated with the off-shell theory. In Section 3,
246: we will discuss the conditions following from the minimization and
247: the relation of the on-shell theory with the Donagi-Witten curve. In
248: Section 4, we discuss the identification of the field theory
249: resolvents with geometrical quantities on the Riemann surface.
250: Finally, in the Appendix we briefly discuss the loop
251: equations and some identities for the off-shell theory.
252: 
253: \newsection{The Riemann surface} 
254: Following \cite{dv3}, we can compute $W_{\rm eff}$ from
255:  the large $\hat N$ expansion of a matrix model:
256: \begin{equation}
257: \int {\cal D} \hat\Phi {\cal D} \hat X {\cal D} \hat Y e^{ \frac{1}{g_s} 
258: {\rm Tr} \{
259: i \hat\Phi [\hat X, \hat Y] + m \hat X \, \hat Y + W(\hat\Phi)\}},
260: \label{path}
261: \end{equation}
262: where  $\hat \Phi, \hat X, \hat Y$ are $\hat{N} \times \hat{N}$ hermitian matrices. The model can be solved integrating out the matrices 
263: $\hat X, \hat Y$ and diagonalizing the  remaining matrix 
264: $\hat \Phi$. 
265: The saddle point 
266: equation of motion reads
267: \begin{equation}
268:   \label{eq:mmeom}
269: W'(\lambda_I) = g_s \sum_{J\neq I}\left[ 
270: 2\frac1{\lambda_I-\lambda_J}
271: -\frac1{\lambda_I-\lambda_J+ im}
272: -\frac1{\lambda_I-\lambda_J- im}\right]\ ,
273: \end{equation}
274: where $\lambda_I$ are the eigenvalues of $\hat\Phi$.
275: As usual, in the large $\hat{N}$ limit, the eigenvalues will be spread 
276: over $n$ cuts around each solution of $W'(\hat\Phi)$. We will
277: denote the fraction of eigenvalues for each cut as $\hat N_i$.
278: According to the Dijkgraaf--Vafa prescription the
279: filling fractions $S_i=g_s\hat N_i$ are identified with the field theory
280: condensates and the effective superpotential 
281: corresponding to  the $\prod_{i=1}^{n} U(N_i)$ vacuum is \cite{dv3} 
282: \begin{equation}
283:   \label{eq:Weff1}
284:   W_{\rm eff}= 
285: \sum_i \left( N_i \frac{\de {\cal F}}{\de S_i} - 2\pi i\tau S_i \right),
286: \end{equation}
287: where ${\cal F}(S_i)$ is the matrix model free energy. 
288: Unless explicitly stated, we will
289: consider the case with the maximal allowed number of cuts.
290: 
291: Information about the model is encoded in the resolvent
292: $\omega(x)\equiv \frac{1}{\hat{N}}{\rm Tr} \frac{1}{x-\hat\Phi}$. 
293: As usual, $\omega(x)$ has cuts corresponding to the distribution of the 
294: matrix model eigenvalues.
295: For this particular model, 
296: it is useful to  define \cite{nek,dv3,dhks} the function 
297: \begin{equation}
298:   \label{eq:G}
299:   G(x) = U(x) + 
300: i S \left[ \omega\Big(x+\frac i2 \,m\Big) -
301: \omega\Big(x-\frac i2 \,m\Big)\right]\ ,
302: \end{equation}
303: where $S\equiv g_s \hat{N}$ is the 't Hooft coupling and 
304: $U(x)$ is defined by
305: the property
306: \begin{equation}
307:   \label{eq:U}
308:   U\Big(x+\frac i2\,m\Big)-U\Big(x-\frac i2\,m\Big) = W'(x)\ .
309: \end{equation}
310: Then as a consequence of (\ref{eq:mmeom})  \cite{nek,dv3,dhks}
311: \begin{equation}
312:   \label{eq:rs}
313: G\Big(x+\frac i2 \,m \pm i\epsilon\Big)=
314: G\Big(x-\frac i2 \,m \mp i\epsilon\Big)\qquad {\rm for}
315: \ \, x \ \, {\rm on\  a\ cut.}
316: \end{equation}
317: %This is not the only consequence of the matrix model equations of
318: %motion (\ref{eq:mmeom}). One can also extract loop equations
319: %that we will consider later in section \ref{sec:res}. For the time
320: %being we stick to (\ref{eq:rs}). 
321: %This equation is important because it
322: %already displays the $\nn=1$ curve.
323: 
324: 
325: %\begin{wrapfigure}[8]{R}{7cm}
326: %\hspace{.5cm}
327: %\begin{picture}(80,50)
328: %\put(180,50){\small\fbox{$x$}}
329: %\includegraphics[width=7cm]{mm1.eps}
330: %\end{picture}
331: %\caption{\small Cuts in the $x$ plane.}
332: %\label{cuts}
333: %\end{wrapfigure}
334: %
335: %\begin{wrapfigure}[8]{R}{7cm}
336: %\hspace{.5cm}
337: %
338: %\begin{figure}[h]
339: %\centerline{{\small\fbox{$x$}}
340: %\includegraphics[width=13cm,height=4.5cm]{mm.eps}}
341: %\caption{\small Cuts in the $x$ plane for the function $G$.
342: %Lower cuts are identified with upper cuts as shown using dashed lines.
343: %$A_i$ and $B_i$ form a basis of cycles for $\Sigma$.}
344: %\label{cuts}
345: %\end{figure}
346: 
347: \begin{figure}[h]
348: \hspace{.5cm}
349: \begin{picture}(80,140)(-10,0)
350: \put(0,0){{\small\fbox{$x$}}}
351: \put(15,0){\includegraphics[width=13cm,height=4.5cm]{mm.eps}}
352: \put(40,80){$B_1$}\put(55,130){$A_1$}
353: \put(165,80){$B_2$}\put(180,130){$A_2$}
354: \put(325,80){$B_3$}\put(310,130){$A_3$}
355: \end{picture}
356: \caption{\small Cuts in the $x$ plane for the function $G$.
357: Lower cuts are identified with upper cuts as shown using dashed lines.
358: $A_i$ and $B_i$ form a basis of cycles for $\Sigma$.}
359: \label{cuts}
360: \end{figure}
361: 
362: 
363: Equation (\ref{eq:rs}) means that $G$ is well-defined
364: on a surface obtained from the $x$ plane after $n$ identifications, as
365: shown in figure \ref{cuts}. If one adds the point $x=\infty$, 
366: this space is topologically 
367: a Riemann surface $\Sigma$ 
368: of genus $n$. Notice that this way of describing a Riemann surface
369: is somewhat reminiscent of the light cone parameterization of moduli
370: space of punctured Riemann surfaces \cite{gw}.
371: 
372: The function $G$ and the Riemann surface $\Sigma$ allow to write
373: the effective superpotential~(\ref{eq:Weff1}) in a more convenient form.
374: First of all, integrating equation~(\ref{eq:G}) around the cuts
375: we obtain the relation
376: \begin{equation}
377: S_i= \frac1{2\pi}\int_{A_i} G\, dx\ ,
378: \label{Si}
379: \end{equation}
380: where $A_i$ are the cycles encircling the cuts in the $x$ plane.
381: Moreover, as shown in \cite{dv3}, $\partial {\cal F}/ \partial S_i$
382: can be written in terms of integrals over the cycles $B_i$ going
383: from a lower cut to an upper one. The superpotential then reads
384: \begin{equation}
385:   \label{eq:Weff}
386:   W_{\rm eff}= 
387: i \sum_i \left( N_i \int_{B_i} G\, dx - 
388:     \tau \int_{A_i} G\, dx \right).
389: \end{equation}
390: 
391: We are interested in the geometrical structure of the problem. 
392: To this purpose, we can re-formulate the matrix model data 
393: in the following way. 
394: The coordinate $x$ of the matrix model plane is not a well-defined 
395: function on the Riemann surface $\Sigma$. Its differential $dx$, though, 
396: is well-defined. It has a double pole in the point $x=\infty$ and 
397: is regular 
398: otherwise; its $A$ periods are zero, and its $B$ periods are $i\,m$.
399: Moreover we can see from eq. (\ref{eq:G})
400: that the function $G$ has a single pole of order $n+1$ 
401: at the point $x=\infty$. By Riemann-Roch this property
402: singles out $G$ up to an additive constant. 
403: Thus our geometrical data are 
404: \begin{itemize}
405: \item a Riemann surface $\Sigma$ of genus $n$ 
406: \item a differential $dx$ with a double pole and periods 
407:   \begin{equation}
408:     \label{eq:dx}
409:     \int_{A_i} dx =0\ ,\qquad\int_{B_i} dx = i\,m\ .
410:   \end{equation}
411: \end{itemize}
412: 
413: Let us count the number of moduli of our data. 
414: We have $3n-3+1$ moduli from the moduli
415: space of Riemann surfaces of genus $n$ with one puncture. Again 
416: from Riemann-Roch one gets $n+1$ for the number of 
417: differentials with a double pole; integrating this to obtain $x$ 
418: involves an extra additive constant. Taking away $2n$ from (\ref{eq:dx}) 
419: one gets finally 
420: \[
421: (3n-2)+ (n+1)+1-(2n)= 2n\ .
422: \]
423: 
424: From the matrix model point of view, 
425: these $2n$ moduli are easily interpreted as the classical vacua
426: (points $\phi_i$ in which $W'(\phi_i)=0$),  
427: and the filling fractions $S_i$. 
428: %Physically, the
429: %latter are gaugino condensates relative to the $N$ factors $SU(N_i)$. 
430: %As for 
431: %the $x_i$, let us consider as in \cite{cv} an $\nn=2$ 
432: %limit in which they are kept 
433: %constant, while sending to zero an overall constant multiplying the 
434: %superpotential ($W'(x)=\epsilon w'(x)$, $\epsilon\to 0$). Then the $x_i$ get
435: %interpreted as moduli of the resulting $\nn=2$ theory, that is the 
436: %Donagi-Witten as we stressed at the beginning. 
437: %Let us count the number of moduli of our data. 
438: %Before we consider that in more detail in next section, 
439: Alternatively, we can also interpret the $2n$ moduli as
440: %some more feature of the data $(\Sigma, dx)$ considered above. 
441: %First of all,
442: the zeros of $dx$. A meromorphic 
443: differential with a double pole must indeed have $2n$ zeros: 
444: in the $x$ plane 
445: they are the end-points of the cuts, denoted by small circles 
446: in figure \ref{cuts}. We see
447: then that $dx$ has precisely the role played by $dw$ in the light-cone 
448: parameterization of the moduli space of Riemann surfaces with punctures in
449: \cite{gw}\footnote{The condition of having specified periods has an analogue 
450: in that case, where the differential was required to have purely
451: imaginary periods.  A differential with a double pole and purely imaginary
452: periods is uniquely determined on any Riemann surface
453: by a procedure similar to that in \cite{gw}; 
454: then imposing the actual value of these
455: periods as in (\ref{eq:dx}) constrains to a sub-variety of the moduli space.
456: The only real difference with \cite{gw} is that $dw$ has two single
457: poles while our $dx$ has one double pole.}. 
458: 
459: It is interesting to compare our data
460: with the case of a pure $U(N)$ gauge theory. In that case,
461: the Riemann surface has genus $n$ and can be explicitly written as
462: \begin{equation}
463: \label{riemann1}
464: y^2=W'(x)^2+f_{n-1}(x) \ ,
465: \end{equation}
466: where $f_{n-1}$ is a polynomial of degree $n-1$, whose coefficients
467: are related to the moduli $S_i$. The role of the meromorphic function $G$ 
468: is played by $y$: all the relevant formulae are obtained with  
469: the substitution
470: $Gdx \rightarrow ydx$.
471: An important difference with respect to our case,
472: is that the matrix model plane for pure $U(N)$ corresponds to just
473: one sheet of the surface~(\ref{riemann1}). On this sheet we have
474: the relation
475: $y=W'(x)-2\omega(x)$, instead of equation~(\ref{eq:G}).
476: All quantities are then continued to the second sheet. In our case,
477: the matrix model plane is topologically identified with the entire
478: Riemann surface.   
479: 
480: 
481: \section{Minimizing the superpotential}
482: 
483: The results obtained so far from the matrix model apply to
484: an off-shell theory. In order to obtain
485: the  vacuum value of the superpotential we further have to 
486: minimized equation~(\ref{eq:Weff}) with respect to $S_i$.
487: We refer to the theory after minimization as the on-shell theory.
488: %We repeat here the idea of this limit \cite{cv}. Choosing a superpotential
489: %$W(\Phi)=\epsilon \Pi_i (x -x_i)$ and sending $\epsilon\to 0$ one can find 
490: %oneself in an arbitrary point of $\nn=2$ moduli space. {\bf Da spiegare 
491: %meglio.} Here we are also {\bf supposing (?)}  that the $x_i$ 
492: %are not modified quantum mechanically, 
493: %that in the pure $\nn=2$ gauge case was 
494: %established \cite{cv} thanks to explicit result that we lack here.
495: %So we have to minimize the superpotential $W_{\rm eff}$ with 
496: %respect to $S_i$. 
497: 
498: %We will see that after minimization the Riemann surface $\Sigma$
499: %is related to the Donagi-Witten curve. 
500: 
501: Define the differentials
502: \begin{equation}
503: \omega_i=\frac{1}{2\pi}\frac{\de}{\de S_i} G\,dx\ .
504: \label{differentials1}
505: \end{equation}
506: The $\omega_i$ have no poles and 
507: are therefore holomorphic differentials.
508: %Indeed, when we differentiate equation~(\ref{eq:G}), $\omega_i$ 
509: %get contributions
510: %only from the resolvent $\omega(x)$, which behaves at infinity
511: %as $\omega(x)\sim 1/x$; the potential simple pole is canceled 
512: %in the difference of the two resolvents
513: Indeed, when we differentiate equation~(\ref{eq:G}), the $\omega_i$ 
514: get contributions
515: only from the difference of the resolvents; 
516: the simple pole in the resolvent, which behaves at infinity
517: as $\omega(x)\sim 1/x$, cancels 
518: in the difference \footnote{Notice the difference
519: with the case of pure gauge \cite{cv,dv3} where one of the
520: derivatives of $ydx$ with respect to the $S_i$ is a meromorphic
521: differential with a simple pole. This difference is consistent with
522: the fact that $\Sigma$ has genus$N$, while the hyperelliptic curve
523: considered in \cite{cv,dv3} has only genus $N-1$.}. 
524: From equation~(\ref{Si}) it also
525: follows that the $\omega_i$ form a basis of canonically normalized
526: holomorphic differentials: $\int_{A_i}\omega_j=\delta_{ij}$. Minimizing
527: equation~(\ref{eq:Weff}) with respect to $S_i$ we obtain
528: \begin{equation}
529: \sum_i N_i \int_{B_i} \omega_j =\int_{B_j} \sum_i N_i \,\omega_i=\tau \ ,
530: \label{mimim}
531: \end{equation}
532: where we used the symmetry of the period matrix $\int_{B_j}\omega_i$.
533: 
534: %The consequences of the minimization can be expressed in terms of
535: %the holomorphic differential $\Omega=\sum_i N_i \omega_i$.
536: The effects of the minimization translate into the behavior of
537: the holomorphic differential $\Omega=\sum_i N_i \omega_i$.
538: Its $A$-periods are $\int_{A_i}\Omega=N_i$. The minimization tells us
539: that, on-shell, the $B$-periods of $\Omega$ are 
540: all equal, $\int_{B_i}\Omega=\tau$. 
541: %It is actually convenient to consider as in \cite{cv} variables $S, S_{i,i+1}$
542: %such that $\de/\de S=\sum_i N_i \,\de/\de S_i$ and $\de/\de S_{i,i+1}$=
543: %$\de/\de S_i -\de/\de S_{i+1}$. In these terms the result of the minimization
544: %is
545: %\begin{equation}
546: %  \label{eq:min}
547: %  \begin{array}{c}\vspace{.2cm}
548: %  \sum_j N_j\int_{A_j} \frac{\de G}{\de S}dx= 1\ ,\quad
549: %  \sum_j N_j\int_{A_j} \frac{\de G}{\de S_{i,i+1}}dx= 0\ ; \\ 
550: %   \sum_j N_j\int_{B_j} \frac{\de G}{\de S}dx= \tau\ ,\quad
551: %  \sum_j N_j\int_{B_j} \frac{\de G}{\de S_{i,i+1}}dx= 0\ .
552: %  \end{array}
553: %\end{equation}
554: %From the definition of $G$ (\ref{eq:G}), 
555: %one sees that 
556: %\begin{equation} 
557: %\omega_0=\de G /\de S,\qquad  \omega_i=\de G/\de S_{i,i+1}, i=1,..,n-1
558: %\label{holo}
559: %\end{equation}
560: %are a basis of holomorphic differentials on the surface.
561: %(\ref{eq:min}) therefore tells us that one line of the period matrix is
562: %$(\tau, 0,\ldots,0)$. {\bf qualche parola in piu'}
563: To see what this means, consider the map
564: \begin{equation}
565:   \label{eq:z}
566: P\quad \mapsto \quad z(P)\equiv\int_{P_0}^P \Omega \,
567: %\int_{P_0}^P \frac{\de G}{\de S_{i+1,i}}dx  
568: \end{equation}
569: from the Riemann surface to $\cc$. 
570: %The map (\ref{eq:z})
571: %can be considered as an ``incomplete Jacobi map'', which, for
572: %a generic Riemann surface, makes no sense at all. However, in our
573: %case, since all the $B$-period are equal and all the $A$-period
574: %are integers, we see that (\ref{eq:z}) is well defined provided
575: %we make the identifications
576: The map (\ref{eq:z}) could be considered an ``incomplete Jacobi map'', in the
577: sense that we only consider the integral of one of the 
578: holomorphic differentials.
579: % another peculiarity is that the
580: %image is $\cc$ and not a compact manifold. 
581: For a generic Riemann surface, one would not be able to do better and identify 
582: points to make the image compact. However, in our
583: case, since all the $B$-periods are equal and all the $A$-periods
584: are integers, (\ref{eq:z}) is a well--defined map to a torus 
585: defined by the identifications
586: \begin{equation}
587: z\sim z+\tau\, \ , \qquad\qquad z\sim z+\tilde N\ , 
588: \label{torus}
589: \end{equation}
590: where $\tilde N$ is the highest common factor of the $N_i$.
591: %Since the
592: %period matrix is symmetric, the holomorphic
593: %differential $\de G/\de S$ has periods only on the cycles 
594: %$A\equiv\sum_i N_i A_i$ and $B\equiv\sum_i N_i B_i$. 
595: %If we start from one point on $\Sigma$ and follow $z$ until we are back, we
596: %will always find the same value if the cycle we followed is anything different
597: %from $A$ and $B$. If we go around these two latter cycle, $z$ does change by
598: %1 and $\tau$ respectively. So the map (\ref{eq:z}) is well defined if we
599: %identify $z\sim z+1 \sim z+\tau$, which means that the image is actually
600: %$T^2$. We also note that 
601: %(\ref{eq:z}) is then an ``incomplete Jacobi map'', and the $T^2$ we just
602: %described is naturally a sub-manifold of the Jacobian of $\Sigma$, a complex
603: %torus of dimension $N$ in which $\Sigma$ can be embedded using the 
604: %usual Jacobi map.
605: Thus, we have shown that using the equations of motion for $S_i$, 
606: $\Sigma$ becomes a covering of a torus of modular parameter 
607: $\tau/\tilde N$. 
608: %The cycles of this torus are $A$ and 
609: %$B$, and the differential $\de G/\de S \,dx= dz$. 
610: 
611: \subsection{The $\nn=2$ case}
612: To understand our result better, it is convenient to consider
613: first the case $n=N$, maximal number of cuts $N$ and $N_i=1$.
614: The gauge group is completely broken to the maximal abelian
615: subgroup $U(1)^N$. This is the 
616: situation where we expect to recover information about the
617: underlying $\nn=2$ theory. Following \cite{cv},
618: we take a potential $W(\Phi)$ of degree $N+1$ 
619: with $W'(x)=\epsilon \, \Pi_i (x -\phi_i)$. Since all $N_i=1$,
620: the eigenvalues of $\Phi$ are all distinct and 
621: classically coincide with the $N$ number $\phi_i$. 
622: The $\nn=1$ theory we are considering differs from an
623: $\nn=2$ theory only for the presence of
624: the potential $W(\Phi)$. 
625: If we turn off the deformation by sending $\epsilon\to 0$, 
626: we recover the $\nn=2$ theory in the point 
627: of the moduli space specified by the VEVs $\phi_i$ \cite{cv}.
628: %To discuss this case, we consider $n=N$ cuts
629: %and all $N_i=1$. In particular $\tilde N=1$.
630: %We will now show that, after minimization, 
631: %the matrix model Riemann surface $\Sigma$ becomes the Donagi-Witten
632: %curve describing the $\nn=2$ theory.
633: %
634: %In order to do so, we need some basic information about the Donagi-Witten
635: %curve \cite{dw} associated with a $U(N)$ gauge group. 
636: 
637: The $\nn=2$ theory we obtain in this way has gauge group $U(N)$,
638: a massive adjoint hypermultiplet \cite{dw}
639: and coupling constant $\tau$. The corresponding Seiberg-Witten curve 
640: %for this $\nn=2$ theory 
641: was determined in \cite{dw}. 
642: The curve can be written as an $N$-sheeted covering 
643: of a base torus (of modular parameter $\tau$) expressed by the equation 
644: \begin{equation}
645: F_N(v,z)={\rm det} (v-\Phi(z))=0\ ,
646: \label{dwcurve}
647: \end{equation}
648: where $\Phi$ is a section of a $U(N)$-Higgs-bundle on the torus.
649: The theory is a massive deformation of $\nn=4$ SYM and the 
650: presence of the torus reflects the S-duality of the original $\nn=4$
651: theory. Equation~(\ref{dwcurve}) gives the Donagi-Witten curve as a
652: degree $N$ polynomial in $v$ with coefficients that are elliptic 
653: functions on the base torus. 
654: The curve~(\ref{dwcurve}) is uniquely characterized by the existence of
655: a meromorphic function, $v$, with $N$ single poles 
656: at the $N$ counter-images $p_i$ 
657: of the point $z=0$ on the base torus. In one of the points, $p_0$, 
658: the residue has to be $-(N-1)m$, while the other $N-1$ points 
659: have residue $m$. The existence of a meromorphic function with these
660: properties uniquely determines the curve~(\ref{dwcurve}) \cite{dw,dph}.
661: 
662: We will now show that, after minimization, 
663: the matrix model Riemann surface $\Sigma$ for $n=N$ and $N_i=1$
664: becomes the Donagi-Witten 
665: curve describing the $\nn=2$ theory.
666: We have already seen that
667: the surface $\Sigma$ becomes a covering of a torus
668: of modular parameter $\tau$ ($\tilde N=1$ in this case). 
669: We will shortly see that the order $\tilde n$ of 
670: this covering is actually $N$. We can 
671: construct, on shell, the meromorphic function $v$ 
672: that characterizes a Donagi-Witten curve.
673: It is easier to find first the differential $dv$, which 
674: should have poles of order 2 with coefficient equal to minus the 
675: residue of $v$. 
676: We have at our disposal at least two differentials 
677: with double poles in some or all the points $p_i$. 
678: One is the pull-back of the differential $\PP(z) dz$ on the base torus. 
679: %where we now call $\Omega=dz$.
680: Once pulled back from $T^2$ to
681: $\Sigma$, this indeed provides double poles with coefficient one on 
682: every point $p_i$ over $z=0$.
683: Another one is $dx$ which has a double pole at $x=\infty$. 
684: It is natural to suppose that $x=\infty$ corresponds to $p_0$\footnote{Notice 
685: that $dz$ is not vanishing in $x=\infty$. 
686: This is consistent with
687: the fact that the point at infinity is not a branch point in the
688: Donagi-Witten curve.}.
689: % is nothing but the point $x=\infty$ of the
690: %matrix model, and to add $dx$ to $\PP(z)dz$. 
691: We can construct a differential with the desired properties
692: by adding $dx$ and $\PP(z)dz$. This easily provides us with
693: a differential with double poles in $p_i$ with coefficients
694: $-m(-\tilde n+1,1,...,1)$.
695: We should also require $dv$ to
696: have zero periods, since we want to integrate it to give a well
697: defined meromorphic function $v$ on $\Sigma$.
698: %The problem of this proposal is
699: %that the sum of these two pieces has periods, and that would not make a
700: %good candidate for $dv$, which should not have any periods. 
701: We can still add a piece with $dz$ without adding any pole. 
702: This way we end up with a differential $adx-m(\PP(z)+b)dz$, 
703: where the constants $a$ and $b$ have to be 
704: determined imposing the vanishing of all periods. 
705: Only two of the $2N$ conditions on periods are non trivial. 
706: %A priori this might seem 
707: %difficult since $\Sigma$ has 2$N$ periods. 
708: Indeed, the pull-back of 
709: any $f(z)dz$ has possibly non--vanishing periods only around the two cycles of $T^2$. 
710: The same statement holds for $dx$; since the periods of $dx$ are all 
711: equal (\ref{eq:dx}), the only 
712: independent non--vanishing period is around the cycle $B$, corresponding 
713: to one of the 
714: cycle of the base torus. 
715: Thus, we reduce to only 
716: two equations, which determine uniquely $a$ and $b$.
717: The result is
718: %La parte commentata ha gli $\omega_1$
719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
720: %\begin{equation}
721: %  \label{eq:dv}
722: %  dv = \frac{2 \pi i}{m \omega_1} dx + \frac1{\omega_1^2}
723: %\Big(\PP(z) -\frac{\pi^2}3 E_2(\tau) \Big) dz\ ,
724: %\end{equation}
725: %where $E_2(\tau)$ is a standard Eisenstein series. This can be 
726: %integrated as
727: %\begin{equation}
728: %  \label{eq:v}
729: %  v= \frac{2 \pi i}{m \omega_1} x + 
730: %\frac\pi{2\omega_1}\frac{\theta_1'(z|\tau)}{\theta_1(z|\tau)}
731: %-\frac{\pi^2 E_2(\tau)}{4\omega_1^2}z 
732: %\end{equation}
733: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
734: \begin{equation}
735:   \label{eq:dv}
736:   dv = 2 \pi dx - m
737: \Big(\PP(z) +\frac{\pi^2}3 E_2(\tau) \Big) dz\ ,
738: \end{equation}
739: where $E_2(\tau)$ is a standard Eisenstein series. This can be 
740: integrated up to a constant to give \footnote{Here $h_1(z)=\zeta(z)-\frac{\zeta(\omega_1)}{\omega_1} z$ for a torus with periods $2\omega_1$ and $2\omega_2$.
741: $\zeta(z)$ is the Weierstrass zeta function: it is defined by
742: having a simple pole in $z=0$ and quasi-periodicity properties $\zeta(z+2\omega_i)=\zeta(z)+2\zeta(\omega_i)$. $h_1(z)$ is periodic along $A$ and
743: $h_1(z+2\omega_2)=h_1(z)-\frac{\pi i}{\omega_1}$. Other useful
744: identities are: $\zeta'(z)=-\PP(z)$, $\omega_2\zeta(\omega_1)-\omega_1\zeta(\omega_2)=\pi i/2$, $\zeta(\omega_1)\omega_1=\pi^2E_2(\tau)/12$.}
745: \begin{equation}
746:   \label{eq:v}
747:   v= 2 \pi x + 
748: m\pi\frac{\theta_1'(\pi z|\tau)}{\theta_1(\pi z|\tau)}
749: \equiv 2\pi x +mh_1(z)\  .
750: \end{equation}
751: This expression is well-defined since $x$ and $h_1(z)$ are
752: both periodic along the $A$ cycle and the coefficients are chosen
753: in order to cancel their jump along $B$. 
754: %This expression is very similar to the one between $k$ and $\tilde k$ in the
755: %expression of the Donagi-Witten curve as coming from the Lax matrix of the
756: %elliptic Calogero-Moser system \cite{dhp}. We will comment more about
757: %this later. 
758: %Notice that we have not bothered
759: %to check that the resulting $v$ has the right pole in $p_0$: this is 
760: %automatically enforced by the vanishing of the sum of residues of the
761: %differential $vdz$. The implicit assumption in this argument
762: %is that the number of sheets is really $N$. 
763: To determine the order $\tilde n$ of the covering, it is now enough to
764: compute the value of the residue of $x$ at $x=\infty$. 
765: From equation~(\ref{eq:G}) it follows 
766: \begin{equation}
767: \frac{\de}{\de S_i}Gdx=m\frac{dx}{x^2} \ ,
768: \end{equation}
769: so that $dz\equiv\Omega= (mN/2\pi)dx/x^2$. It follows that, 
770: in local coordinates,
771: $x=- mN/(2\pi \, z)$. This fixes the residue of $v$ near $x=\infty$ to be 
772: $-(N-1)m$. Thus $\tilde n=N$, and our proof is completed.
773: 
774: Once the algebraic curve is given, eq.~(\ref{eq:v}) determines the
775: map to the matrix model plane $x$. The function $G$ is also 
776: uniquely determined by the requirement of having a single pole of order
777: $N+1$ at $x=\infty$, even though its explicit form can be difficult
778: to find. Using the results of \cite{dw},  we can determine  the expression 
779: of the curve and $G$ for small values of $N$. The spectral curve 
780: (\ref{dwcurve}) can be written as a pair of equations \cite{dw}
781: \begin{eqnarray}
782: &&y^2 = (t-e_1)(t-e_2)(t-e_3) \ ,\nonumber\\
783: &&F_N(v,t,y)=0 \ ,
784: \label{covering}
785: \end{eqnarray} 
786: where the first equation is the standard representation 
787: of the torus as a cubic, while the second 
788: is a polynomial of degree $N$ in $v$, giving the $N$-sheeted covering 
789: of the torus. 
790: As shown in \cite{dw}, $F$ is also a polynomial in $x$ and $y$.
791: 
792: For example, for $N=2$ and $W'(\Phi)=\Phi^2-A_2$, 
793: %which is already nontrivial if one considers
794: %two cuts, 
795: %by imposing as we said above that $G$ have only a
796: %single pole of order $N$ in $p_0$, 
797: we have 
798: \begin{eqnarray}
799: &&F_2(v,x,y)= v^2 -t - A_2 \ , \nonumber\\
800: &&G=y +v^3 -\frac{3A_2}{2} v \ .
801: \end{eqnarray}
802: %$A_2$ is the only SU(2) modulus.
803: Analogously, for $N=3$ and $W'(\Phi)=\Phi^3-A_2\Phi-A_3$
804: we have 
805: \begin{eqnarray}
806: &&F_3(v,x,y)= v^3 -v(3t+A_2) +2y -A_3\ ,\nonumber\\
807: &&G=v^4 -v^2(6t-\frac{2}{3}A_2) +8vy - 3t^2-\frac{2}{3}A_2 t \ .
808: \end{eqnarray}
809: The degree $N$ 
810: polynomial in $F$ can be related to $W'$ by analyzing the large $x$
811: behavior of $G$ (see equation \ref{eq:G}).
812: %{\bf In general}, a way of finding $G$
813: %would be to get its dependence on $v$ via (\ref{eq:G}) and
814: %(\ref{eq:v}), and determine afterwards its dependence on $t$ and
815: %$y$. {\bf Forse ci si pu{\`o} pensare\ldots}
816: In both cases, $x$ is determined by equation~(\ref{eq:v}). The structure
817: of these two examples suggests that in general $G$ is a linear combination of
818: the polynomials $P_k$ defined in \cite{dw}.
819: 
820: For generic $N$ it is probably more convenient to use
821: an alternative expression for the Donagi-Witten curve \cite{dph},
822: \begin{equation}
823: \sum_{n=-\infty}^\infty (-)^n q^{\frac{1}{2}n(n-1)} e^{nz}H(x-nm)=0 \ , 
824: \label{dh}
825: \end{equation}
826: where $H$ is a degree $N$ polynomial, which we can roughly identify
827: with $W'$. 
828: Another advantage of this expression is that it is naturally 
829: written in terms of the matrix model variable $x$. Indeed,
830: as shown in \cite{dph},
831: to go from the polynomial equation~(\ref{covering}) to the expression
832: above, a change of variables $v\rightarrow v-mh_1(z)$ is required, 
833: which, by equation~(\ref{eq:v}) 
834: exactly defines the function $x$. 
835: 
836: %In this formalism,
837: %the function $G$ is explicitly given by
838: %{\bf formula per G hopefully....} 
839: 
840: \subsection{The general case}
841: We can also study the general case with arbitrary $n$ and $N$.
842: In this case, the group is broken to $\prod_{i=1}^n U(N_i)$ and
843: there are some non-abelian gauge factors at low energy.
844: For these vacua, we could also introduce, as in \cite{csw2},
845: integer numbers $b_i$ labeling the type of confinement in each factor.
846: The numbers $b_i$ appear in the matrix model expression for $W_{eff}$
847: as \cite{csw2}
848: \begin{equation}
849:   \label{eq:Weff2}
850:   W_{\rm eff}= 
851: \sum_i \left( N_i \frac{\de {\cal F}}{\de S_i} - 2\pi i\tau S_i\right ) 
852: - \sum_{i=2}^n 2\pi i b_i S_i\ .
853: \end{equation}
854: The minimization procedure then fixes the periods of $\Omega$
855: to be  $N_i$ and $\tau+b_i$. As before, we conclude that the map
856: $z:\Sigma\rightarrow \mathbb{C}$ is well defined if we make the
857: identifications $z\sim z+\tau$ and $z\sim z+t$, where $t$
858: is the highest common factor of the integers $N_i$ and $b_i$.
859: It was shown in \cite{csw2} that $t$ defines the index of confinement
860: of the vacuum.
861: 
862: We see that $\Sigma$ becomes a covering of a torus of modular parameter
863: $\tau/t$. We can further show that $\Sigma$ is a 
864: $N/t$-sheeted covering
865: of the base torus, and express $\Sigma$ as an
866: algebraic equation. Indeed, the argument  we used to identify the function
867: $v$ can be repeated almost verbatim in the general case.
868: The only difference is now the ratio
869: of the residues in the points $p_i$: 
870: equation~(\ref{eq:v}) is replaced by
871: \begin{equation}
872: v=\frac{2\pi}{t} dx +mh_1(z) \ .
873: \label{eq:v2}
874: \end{equation}
875: Since the behavior of $x$ at infinity is still given by $x= - mN/(2\pi \, z )$,
876: the residue at infinity is now $N/t$. It follows
877: that the number of sheets is $\tilde n=N/t$. The genus $n$ 
878: curve $\Sigma$
879: is then expressed as an element of the spectral family $F_{\tilde n}(v,z)=0$.
880: Since the arithmetic genus of a curve of the family $F_{\tilde n}(v,z)=0$ is
881: $\tilde n>n$, the algebraic curve will be singular and $\Sigma$ will
882: correspond to its normalization.    
883: 
884: This result deserves some comments.
885: We know that the relevant geometry 
886: for arbitrary $n$ and $N$ can be determined as a particular point
887: in the moduli space of the underlying $U(N)$ $\nn=2$ theory.
888: In a vacuum with $\prod_{i=1}^n U(N_i)$ only $n$ photons remain massless
889: and this requires that $N-n$ monopoles are massless and condensate
890: in order to give mass to all the other degrees of freedom. 
891: The associated curve has then $N-n$ nodes. The
892: sub-variety of the moduli space where $N-n$ monopoles are massless
893: has dimension $n$. The point in this sub-variety associated to the
894: $\nn=1$ vacuum can be determined by explicitly solving the equations
895: of motion for the potential $W(\Phi)$. This minimization was
896: explicitly done in the case of a pure gauge theory in \cite{int}.
897: In that case,
898: one can explicitly check that a potential $W(\Phi)$ of degree $n+1$
899: selects a point in the $\nn=2$ moduli space of the form
900: \begin{equation}
901: y^2=P_N(x)^2-1=\prod_{k=1}^{N-n}(x-u_i)^2 (W'(x)^2+f_{n-1}) \ , 
902: \label{curvepure}
903: \end{equation}
904: where the $N-n$ double zeros correspond to a degeneracy of the curve
905: associated with $N-n$ massless monopoles. The reduced genus $n-1$ curve 
906: $y^2=(W'(x)^2+f_{n-1})$ is exactly that describing the matrix model geometry.
907: It would be interesting to repeat this analysis for the case with
908: a massive hypermultiplet. The $\nn=1$ vacuum $\prod_{i=1}^n U(N_i)$
909: should be associated with an element of the spectral family
910: $F_N(v,z)=0$ with $N-n$ nodes, thus degenerating to a genus $n$ surface. 
911: The matrix model computation suggests that the normalization of such curve 
912: can be also find by  normalizing a  curve $F_{\tilde n}(v,z)$ of 
913: Donagi-Witten type
914: but with a different base torus of modular parameter $\tau/t$.
915: %It would be interesting to explicitly check this result starting
916: %from the $\nn=2$ theory.
917: 
918: Finally we would like to comment about the 
919: cases where some of the cuts coincide.
920: In such cases, the genus $n$ curve $\Sigma$ degenerates.
921: The normalization of the resulting curve
922: has genus equal to the number of cuts. We can for example make contact
923: with the examples discussed in \cite{dv3,dhks}. These 
924: correspond to completely massive vacua and are associated with
925: a single cut. The previous discussion can be repeated with
926: $n$ replaced by the number of cuts ($n\rightarrow 1$) and
927: the only non trivial $N_i=N$. We exhibited $\Sigma$ 
928: as a $1$-sheeted covering (i.e. an isomorphism) of a torus 
929: of modular parameter $\tau/N$. 
930: For $n=1$, the Donagi-Witten curve becomes quite
931: trivial, $F_1(v,x,y)= v-A$, and indeed describes a torus. 
932: Equation~(\ref{eq:v}) then gives 
933: $\frac{2 \pi }m  x =
934: A -\pi  \frac{\theta_1'(z|\tau)}{\theta_1(z|\tau)}
935: $. 
936: This is exactly 
937: the map from the torus to the matrix model plane 
938: given in \cite{dv3,dhks}. 
939: %Moreover, equation~(\ref{eq:min})
940: %adapted to the case of $N$ colors and a single cut gives 
941: %$\int_B \omega_o =\tau/N$. Therefore, as in \cite{dv3,dhks},
942: %the matrix model plane is mapped to a torus of modular parameter
943: %$\tau/N$. 
944: The field theory interpretation of this result is simple.
945: Indeed, as it is well known \cite{dw}, the massive vacua
946: of the $U(N)$ theory with bare coupling $\tau$ 
947: correspond to points in the moduli space of the $\nn=2$ theory 
948: where the genus $N$ curve $F_N(v,x,y)=0$ maximally degenerates
949: and becomes a torus of modular parameter
950: $\tau/N$. 
951: 
952: 
953: %Another point deserves a comment. In the general case the mapping $x$ between
954: %the matrix model plane and the Riemann surface $\Sigma$ 
955: %has a priori nothing
956: %to do with the function $G$. In the single cut case and with quadratic $W$
957: %it so happened that (up to a constant) $G(x)=\PP(z)$, 
958: %the Weierstrass function on the torus, which is also one of the coordinates 
959: %$(t,y)$ in the standard embedding. Thus in a way $G$ and $dx$ happened to be 
960: %related. Though this particular relation cannot be true in general, the 
961: %properties of $G$ are still enough to find it explicitly after having gone 
962: %to the $\nn=2$ limit, as we are going to see {\bf in examples}.
963: 
964: 
965: 
966: 
967: 
968: \newsection{Resolvents}
969: \label{sec:res}
970: 
971: The interesting quantities to compute in quantum field theory
972: are the resolvents \cite{cdsw}
973: \begin{eqnarray}
974: R(x)&=&{\rm Tr}(\frac{{\cal W}_\alpha {\cal W}^\alpha}{ x-\Phi})\ ,
975: \nonumber\\
976: T(x)&=&{\rm Tr}(\frac{1}{x-\Phi})\ .
977: \label{QFTresolvent}
978: \end{eqnarray}
979: The knowledge of $R(x)$ and $T(x)$ allows to compute all
980: the vacuum expectation values of operators in the chiral
981: ring \cite{cdsw}. In this section, we will discuss a possible
982: identification of $R$ and $T$ with geometrical quantities. 
983: In the case of pure gauge, $R$ can be related to the meromorphic function
984: $y$ defining the curve (see equation~(\ref{riemann1})) while $T(x)dx$
985: becomes a meromorphic differential on the curve \cite{cv,dv3,cdsw,gopa,csw3}.
986:  
987: \subsection{Identification of the resolvents} 
988: We expect that, as in \cite{cdsw}, $R(x)$ is identified with 
989: the matrix model resolvent 
990: \begin{equation}
991: R(x)= S \, \omega (x) \ . 
992: \label{res}
993: \end{equation}
994: In the case of pure gauge, 
995: this identification descends from  the comparison of the matrix model
996: loop equations with the Ward identities in field theory \cite{cdsw}.
997: As shown in \cite{cdsw}, the equations for $R(x)$,  which  
998: can be deduced in the quantum field
999: theory using the Konishi anomaly, are formally identical
1000: to the matrix model Ward identities for $\omega(x)$. 
1001: In our case, the Ward identities are more complicated 
1002: to write (some explicit
1003: relations are discussed in the Appendix), but we expect that 
1004: the general philosophy still applies.
1005: %this relation is derived
1006: %by comparison of the matrix model loop equation with
1007: %the quantum field theory Ward identities \cite{cdsw}.
1008: 
1009: Quantum mechanically, we may expect that
1010: small cuts are opened around the classical eigenvalues of $\Phi$,
1011: analogously to what happens for the matrix model. Integrals of
1012: $R(x)$ around a critical point of $W$ define the condensates 
1013: ${\rm Tr}(W_{\alpha}W^{\alpha})_{(i)}$.
1014: The contour integrals of $R$ around a cut 
1015:  are mapped to integrals of
1016: $\omega(x)$ around its cut in the matrix model plane (these cuts are
1017: indicated with dotted lines in figure 2),
1018: which define the quantities $S_i$
1019: \begin{equation}
1020: {\rm Tr}(W_{\alpha}W^{\alpha})_{(i)}=
1021: \frac{1}{2\pi i} \oint R(x) dx =\frac{1}{2\pi i} \oint S \, \omega(x) dx=
1022: S_i\ . 
1023: \label{R}
1024: \end{equation}
1025: 
1026: %\begin{figure}[h]
1027: %\centerline{{\small\fbox{$x$}}
1028: %\includegraphics[width=13cm,height=4.5cm]{mm2.eps}}
1029: %\caption{\small The $C_i$ contours encircle the cuts of $\omega(x)$,
1030: %denoted by dotted lines.
1031: %The contours $A,A^*,B,B^*$ delimit a simply connected region in the $x$ plane.}
1032: %\label{cuts2}
1033: %\end{figure}
1034: \begin{figure}[h]
1035: \begin{picture}(80,140)(-20,-5)
1036: \put(0,0){{\small\fbox{$x$}}}
1037: \put(15,10){\includegraphics[width=13cm,height=4.5cm]{mm2.eps}}
1038: \put(40,140){$A$}\put(40,0){$A^*$}\put(40,85){$C$}
1039: \put(115,32){$B$}\put(105,51){$B^*$}
1040: \end{picture}
1041: \caption{\small The $C_i$ contours encircle the cuts of $\omega(x)$,
1042: denoted by dotted lines.
1043: The contours $A,A^*,B,B^*$ delimit a simply connected region in the $x$ plane.}
1044: \label{cuts2}
1045: \end{figure}
1046: 
1047: From quantum field theory it is also obvious that the integrals of
1048: $T(x)$ around a critical point of $W$ are integers
1049: \begin{equation}
1050: \frac{1}{2\pi i} \oint T(x) dx =N_i \ .
1051: \label{T}
1052: \end{equation}
1053: In analogy with \cite{csw3}, we may expect $T(x)$ to be associated
1054: with a differential on the Riemann surface.
1055: Consider the differential 
1056: \begin{equation}
1057: t(x)dx=\Big[T\Big(x-\frac{i}{2}m\Big)- T\Big(x+\frac{i}{2}m\Big)\Big]dx\ ;
1058: \label{t}
1059: \end{equation}
1060: in a similar way as for the matrix model resolvent, the residue 
1061: cancels from the two $\pm im/2$ pieces, and the differential is 
1062: holomorphic around $x=\infty$. We now conjecture that $t(x)dx$ can be extended
1063: to a holomorphic differential defined on the entire Riemann surface.
1064: Its periods  
1065: around the $A_i$ cycles are $N_i$, by equation~(\ref{T}).
1066: %the usual argument about the jump in the
1067: %resolvent being equal to the distribution ({\bf assumption} about distribution
1068: %of eigenvalues having the same support as the mm one) nothing but the 
1069: %number of eigenvalues of $\Phi$ in the $i$th vacuum, $N_i$. 
1070: These are the same $A_i$ periods as $dz$. 
1071: Since an holomorphic differential is completely specified by its
1072: $A_i$ periods
1073: %it then follows that also the $B_i$ periods
1074: %coincide 
1075: we conclude that
1076: \begin{equation}
1077:   \label{eq:dzFT}
1078: \Omega\equiv dz=\frac{1}{2\pi i}t(x) dx=\frac{1}{2\pi
1079:   i}\Big[T\Big(x-\frac{i}{2}m\Big)- T\Big(x+\frac{i}{2}m\Big)\Big] dx\ .
1080: \end{equation}
1081: It then follows that also the $B_i$ periods
1082: are completely specified (this argument was also 
1083: used in a similar way in \cite{csw3})
1084: \begin{equation}
1085: \frac{1}{2\pi i} \int_{B_i} t(x) dx =\tau \ .
1086: \label{T2}
1087: \end{equation}
1088: 
1089: The validity of equation~(\ref{eq:dzFT}) is strengthened by the following
1090: formal %but suggestive 
1091: argument. Using the definition of $dz$ and
1092: the identification~(\ref{res}), we have
1093: \begin{equation}
1094: dz=\frac{1}{2\pi}\sum_iN_i\frac{d}{dS_i}Gdx=-\frac{1}{2\pi i}
1095: \sum_iN_i\frac{d}{dS_i}\Big[R\Big(x+i\frac{m}{2}\Big)
1096: -R\Big(x-i\frac{m}{2}\Big)\Big] \ ,
1097: \label{di}
1098: \end{equation}
1099: which, by comparison with~(\ref{eq:dzFT}), leads to the suggestive equation
1100: \begin{equation}
1101: \label{eq:dG}
1102: \frac1N \sum_i N_i \frac\de{\de S_i} {\rm Tr}
1103: \Big( {\cal W}_\alpha {\cal W}^\alpha 
1104: \frac1{x-\Phi}\Big)
1105: ={\rm Tr}\Big( \frac1{x-\Phi}\Big) \ .
1106: \end{equation}
1107: %{\bf up to a quantity} periodic under $x\to x+im$. 
1108: %Finally, we also know that the equations for the 
1109: %resolvents $T(x)$ always have the form of a
1110: %derivative of the ones for $R(x)$, due essentially to the ``superfield''
1111: %formalism of \cite{cdsw}. 
1112: %One can compare this procedure with actually
1113: %hitting the equation for the resolvent with ${\cal W}_\alpha{\cal
1114: %  W}^\alpha$ by $\de/\de S$ gives the same equation, and thus
1115: %(\ref{eq:dG}) strongly suggests itself.}
1116: %This equation of course 
1117: %could have formally guessed, and it is remarkable that it gets confirmed by
1118: %geometrical techniques 
1119: %
1120: %In the case of pure gauge, $T(x)$ are then obtained
1121: %by differentiating those for $R(x)$, as a consequence of the
1122: %superfield formalism introduced in \cite{cdsw}.
1123: 
1124: \subsection{Field theory expectation values} 
1125: We can use $t(x)dx$ to compute the field theory expectation values 
1126: of $\langle{\rm Tr} \Phi^k\rangle$. To this purpose, we can integrate
1127: $x^k t(x)dx$ on a small contour around $x=\infty$,
1128: \begin{equation}
1129: \langle{\rm Tr} \Big[(\Phi+\frac{i}{2}m)^k-(\Phi-\frac{i}{2}m)^k \Big]\rangle=
1130: \int_{\infty} x^k dz \ .
1131: \label{moments}
1132: \end{equation}
1133: We can deform the previous contour integral in the $x$ plane 
1134: until it encircles the cycles $A_i$ and $A_i^*$ (see figure 2), 
1135: obtaining
1136: \begin{equation}
1137: \int_{\infty} x^k dz= \sum_i (\int_{A_i}x^k dz+ \int_{A_i^*}x^k dz)=
1138: \sum_i \int \Big[(x_0^{(i)}+i\frac{m}{2})^k-(x_0^{(i)}-i\frac{m}{2})^k\Big]dz\ ,
1139: \label{moments2}
1140: \end{equation}
1141: where $x_0^{(i)}(z)$ are maps from the $A$ cycle of the base torus
1142: to contours  around the cuts of the resolvents on the real axis
1143: (indicated by dots in figure \ref{cuts2}). 
1144: We call $C_i$ these contours.
1145: Comparing equations~(\ref{moments}) and~(\ref{moments2})
1146: we obtain the useful formula
1147: \begin{equation}
1148: \langle{\rm Tr}\Phi^k\rangle = \sum_i \int_{C_i} x^k dz =\sum_i
1149: \int x_0^{(i)}(z)^k dz\ .
1150: \label{condensate}
1151: \end{equation}
1152: %$x_0(z)$ can be interpreted as the quantum distribution of field theory
1153: %eigenvalues. 
1154: Formula~(\ref{condensate}) was derived in \cite{dhks}
1155: in the case of a single cut. 
1156: 
1157: 
1158: 
1159: We can also compute the vacuum value of the effective potential 
1160: in terms of the function $x_0$. This will also strengthen our 
1161: identification of $t(x)dx$ with $dz$.
1162: On shell, we can write
1163: $W_{\rm eff}$ as
1164: \[
1165: \sum_i \left( \int_{A_i}dz \int_{B_i} G\, dx - 
1166:     \int_{B_i}dz \int_{A_i} G\, dx \right)\ .
1167: \]
1168: By Riemann bilinear relations \cite{gh}, this expression is also equal to
1169: Res$_\infty (z(x) G(x)dx)$. The $U$ piece in the definition (\ref{eq:G}) of $G$
1170: is the only one which can contribute to this residue. 
1171: We can at this point
1172: try to invert the proof of the Riemann bilinear relations. To this purpose
1173: we choose a base point, say on the real 
1174: axis of $x$, and modify all the $A_i$ and $B_i$ periods in such a way that
1175: they bound a simply connected region (see figure \ref{cuts2}). 
1176: This way we build a polygon whose sides 
1177: are $A_i$, $B_i$ and their opposites $A_i^*$, $B_i^*$, 
1178: identified in such a way as to reconstruct the Riemann surface $\Sigma$,
1179: similarly to \cite{gh}. In the interior of this polygon 
1180: (a simply connected region) the function $z$ is 
1181: well-defined. The Riemann bilinear relations are then demonstrated by
1182: deforming  a contour integral of $z \, Gdx$ around $x=\infty$ to
1183: the perimeter of the polygon and by exploiting the periodicities of $z$.
1184: If we now substitute $G(x)$ 
1185: with $U(x)$, we obtain extra contributions 
1186: coming from the fact that $U(x)$ is a well defined function 
1187: on the plane $x$ but not on the Riemann surface. But we can now
1188: exploit that to our advantage:
1189: \begin{eqnarray}
1190: \label{eq:Ci}
1191: &&W_{\rm eff}={\rm Res}_\infty (z \,Udx)= 
1192: -\sum_i \left( \int_{A_i+A_i^*} z\,U dx + \int_{B_i+B_i^*} z\,U dx \right)
1193: \nonumber \\
1194: && = -\sum_i \left( \int_{A_i}(z+\tau)U(x_0+i\frac{m}{2})+\int_{A_i^*} z\,U(x_0-i
1195: \frac{m}{2}) dx - \int_{B_i} U dx \right )\nonumber\\
1196: && = -\sum_i\left(\int_{C_i}z W'(x)dx - \int_{B_i}U(x)dx\right) \nonumber\\
1197: && = \sum_i \left(\int_{C_i} W(x)dz\right) + N\int_{-im/2}^{im/2}U(x)dx \ .
1198: \end{eqnarray}
1199: In second line of \ref{eq:Ci}, we first used equation (\ref{eq:U}) and 
1200: the fact that  $U(x)$ is not well-defined on $\Sigma$ 
1201: %and equation (\ref{eq:U}) 
1202: to evaluate the integrals over $A_i$; the integrals over $B_i$
1203: instead just behave as for the usual Riemann argument. 
1204: We have also used that $\int_{A_i}U=0$. We then integrated
1205: by parts the integrals over $C_i$, and then used appropriate integrals of
1206: (\ref{eq:U}) again to put the second piece in the final form. 
1207: 
1208: 
1209: The final expression for $W_{\rm eff}$ in (\ref{eq:Ci}) is consistent
1210: with the identification~(\ref{condensate}). Indeed, if
1211: $W\equiv\sum_p g_p {\rm Tr}\Phi^p$,
1212: we know $W_{\rm eff}= \sum g_p \langle{\rm Tr}\Phi^p \rangle$. 
1213: So equation~(\ref{eq:Ci}) can be written as
1214: \begin{equation}
1215:   \label{eq:up}
1216:   \sum_i \int_{C_i} W(x)dz 
1217: \sim \langle\sum_k g_k{\rm Tr} \Phi^k\rangle\ ,
1218: \end{equation}
1219: modulo pieces which, for a given $W$, 
1220: depend on $g_k$, but not on $\tau$ and on the
1221: choice of a particular vacuum. Formula
1222: (\ref{eq:Ci})
1223: was derived in a different way in the one-cut case in \cite{dhks}.
1224: 
1225: Formula~(\ref{condensate}) gives a prescription
1226: for computing the quantities $<{\rm Tr}\Phi^k>$ purely in terms
1227: of matrix model data; the function
1228: $x_0(z)$ can be interpreted as the quantum distribution of field theory
1229: eigenvalues. We should note, however, that there is
1230: an ambiguity in the definition of $<{\rm Tr}\Phi^k>$
1231: in field theory. The condensates can be computed as the order
1232: parameters $u_k=<{\rm Tr}\Phi^k>_{\nn=2}$
1233: of the $\nn=2$ vacuum that is selected by the
1234: potential $W(\Phi)$. In presence of a mass $m$, $u_k$ can mix
1235: with all the other order parameters $u_j, j<k$ \cite{dw,ad}.
1236: The results for the condensates obtained with different methods
1237: could be related by a change of basis in the $u_i$; consistency
1238: requires the coefficients of such redefinition to be
1239: vacuum independent.
1240: In the case of a single cut, it was explicitly checked in \cite{dhks} 
1241: that the condensates computed with the matrix model prescription 
1242: are indeed related by a vacuum-independent redefinition to
1243: the condensate computed using the Donagi-Witten curve.
1244:   
1245: 
1246: \vskip .2in \noindent \textbf{Acknowledgments}\vskip .1in \noindent   
1247: We would like to thank Annamaria Sinkovics and 
1248: Stefan Theisen for interesting discussions.
1249: This work is partially supported by the EU contract
1250: HPRN--CT--2000--00122. M.P. is supported by the European Commission Marie
1251: Curie Postdoctoral Fellowship under contract number HPRN--CT--2001--01277.
1252: A. Z. are partially supported by INFN and MURST, and
1253: by the European Commission TMR program HPRN-CT-2000-00131,
1254: wherein he is associated to the University of Padova.
1255: 
1256: \section{Appendix: Loop equations}
1257: 
1258: The identifications of Section 4 could be strengthened
1259: by comparing the matrix model loop equations with the Konishi
1260: anomaly equations in field theory. A complete set of loop equations
1261: uniquely determining the resolvents would also give a convenient
1262: description of the off-shell theory. In the case of pure gauge,
1263: the loop equations for the matrix model give relation~(\ref{riemann1})
1264: for $y=W'(x)-2\omega(x)$, which uniquely determines the
1265: Riemann surface associated with the matrix model; this result
1266: is also valid off-shell. In the case of $\nn=1^*$, the
1267: loop equations are more complicated. 
1268: Here we will make a first step in the study of the
1269: loop equations and we
1270: derived a suggestive relation satisfied by the resolvent.
1271: In the following, we will refer to the Ward identity for the matrix model, 
1272: which are identities for $\omega(x)$. As shown in \cite{cdsw},
1273: identical equation for $R(x)$ can be deduced from the quantum field
1274: theory Ward identities. Identities for $T(x)$ are then obtained
1275: by differentiating those for $R(x)$, as a consequence of the
1276: superfield formalism introduced in \cite{cdsw}.
1277: 
1278: %We try now to extract more informations about the matrix model resolvent,
1279: %of which (\ref{eq:rs}) was a priori not all, as commented there. 
1280: The best
1281: way to find the loop equations is to consider the matrix model {\sl before}
1282: integrating out $\hat X$ and $\hat Y$ (\ref{eq:supot}), 
1283: and making appropriate changes of coordinates \cite{schnitzer,ragazzi}.
1284: The ones we consider here are
1285: \[
1286: \delta\hat \Phi= \frac1{x-\hat\Phi}\ , \qquad 
1287: \delta \hat X= \frac1{x-\hat\Phi+i\frac m2}\,\hat X\,\frac1{x-\hat\Phi-i\frac
1288:   m2} \ .
1289: \]
1290: The equations that we get from these are
1291: \
1292: \begin{equation}
1293: \label{eq:lxy}
1294: \begin{array}{ccl}\vspace{.2cm}
1295: && \omega^2(x)= - {\rm Tr}\Big[\Big(W'(\hat \Phi)+[\hat X,\hat Y]\Big)\frac1{x-\hat\Phi}\Big] \ ,
1296: \\ 
1297: && \omega\Big(x+i\frac m2\Big)\omega\Big(x-i\frac m2\Big)
1298: ={\rm Tr}\left(\hat Y\hat X\frac1{x-\hat\Phi-i\frac m2}-\hat X\hat Y\frac1{x-\hat\Phi+i\frac m2}\right)\ . 
1299: \end{array}
1300: \end{equation}
1301: 
1302: By considering repeated translations of (\ref{eq:lxy}) by $\pm im/2$, 
1303: we can find a combination in which $\hat X$ and $\hat Y$ disappear:
1304: \begin{equation}
1305: \label{eq:loop}
1306: \sum_{n=-\infty}^{\infty} \Big[\omega\Big(x+i(n+2)\frac m2\Big)-
1307:   \omega\Big(x+in \frac m2\Big)\Big]^2=
1308: 2\sum_{n=-\infty}^{\infty}{\rm Tr} \Big[\frac{W'(\hat \Phi)}{x-\hat
1309:   \Phi+in\frac m2} \Big]\ .
1310: \end{equation}
1311: One can actually write this equation in terms of $G$ by completing a
1312: square. 
1313: Defining  $X_n(x)\equiv X(x+in\frac m2)$ for any function $X(x)$, we have
1314: \begin{equation}
1315:   \label{eq:loopG}
1316:   \sum_{n=-\infty}^{\infty} G_n^2=\sum_{n=-\infty}^{\infty} U_n^2 - 
1317: 2\sum_{n=-\infty}^{\infty} f_n \ ,
1318: \end{equation}
1319: where we have defined 
1320: $f(x)= {\rm Tr} [W'(\hat \Phi)-W'(x))/(x-\hat \Phi)]$. 
1321: The formal analogy
1322: with the pure gauge case is evident. In that case, the loop equations are
1323: $y^2=W'^2+f$. In our case, $G$ plays the role of $y$ and $U$ the role
1324: of $W'$. This would become of course more
1325: than an analogy if one considers the limit $m\to \infty$.
1326: 
1327: It is not clear whether equation (\ref{eq:loopG}), involving
1328: and infinite series of shifts, is well-defined. If so, it
1329: could be interpreted as an equation determining $G$, and implicitly 
1330: $\Sigma$, in dependence of the polynomials $U$ and $f$. It would
1331: be interesting to investigate whether $G$ is uniquely determined 
1332: by such equation. Similar issues were studied in \cite{ber}.
1333: It would be also interesting to study 
1334: equation~(\ref{eq:loopG}) after minimization, where it should be
1335: related to the Donagi-Witten curve.
1336: The form of equation~(\ref{eq:loopG}) is very suggestive in comparison to the
1337: Lax matrix form of the Donagi-Witten curve given in equation~(\ref{dh}).
1338: %situation, as already noticed above, after (\ref{eq:v}). The formal
1339: %similarities of (\ref{eq:loopG}) with the equation for the curve in
1340: %the $m\to\infty$ and $\nn=2$ limit lead one to conjecture that this is
1341: %the equation for the $\nn=1$ curve.
1342: We leave the investigation of all these issues to future work. 
1343: 
1344: %The geometry of the
1345: %off-shell theory, which is a $n$-parameter deformation of the
1346: %Donagi-Witten curve, is certainly interesting
1347: %in itself and deserve further investigation. 
1348: 
1349: 
1350: 
1351: \begin{thebibliography}{99}
1352: 
1353: \bibitem{dv3}
1354: R.~Dijkgraaf and C.~Vafa,
1355: ``A perturbative window into non-perturbative physics,''
1356: [arXiv:hep-th/0208048].
1357: %%CITATION = HEP-TH 0208048;%%
1358: 
1359: \bibitem{int}
1360: F.~Cachazo, K.~A.~Intriligator and C.~Vafa,
1361: ``A large N duality via a geometric transition,''
1362: Nucl.\ Phys.\ B {\bf 603} (2001) 3
1363: [arXiv:hep-th/0103067]; 
1364: F.~Cachazo, B.~Fiol, K.~A.~Intriligator, S.~Katz and C.~Vafa,
1365: ``A geometric unification of dualities,''
1366: Nucl.\ Phys.\ B {\bf 628} (2002) 3
1367: [arXiv:hep-th/0110028].
1368: 
1369: \bibitem{cv}
1370: F.~Cachazo and C.~Vafa,
1371: ``N = 1 and N = 2 geometry from fluxes,''
1372: [arXiv:hep-th/0206017].
1373: %%CITATION = HEP-TH 0206017;%%
1374: 
1375: \bibitem{cdsw}
1376: F.~Cachazo, M.~R.~Douglas, N.~Seiberg and E.~Witten,
1377: ``Chiral rings and anomalies in supersymmetric gauge theory,''
1378: JHEP {\bf 0212} (2002) 071
1379: [arXiv:hep-th/0211170].
1380: %%CITATION = HEP-TH 0211170;%%
1381: 
1382: \bibitem{csw3}
1383: F.~Cachazo, N.~Seiberg and E.~Witten,
1384: ``Chiral Rings and Phases of Supersymmetric Gauge Theories,''
1385: [arXiv:hep-th/0303207].
1386: %%CITATION = HEP-TH 0303207;%%
1387: 
1388: \bibitem{dhks}
1389: N.~Dorey, T.~J.~Hollowood, S.~Prem Kumar and A.~Sinkovics,
1390: ``Exact superpotentials from matrix models,''
1391: JHEP {\bf 0211} (2002) 039
1392: [arXiv:hep-th/0209089];
1393: ``Massive vacua of N = 1* theory and S-duality from matrix models,''
1394: JHEP {\bf 0211} (2002) 040
1395: [arXiv:hep-th/0209099].
1396: %%CITATION = HEP-TH 0209089;%%
1397: 
1398: 
1399: \bibitem{dw}R.~Donagi and E.~Witten,
1400: ``Supersymmetric Yang-Mills Theory And Integrable Systems,''
1401: Nucl.\ Phys.\ B {\bf 460} (1996) 299
1402: [arXiv:hep-th/9510101].
1403: %%CITATION = HEP-TH 9510101;%%
1404: 
1405: \bibitem{nek}
1406: V.~A.~Kazakov, I.~K.~Kostov and N.~A.~Nekrasov,
1407: %``D-particles, matrix integrals and KP hierarchy,''
1408: Nucl.\ Phys.\ B {\bf 557} (1999) 413
1409: [arXiv:hep-th/9810035].
1410: %%CITATION = HEP-TH 9810035;%%
1411: 
1412: 
1413: 
1414: \bibitem{gw}
1415: S.~B.~Giddings and S.~A.~Wolpert,
1416: ``A Triangulation Of Moduli Space From Light Cone String Theory,''
1417: Commun.\ Math.\ Phys.\  {\bf 109} (1987) 177.
1418: %%CITATION = CMPHA,109,177;%%
1419: 
1420: 
1421: 
1422: \bibitem{dph}
1423: E.~D'Hoker and D.~H.~Phong,
1424: ``Calogero-Moser systems in SU(N) Seiberg-Witten theory,''
1425: Nucl.\ Phys.\ B {\bf 513} (1998) 405
1426: [arXiv:hep-th/9709053].
1427: %%CITATION = HEP-TH 9709053;%%
1428: 
1429: \bibitem{gopa}
1430: R.~Gopakumar,
1431: ``N = 1 theories and a geometric master field,''
1432: arXiv:hep-th/0211100.
1433: %%CITATION = HEP-TH 0211100;%%
1434: 
1435: \bibitem{csw2}
1436: F.~Cachazo, N.~Seiberg and E.~Witten,
1437: ``Phases of N = 1 supersymmetric gauge theories and matrices,''
1438: JHEP {\bf 0302} (2003) 042
1439: [arXiv:hep-th/0301006].
1440: %%CITATION = HEP-TH 0301006;%%
1441:   
1442: \bibitem{gh}
1443: Ph.~Griffiths and J.~Harris, {\it Principles of algebraic geometry},
1444: Wiley and sons, New York, 1979.
1445: 
1446: \bibitem{ad}
1447: O.~Aharony, N.~Dorey and S.~P.~Kumar,
1448: ``New modular invariance in the N = 1* theory, operator mixings and  supergravity singularities,''
1449: JHEP {\bf 0006} (2000) 026
1450: [arXiv:hep-th/0006008].
1451: %%CITATION = HEP-TH 0006008;%%
1452: 
1453: \bibitem{schnitzer}  S.~G. Naculich, H.~J. Schnitzer and  N. Wyllard,
1454: ``Cubic curves from matrix models and generalized Konishi anomalies", 
1455: [arXiv:hep-th/0303268].
1456:  
1457: \bibitem{ragazzi} R. Casero and E. Trincherini, 
1458: ``Quivers via anomaly chains", [arXiv:hep-th/0304123]. 
1459: 
1460: \bibitem{ber}
1461: D.~Berenstein,
1462: ``Solving matrix models using holomorphy,''
1463: JHEP {\bf 0306}, 019 (2003)
1464: [arXiv:hep-th/0303033].
1465: %%CITATION = HEP-TH 0303033;%%
1466: 
1467: \end{thebibliography}
1468: 
1469: \end{document}
1470: 
1471: 
1472: 
1473: 
1474: 
1475: 
1476: 
1477: 
1478: