hep-th0606261/p.tex
1: \documentclass[12pt]{article}
2: %  \usepackage[notref, notcite]{showkeys}
3: 
4: \topmargin      -0.3in  % distance to headers
5: \headheight      0.2in  % height of header box
6: \headsep         0.3in  % distance to top line
7: \textheight      8.9in  % height of text
8: \footskip        0.3in  % distance from bottom line
9: \oddsidemargin   0.0in  % Horizontal alignment
10: \evensidemargin  0.0in  % Horizontal alignment
11: \textwidth       6.5in  % Horizontal alignment
12: 
13: \renewcommand{\baselinestretch}{1.1}
14: \def\rem{$\clubsuit$} % remarks in the draft
15: 
16: \usepackage{amsmath,amssymb,latexsym,theorem,graphicx,amscd}
17: \usepackage[all]{xy}
18: \CompileMatrices
19: 
20: %%%%%% to help proofreading
21: \usepackage{color}
22: \definecolor{dblue}{rgb}{0,0,.7}
23: \definecolor{indigo}{RGB}{50,0,105}
24: 
25: \usepackage[hyperindex=true, colorlinks=true, pagecolor=dblue, bookmarksopen=false, urlcolor=dblue, linktocpage,linkcolor=dblue, citecolor=indigo]{hyperref}
26: \def\P{{\mathbb P}} \def\O{\mathcal{O}} \def\C{{\mathbb C}}\def\ff#1#2{{\textstyle\dfrac{#1}{#2}}}
27: %%%%%%%%%% Some commands  %%%%%%%%%%%%%%
28: \newcommand{\be}{\begin{equation}}
29: \newcommand{\ee}{\end{equation}}
30: \newcommand{\eq}[1]{Eq. (\ref{eq:#1})}
31: %
32: \newcommand{\Aphys}{{A^{(phys)}}}
33: \newcommand{\bAphys}{{{\bar A}^{(phys)}}}
34: \newcommand{\Amath}{{A^{(math)}}}
35: \newcommand{\bAmath}{{{\bar A}^{(math)}}}
36: %
37: \newcommand{\bbc}{{\mathbb C}}
38: \newcommand{\bbz}{{\mathbb Z}}
39: \newcommand{\bbf}{{\mathbb F}}
40: \newcommand{\bbr}{{\mathbb R}}
41: \newcommand{\rk}{{\text{ rk} }}
42: \newcommand{\Aut}{{\operatorname{Aut}}}
43: \newcommand{\vol}{{\text{vol} }}
44: \newcommand{\Tr}{{\text{ Tr}~ }}
45: \renewcommand{\Re}{{\text{ Re}~ }}
46: \newcommand{\End}{{\text{ End}~ }}
47: \newcommand{\Coh}{{\text{ Coh} }}
48: \newcommand{\effective}{{\text{effective} }}
49: \newcommand{\ch}{{\text{ch} }}
50: \newcommand{\sgn}{{\text{sgn}\,}}
51: \newcommand{\KC}{{\it KC}}
52: \newcommand{\tH}{{\tilde H}}
53: \newcommand\ba{\bar{a}}
54: \newcommand\bb{{\bar{b}}}
55: \newcommand\bc{\bar{c}}
56: \newcommand\bd{\bar{d}}
57: \newcommand\bz{\bar{z}}
58: \newcommand\bZ{\bar{Z}}
59: \newcommand\bW{\bar{W}}
60: \newcommand\bD{\bar{D}}
61: \newcommand\bA{\bar{A}}
62: \newcommand\bB{\bar{B}}
63: \newcommand\bR{\bar{R}}
64: \newcommand\bS{\bar{S}}
65: \newcommand\bT{\bar{T}}
66: \newcommand\bU{\bar{U}}
67: \newcommand\bPi{\bar{\Pi}}
68: \newcommand\bphi{\bar{\phi}}
69: \newcommand\balpha{{\bar{\alpha}}}
70: \newcommand\bbeta{{\bar{\beta}}}
71: \newcommand\bOmega{\bar{\Omega}}
72: \newcommand\bpartial{\bar{\partial}}
73: \newcommand\bj{{\bar{j}}}
74: \newcommand\bi{{\bar{i}}}
75: \newcommand\bk{{\bar{k}}}
76: \newcommand\bm{{\bar{m}}}
77: \newcommand\btheta{\bar{\theta}}
78: \newcommand\bpsi{\bar{\psi}}
79: \newcommand\bF{\bar{F}}
80: \newcommand\bs{{\bar{s}}}
81: \newcommand\bt{\bar{t}}
82: \newcommand\bv{\bar{v}}
83: \newcommand\bx{\bar{x}}
84: \newcommand\by{\bar{y}}
85: \newcommand\btau{\bar{\tau}}
86: \newcommand\bC{\bar{C}}
87: \newcommand\bX{\bar{X}}
88: \newcommand\bY{\bar{Y}}
89: \newcommand{\bra}[1]{{\langle}#1|}
90: \newcommand{\ket}[1]{|#1\rangle}
91: \newcommand{\bbra}[1]{{\langle\langle}#1|}
92: \newcommand{\kket}[1]{|#1\rangle\rangle}
93: \newcommand{\vev}[1]{\langle{#1}\rangle}
94: %
95: \newcommand{\nch}{{\text{ch}}}
96: \newcommand{\td}{{\text{td}\; }}
97: \newcommand{\pt}{{\text{pt} }}
98: %
99: \newcommand{\cp}[1]{{\mathbb P}^{#1}}
100: \newcommand{\op}[1]{\operatorname{#1}}
101: \newcommand{\ocp}[1][]{{\mathcal O}_{\cp{1}}{#1}}
102: \newcommand{\ocpn}[2][]{{\mathcal O}_{\cp{#1}}{#2}}
103: \newcommand{\oy}[1][]{{\mathcal O}_B{#1}}
104: \newcommand{\oz}[1][]{{\mathcal O}_{B^{'}}{#1}}
105: \newcommand{\ox}[1][]{{\mathcal O}_X{#1}}
106: \newcommand{\om}{{\mathcal O}_M}
107: \newcommand{\MW}{{\mathbb M}{\mathbb W}}
108: \newcommand{\mw}[1]{[#1]}
109: \newcommand{\Pic}{\op{Pic}}
110: \newcommand{\FM}{\boldsymbol{F}{\boldsymbol{M}}}
111: \newcommand{\fm}{\boldsymbol{f}\boldsymbol{m}}
112: \newcommand{\T}{\boldsymbol{T}}
113: \newcommand{\ct}{\boldsymbol{t}}
114: \newcommand{\D}{\boldsymbol{D}}
115: \newcommand{\num}[1]{{\#}\text{\bfseries #1}}
116: \newcommand{\rhom}{R^{\bullet}{\mathcal H}om}
117: \newcommand{\cH}{{\mathcal H}}
118: \newcommand{\cL}{{\mathcal L}}
119: \newcommand{\cV}{{\mathcal V}}
120: \newcommand{\cW}{{\mathcal W}}
121: \newcommand{\heck}{\boldsymbol{H}\boldsymbol{e}\boldsymbol{c}
122: \boldsymbol{k}\boldsymbol{e}}
123: \newcommand{\hup}[2]{\heck^{+}_{(#1)}(#2)}
124: \newcommand{\hdown}[2]{\heck^{-}_{(#1)}(#2)}
125: \newcommand{\homsh}[3]{{\mathcal H}om_{#1}(#2,#3)}
126: %\newcommand{\extsh}[4]{{\mathcal E}xt_{#2}^{#1}(#3,#4)}
127: \newcommand\Hom{{\rm Hom}}
128: \newcommand\Ext{{\rm Ext}}
129: \newcommand{\p}[1]{{\mathbb P}(#1)}
130: \newcommand{\bl}[2]{\operatorname{Bl}_{#1}#2}
131: \newcommand{\ses}[3]{\xymatrix{
132: 0 \ar[r] & {#1} \ar[r] & {#2} \ar[r] & {#3} \ar[r] & 0 \\ }}
133: 
134: %%%%%%%%%  Theorems and the like  %%%%%%%%%
135: \newtheorem{theo}{Theorem}[section]
136: \newtheorem{lem}[theo]{Lemma}
137: \newtheorem{cor}[theo]{Corollary}
138: \newtheorem{prop}[theo]{Proposition}
139: \newtheorem{defi}[theo]{Definition}
140: \newtheorem{conj}[theo]{Conjecture}
141: {\theorembodyfont{\rmfamily} \newtheorem{remark}[theo]{Remark}}
142: {\theorembodyfont{\rmfamily} \newtheorem{ex}[theo]{Example}}
143: 
144: % Put preprint number in top-right.
145: \def\pplogo{\vbox{\kern-\headheight\kern -29pt
146: \halign{##&##\hfil\cr&{\ppnumber}\cr\rule{0pt}{2.5ex}&\ppdate\cr}}}
147: \makeatletter
148: \def\ps@firstpage{\ps@empty \def\@oddhead{\hss\pplogo}%
149:   \let\@evenhead\@oddhead % in case an article starts on a left-hand page
150: }%      The only change in \maketitle is \thispagestyle{firstpage}  instead of \thispagestyle{plain}
151: \def\maketitle{\par
152:  \begingroup
153:  \def\thefootnote{\fnsymbol{footnote}}
154:  \def\@makefnmark{\hbox{$^{\@thefnmark}$\hss}}
155:  \if@twocolumn
156:  \twocolumn[\@maketitle]
157:  \else \newpage
158:  \global\@topnum\z@ \@maketitle \fi\thispagestyle{firstpage}\@thanks
159:  \endgroup
160:  \setcounter{footnote}{0}
161:  \let\maketitle\relax
162:  \let\@maketitle\relax
163:  \gdef\@thanks{}\gdef\@author{}\gdef\@title{}\let\thanks\relax}
164: \makeatother
165: 
166: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
167: 
168: \begin{document}
169: \setcounter{page}0
170: \def\ppnumber{\vbox{\baselineskip14pt
171: \hbox{hep-th/0606261}}}
172: \def\ppdate{RUNHETC-06-15} \date{}
173: 
174: \title{\bf \LARGE 
175: 	Numerical solution to the hermitian Yang-Mills equation on the Fermat quintic	\\[10mm]}
176: \author{\small
177:      {{\bf Michael R. Douglas}$\stackrel{\&}{\bf ,}$
178:      {\bf Robert L.~Karp}, 
179:      {\bf Sergio Lukic}
180:      {\bf and Ren\'e Reinbacher}} \\	%$^1$
181: [10mm]
182: \normalsize  Department of Physics, Rutgers University \\
183: \normalsize Piscataway, NJ 08854-8019  USA			\\	[10mm]
184: \normalsize $^\&$ I.H.E.S.,  Le Bois-Marie, Bures-sur-Yvette, 91440 France}
185: 
186: {\hfuzz=10cm\maketitle}
187: 
188: \vskip 1cm
189: 
190: \begin{abstract}
191: \normalsize
192: \noindent
193: We develop an iterative method for finding solutions
194: to the hermitian Yang-Mills equation on stable holomorphic vector
195: bundles, following ideas recently developed by Donaldson. 
196: As illustrations, we construct numerically the hermitian Einstein metrics
197: on the tangent bundle and a rank three vector bundle on
198: $\mathbb{P}^2$. In addition, we find a hermitian Yang-Mills connection
199: on a stable rank three vector bundle on the Fermat quintic.
200: \end{abstract}
201: 
202: \vfil\break
203: 
204: \tableofcontents
205: 
206: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
207: \section{Introduction}    \label{s:intro}
208: 
209: The modern study of compactification of higher dimensional theories
210: can be divided into two general branches.  The first makes use of
211: compactification manifolds with a good deal of symmetry, such as the
212: torus, sphere, squashed spheres and so on.  Such spaces have
213: explicitly known metrics, allowing explicit solutions of the equations
214: of motion, and explicit Kaluza-Klein reduction.  These solutions have many
215: applications, such as supergravity duals of large $N$ gauge theories;
216: however their high degree of symmetry tends to be a problem in trying to
217: obtain models with the level of complexity of the Standard Model or its
218: often-postulated extensions.
219: 
220: The second branch makes use of manifolds for which the relevant metrics
221: are known to exist by general theorems, but for which explicit
222: expressions are not known.  The most famous examples are the
223: Ricci-flat Kahler metrics conjectured to exist by Calabi and proven
224: to exist by Yau \cite{Yau}.
225: In 1985, it was proposed by Candelas {\it et al}  \cite{CHSW}
226: that compactification of the heterotic string on a Calabi-Yau manifold
227: could lead to quasi-realistic theories of particle physics, containing
228: grand unified extensions of the Standard Model and low energy
229: supersymmetry.
230: Since then, other metrics of this type, such as $G_2$
231: holonomy metrics, have been used in quasi-realistic compactifications;
232: see for example \cite{Acharya:Witten}.
233: 
234: Over the subsequent years, many tricks were developed to bypass the
235: difficulties posed by not knowing the compactification metric.  These
236: tricks began with the algebraic geometry behind the theorems of Yau
237: and Donaldson-Uhlenbeck-Yau, and gradually evolved into entire
238: branches of mathematical physics, such as topological string theory and
239: special geometry.  To drastically oversimplify, the
240: general picture is that certain ``protected'' quantities in the four
241: dimensional Effective Field Theory (EFT), such as the superpotential
242: in theories with four supercharges, and the prepotential in theories
243: with eight supercharges, can be computed using techniques combining
244: algebraic geometry with physical ideas.  Other quantities, such as the
245: Kahler potential in theories with four supercharges, cannot be
246: computed directly.  Since a good deal of important physics depends on
247: the Kahler potential -- precise values of particle masses, and the
248: existence and stability of supersymmetry breaking vacua, this
249: situation is not very satisfactory.
250: 
251: Almost all present knowledge about the Kahler potential in
252: the EFT comes from studying expansions around more computable limits.
253: The best known example is the case of $N=1$ compactifications which
254: contain $N=2$ subsectors, such as heterotic $(2,2)$ models, or type II
255: on Calabi-Yau orientifolds.  In these cases, there is a limit in which
256: part of the $N=1$ Kahler potential becomes equal to that of the
257: related $N=2$ theory, which is computable using special geometry.
258: Other examples include the solvable orbifold or Gepner model limits,
259: at which the entire Kahler potential is computable in principle
260: using CFT techniques.  However, it is not clear at present how
261: representative such results are of the general case.  Even a limited
262: ability to compute in the general case would allow studying this
263: question.
264: 
265: One completely general technique for addressing such problems is to
266: compute the Ricci-flat metrics and related quantities numerically.
267: Numerical methods are unavoidable in other areas of physics, beginning
268: with such seemingly elementary problems as computing the
269: spectrum of the helium atom or integrating Newton's equations for
270: the three body problem in
271: celestial mechanics; it would be surprising if string theory 
272: could avoid this.  To bring string theory closer to a possible
273: confrontation with real data, for example from collider physics, it
274: may be valuable to develop these missing parts of the theory of
275: compactification.
276: 
277: In this work, we make a start in this direction by showing two %three
278: things. First, we review how to use existing mathematical techniques
279: to numerically approximate metrics on Kahler manifolds, along lines
280: recently developed by Donaldson \cite{Donaldson:numeric}.  Second, we
281: extend these mathematical techniques to hermitian Yang-Mills
282: connections.  
283: It will be clear that these techniques could
284: be pushed to compute higher order terms, metrics on moduli spaces, and
285: the like.  A subsequent paper will explain the numerical methods in
286: more detail and do some simple computations of terms in the
287: EFT for compactification on a
288: quintic Calabi-Yau 3-manifold.
289: 
290: Our direct inspirations are Donaldson's work \cite{Donaldson:numeric}
291: on numerical approximation of metrics, and of Wang
292: \cite{Wang:metricsbundles} developing the corresponding mathematics
293: for vector bundles.  We can also cite Headrick and Wiseman
294: \cite{Headrick:Wiseman}, who made a pioneering numerical study of the
295: K3 metric using position-space methods.  Finally, the first author is 
296: particularly indebted to Bernie Shiffman and Steve Zelditch for teaching him
297: the basics of asymptotic analysis on holomorphic line bundles, and for
298: advice in the early stages of this project, in particular for
299: pointing out Wang's work.
300: 
301: Let us briefly explain the problem and survey some of the approaches
302: one might take towards it, before beginning the detailed development
303: in section 2.  Following \cite{CHSW}, the derivation of the matter
304: Lagrangian in a heterotic compactification on a Calabi-Yau $X$ 
305: carrying a bundle $V$ involves the following steps:
306: \begin{enumerate}
307: \item Find the Ricci-flat metric $g_{ij}$ (with specified moduli) on $X$.
308: \item Find the hermitian Yang-Mills connection $A_i$ on $V$.
309: \item Find the zero modes $\psi^\alpha$
310: of the Dirac operator.  As is standard, on a Kahler manifold this
311: amounts to finding harmonic differential forms $\psi$ valued in $V$,
312: {\it i.e.} solutions of 
313: $0=(\bpartial+\bA)\psi=(\bpartial+\bA)^*\psi$,
314: where $*$ denotes the adjoint operator.  
315: \item Find an orthonormal basis of forms $\psi$.
316: \item Compute the integrals over $X$ of wedge products of these forms
317: to get the superpotential.
318: \end{enumerate}
319: The key step for us is (4).  Existing methods for computing the
320: superpotential, such as \cite{GSW2,Candelas:1987rx,Donagi:2006yf}, accomplish step (5) without needing the
321: results of (1) and (2), by using unnormalized zero modes.  This leads
322: to a superpotential defined in terms of fields whose kinetic term is
323: obtained from ``some'' unknown Kahler potential.  To do better, we must
324: either derive normalized zero modes in (4) for use in (5), or else take
325: the zero modes used in (5) and compute their normalizations using the
326: explicit metric from (1).
327: 
328: There seems to be no way of doing this without some knowledge of the
329: Ricci-flat metric and thus the first step is to choose some
330: approximation scheme for this metric.  One's first thought might be to
331: follow standard practice in numerical relativity, as done in
332: \cite{Headrick:Wiseman}, and introduce a six dimensional lattice which
333: is a discrete approximation to the manifold $X$; in other words a
334: position space approach.  Taking the Kahler potential $K$ as the basic
335: dynamical variable, Einstein's equations reduce to the complex
336: Monge-Ampere equation \be\label{eq:mongeampere} \det(\partial\bpartial
337: K) = \Omega\wedge\bOmega \ee which can be solved by relaxation
338: methods.  One would then need to find similar lattice approximations
339: for the connection on $V$ and the zero modes.
340:  
341: An alternative approach, introduced by Donaldson
342: \cite{Donaldson:numeric}, is to use the natural embedding of $X$ into
343: $\P^{N-1}$ provided by the $N$ sections of an ample line bundle $L^k$ (we will
344: explain this in detail below).  We then take as a candidate
345: approximating metric on $X$ the pull-back of a Fubini-Study metric on
346: $\P^{N-1}$.  Such a metric is defined by an $N\times N$ hermitian
347: matrix. By suitably choosing this matrix we can try to make the
348: associated Fubini-Study metric restrict to $X$ in such a way that it
349: gives a good approximation to the Ricci-flat metric on $X$.
350: 
351: A major advantage of this approach is that it avoids the complications and
352: arbitrariness involved in choosing an explicit discretization of $X$;
353: rather the entire approximation scheme follows from a single parameter
354: $k$, the scale of the first Chern class of $L$.  Subsequent
355: mathematical development reveals more structure which can be used to our
356: advantage.  For example, a very natural approximation to the Ricci-flat
357: metric, which becomes exact as $k\rightarrow\infty$, is the so-called
358: ``balanced'' metric.  In a sense, to be described below, this is
359: the metric for which the embedding of $X$ into $\P^{N-1}$ has its center
360: of mass at the ``origin''.  It also satisfies a simple fixed point condition
361: which can be used for relaxation, solving step (1).
362: 
363: Another advantage, which is key for the present application, is that
364: Donaldson's method can be naturally extended to study holomorphic
365: vector bundles on $X$.  There is a standard relation between
366: holomorphic connections and hermitian metrics, which we review in
367: section 2, in which step (2) of the above prescription is turned into
368: the problem of finding a hermitian-Einstein metric on a vector
369: bundle. For illustrative purposes we will explicitly study
370: hermitian-Einstein metrics on two spaces: complex projective space
371: $\P^n$ and the Fermat quintic threefold.
372: 
373: The organization of the paper is as follows. In Section~2 we provide
374: an overview of the geometric background needed for our construction,
375: in particular we will describe Donaldson's approach for getting
376: metrics of constant scalar curvature. In Section 3 we explain a
377: numerical approximation to the hermitian Einstein metric on a
378: holomorphic vector bundle by a simple adaptation of Donaldson's
379: scheme, building on mathematical work of Wang.  
380: In section 4 we focus
381: on several explicit examples. Here we describe some of our numerical
382: methods and results in detail. By design we are also able to test our
383: approximation scheme for $T\P^2(k)$, where $T\P^2$ is the holomorphic
384: tangent bundle of $\P^2$, since in this case one has an analytic
385: solution.
386: 
387: \section{Metrics of constant scalar curvature}
388: 
389: We follow the plan outlined in the introduction, beginning with step (1). Let
390: $X$ be an $n$-dimensional complex Kahler ``compactification manifold.''
391: Since we are not assuming it is a valid string theory background, we 
392: can generalize the discussion to arbitrary $n$ and first Chern class $c_1(X)$.
393: 
394: The basic example we have in mind is the complex projective space
395: $\cp{n}$, parameterized by the standard homogenous coordinates
396: $\{Z_i\}_{i=0}^n$, up to the identification $\{Z_i\} \cong \{\lambda
397: Z_i\}$ for $\lambda\in\bbc^*$.  The Kahler potential
398: \be\label{eq:unitFS} K_{FS} = k~ \log \left(\sum_{i=0}^n |Z_i|^2
399: \right) \ee defines the Fubini-Study metric on $\P^n$, with $SU(n+1)$
400: symmetry $Z_i\rightarrow g_i^j Z_j$, and $g\in U(n+1)$.  The parameter
401: $k$ controls the Kahler class $\omega=\partial\bpartial K_{FS}$.
402: 
403: Our other general example 
404: is the hypersurface defined by the vanishing of a degree $d$ polynomial 
405: in $\P^n$:
406: \be\label{eq:hyperQ}
407: f(Z) = \sum_{i_1\cdots i_d} c^{i_1\cdots i_d} Z_{i_1}\ldots Z_{i_d}.
408: \ee
409: For $n=4$ and $d=5$ we get a quintic threefold $Q$.  Its complex
410: structure is determined by the $126$ parameters $c^{i_1\cdots i_5}$,
411: modulo the action of $GL(5,\C)$ on the $Z_i$'s,
412: which leaves $101$ parameters.  The generic member of 
413: this family is smooth, and has $b^{1,1}=1$. Therefore the
414: Kahler class is determined by a single real number.  A simple one
415: parameter family of Kahler metrics on $Q$ is obtained by pulling back the
416: Fubini-Study metric on $\cp{4}$, or equivalently interpreting \eq{unitFS}
417: as a Kahler potential on $Q$.  Of course this will not be Ricci-flat.
418: 
419: \subsection{Approximating Ricci-flat metrics by projective embedding}
420: 
421: We now want to find a larger space of Kahler metrics in which
422: to find a better approximation to the Ricci flat metric.
423: One simple generalization of \eq{unitFS} can be obtained by choosing an
424: $(n+1)\times (n+1)$ hermitian matrix $h^{i\bj}$, and writing
425: \be\label{eq:hFS}
426: K_{h} = k \log \left(\sum_{i,\bj=0}^n h^{i\bj} Z_i \bZ_j \right).
427: \ee
428: Of course, by making a linear redefinition of coordinates, we could
429: turn this back into \eq{unitFS}, but doing so would modify
430: \eq{hyperQ}.  Rather, by fixing \eq{hyperQ}, this way we get an $(n+1)^2$-parameter family of Kahler potentials.
431: 
432: Another way to think about this definition is to make the linear
433: redefinition taking $h$ to the identity.  In this case, the parameters
434: we are varying to control the metric are the extra $25$ parameters in
435: \eq{hyperQ} determining a specific embedding of $Q$ into $\P^4$.
436: While all of the embeddings are equivalent under a $GL(5,\C)$ action, once
437: we use the metric, we break this to $U(5)$; thus the set of metrics
438: we can obtain this way is parameterized by a $GL(5,\C)/U(5)$ homogeneous space.
439: 
440: A simple generalization to get more parameters could be motivated by
441: noticing that \eq{unitFS} is also equal to
442: \begin{eqnarray}
443: K_{FS,k} &=& \log \left(\sum_{i=0}^n |Z_i|^2 \right)^k \\
444:  &=& \log \left(\sum_{i_1,\cdots,i_k=0}^n
445:  Z_{i_1}\cdots Z_{i_k}\bZ_{i_1}\cdots\bZ_{i_k} \right) 
446: \end{eqnarray}
447: and generalizing this to
448: \be\label{eq:Kh}
449: K_{h,k} = \log \left(\sum_{i_1,\cdots,i_k,\bj_1,\cdots,\bj_k=0}^n
450:  h^{i_1\cdots i_k\bj_1\cdots \bj_k}
451:  Z_{i_1}\cdots Z_{i_k}\bZ_{\bj_1}\cdots\bZ_{\bj_k}\right) ,
452: \ee
453: which can again be interpreted as a Kahler potential on $Q$.
454: In simple terms, we are using higher degree polynomials as basis functions.
455: Now we have an $(n+1)^{2k}$-parameter family of metrics, and by taking
456: $k$ large we can imagine finding an arbitrarily good approximating metric
457: within this class.  
458: 
459: One way to find the best approximation to the Ricci-flat metric on $Q$
460: would be to write \eq{mongeampere} directly in these variables.  Note that
461: the holomorphic $(n,0)$-form $\Omega$ is known explicitly. For example, in the  coordinate patch where $Z_0\neq 0$ we can choose the local coordinates $w_i=Z_i/Z_0$, in terms of which
462: $$
463: \Omega = \frac{dw_1\, dw_2\, dw_3}{\partial f/\partial w_4} ,
464: $$
465: and thus one can write the volume form for the Ricci flat metric
466: explicitly,
467: \be\label{eq:cyvol}
468: {d\vol}_X  = \Omega\wedge\bOmega
469: \ee
470: without solving any equations.  One might then substitute \eq{Kh} into
471: \eq{mongeampere}, evaluate this at a set of points $p_i$, and solve
472: the resulting system of nonlinear equations.  These are rather
473: complicated, however, and furthermore we have introduced arbitrariness
474: in the choice of the $p_i$.  Now this arbitrariness can have
475: its uses, for example we might use it to place more points in
476: regions of large curvature.  On the other hand, it means that the
477: results will not have simple mathematical or physical interpretations,
478: except in the limit in which the number of points is so large that we
479: can ignore the discretization.\footnote{ Or unless we can come up with
480: a construction in which some sort of physical objects at the points
481: $p_i$ enforce the equations.}  Before investing a lot of effort into
482: their study, we should try to improve on this point.
483: 
484: \subsection{Balanced metrics}
485: 
486: There is a pretty construction that goes back to \cite{Tian:metrics}
487: which provides a more natural approximating metric, and a numerical
488: scheme which is guaranteed to converge to it.
489: 
490: First, we can systematize the construction which led to \eq{Kh}, by
491: noting that the basis functions are products of degree
492: $k$ holomorphic times  degree
493: $k$ antiholomorphic monomials.  Let the number
494: of independent holomorphic degree $k$ monomials be $N+1$; this is
495: the binomial coefficient $\binom{n+k} {k}$ 
496: for $\cp{n}$, and we will give it for $Q$ later.
497: 
498: Let us phrase this construction in a way 
499: which can be used for an arbitrary manifold $X$.  
500: We choose a holomorphic line bundle $\cL$ over $X$, 
501: with $N$ global sections.  Denote a complete basis of these as
502: $s_\alpha$, where $1\le\alpha\le N$, and consider the map
503: $$
504: i_k\colon X\longrightarrow \cp{N-1}
505:  \qquad i_k(Z_0,\ldots,Z_n) = (s_1(Z),s_2(Z),\ldots,s_N(Z)).
506: $$
507: The geometric picture is that each point in our original manifold $X$
508: (parameterized by the $Z_i$)
509: corresponds to a point in $\bbc^{N}$ parameterized by the sections
510: $s_\alpha$.  Since choosing a different frame for $\cL$ would produce
511: an overall rescaling $s_\alpha\rightarrow \lambda s_\alpha$, the overall
512: scale is undetermined. Granting that 
513: $s_1(Z),s_2(Z),\ldots,s_N(Z)$ 
514: do not vanish simultaneously, this gives us a map to $\P^{N-1}$. 
515: 
516: The simplest example is to embed $\cp{1}$ using $\cL=\O_{\P^1}(k)$ 
517: into $\cp{k}$. In this case the map is
518: $$
519: i_k(Z_0,Z_1) =
520:  (Z_0^k,~ Z_0^{k-1}Z_1,~Z_0^{k-2}Z_1^2,~\ldots,~ Z_0 Z_1^{k-1},~Z_1^k).
521: $$
522: 
523: In general we want this map to be an
524: embedding, {\it i.e.} that distinct points map to distinct points with
525: non-vanishing Jacobian.  In general, we can appeal to the Kodaira embedding
526: theorem, which asserts that for positive $\cL$
527: this will be true for all $\cL^k$ for some
528: $k\ge k_0$.  For non-singular quintics, this is true for $\om(k)$ for
529: all $k\ge 1$.  As a point of language, the pair of a
530: manifold $X$ with a positive line bundle $\cL$ is referred to as 
531: a {\em polarized manifold} $(X,\cL)$; the condition that this construction
532: provides an embedding for some $k$ is that $\cL$ is {\it ample}.
533: 
534: Now, we consider our family \eq{Kh} of candidate K\"ahler
535: potentials, and rewrite them as
536: $$
537: K_{h} = \log \left(\sum_{\alpha,\bbeta} h^{\alpha\bbeta}
538:  s_\alpha \bs_\bbeta \right)
539: $$
540: or simply
541: \be\label{eq:Ks}
542: K_{h} \equiv \log ||s||^2_h
543: \ee
544: for short, where $s_\alpha$ plays the role of a degree $k$ monomial.
545: We now have an $N^2$-parameter family of K\"ahler potentials, and will
546: seek a good approximating metric in this family.
547: Just as before, this amounts to using the pull-back of a Fubini-Study
548: metric from $\cp{N-1}$ as our trial metric.
549: 
550: \iffalse
551: To what extent is the definition of this family of metrics free of
552: arbitrary choices?  While choosing a specific basis $s_\alpha$ gives a
553: further arbitrariness in the embedding, this arbitrariness is exactly
554: parameterized by the action $g\in SL(N)$ as $s_\alpha \rightarrow
555: g_\alpha^\beta s_\beta$.  Any basis of holomorphic sections can be
556: taken to any other by this action, and in this sense the embedding is
557: entirely determined by the choice of $\cL$.  However once we use the
558: metric \eq{Ks} on $\P^{N}$, we again break this symmetry to
559: $U(N+1)$.  Thus we can again think of the metric as parameterized by
560: the matrix $h_{\alpha\bbeta}$ up to $U(N+1)$ equivalence, or else as
561: parameterized by a specific choice of embedding.  In this sense, given
562: a definition of ``best approximating metric,'' the embedding which leads
563: it is uniquely defined.
564: 
565: The remaining choice of basis and breaking of $U(N+1)$ symmetry is an
566: arbitrariness in any practical implementation of this scheme; whether
567: this is better or worse than choosing (say) a set of points depends on
568: precisely what one is doing.  But since the theory of $U(N+1)$
569: invariants is very familiar, one can easily derive basis-independent
570: quantities which characterize the metric at this point.
571: \fi
572: 
573: Mathematically, the simplest interpretation of \eq{Ks} is that it
574: defines a hermitian metric on the line bundle
575: $\cL = \om(k)$.  This is a sesquilinear map 
576: from $\bar\cL\otimes \cL$ to smooth functions $C^\infty(X)$, here defined by
577: $$
578: (s,s') = e^{-K_{h}}\cdot\bar s\cdot s' =
579:  \frac{\bar s\cdot s'}{\sum_{\alpha,\bbeta} h^{\alpha\bbeta}
580:  s_\alpha \bs_\bbeta} .
581: $$
582: The point is that a change of frame, which acts on our explicit sections
583: as $s_\alpha\rightarrow \lambda s_\alpha$,
584: cancels out of this expression.\footnote{A possibly more familiar
585: physics use of this is in $N=1$ supergravity: taking
586: $K\rightarrow -K$ and  $s\rightarrow W$, one gets the standard 
587: expression for the gravitino mass $e^K|W|^2$.  In an example such as
588: the flux superpotential, in which $W$ is a sum of various terms $s_\alpha$
589: with constant coefficients, \eq{Ks} also applies to give $K$.}
590: 
591: This metric allow us to define an inner product between the global sections:
592: \be\label{eq:defH1}
593: H_{\alpha\bbeta} = \vev{s_\beta|s_\alpha} = i\int_X 
594: \frac{s_{\alpha}\bs_{\bbeta}}{||s||^2_h} \,d{\vol_X}.
595: \end{equation} 
596: This is the ``physical'' inner product in a sense we will explain further
597: below.
598: Note that it depends on $h$ in a nonlinear way, since $h$ appears
599: in the denominator.
600: 
601: Here $d\vol_X$ is a volume form on $X$, which has to be chosen.
602: If $X$ is Calabi-Yau, it is simplest to use \eq{cyvol} to define $d\vol_X$.
603: If $X$ is not Calabi-Yau, the standard choice of $d\vol_X$ is
604: to take $d\vol_\omega=\omega^n/n!$, where $\omega$ is the 
605: Kahler metric derived from \eq{Ks}.  This depends on $h$ as well, so
606: the expression is even more non-linear in $h$.
607: 
608: Thus, given $h$ and a basis of global sections $s_\alpha$, we could
609: compute the matrix of inner products \eq{defH1}.
610: Once we have it, we 
611: could make a linear redefinition, say $\tilde s = H^{-1/2} s$, and
612: go to a basis of orthonormal sections where
613: \be\label{eq:onH}
614: H_{\alpha\bbeta} = {\delta}_{\alpha\bbeta} .
615: \ee
616: 
617: On the other hand, \eq{Ks} also implicitly
618: defines a notion of orthonormal basis locally in the bundle, in which
619: \be
620: h^{\alpha\bbeta} = {\delta}^{\alpha\bbeta} .
621: \ee
622: This is {\it a priori} different from \eq{onH};
623: indeed we can freely postulate it when we write \eq{Ks}.
624: However, if the two notions agree,
625: $$
626: H_{\alpha\bbeta} = (h^{-1})_{\alpha\bbeta} ,
627: $$
628: then we can go to a basis of sections in which
629: \be\label{eq:balanced}
630: H_{\alpha\bbeta} = h^{\alpha\bbeta}  = {\delta}_{\alpha\bbeta} .
631: \ee
632: In this case, 
633: the embedding of $X$ in $\mathbb{P}^{N-1}$ using these sections
634: is called {\it balanced}.
635: More generally, we call
636: a polarized manifold $(X,\cL^k)$ balanced 
637: if such an embedding exists.
638: 
639: An equivalent definition of 
640: the balanced embedding is arrived at if we consider the
641: function on $X$ defined as
642: \begin{equation}\label{eq:bk}
643:  \rho(\omega)(x)=
644:  \sum_{\alpha,\bbeta} (H^{-1})^{\alpha\bbeta} (s_\alpha(x),\bs_\bbeta(x))
645: \end{equation}
646: or equivalently
647: $$
648: \rho(\omega)(x)= \sum_{\alpha} ||s_\alpha(x)||^2
649: $$
650: where the second sum is taken over an orthonormal basis in which
651: $H={\delta_{\alpha\beta}}$.  $X$ is balanced precisely when
652: $\rho(\omega)(x)$ is the constant function.
653: 
654: Many theorems have been proven about balanced manifolds.  Let us first
655: recall the following theorem of Donaldson (Theorem 1 in
656: \cite{Donaldson1}):
657: \begin{theo}\label{thm:unique} 
658: Suppose the automorphism group $\Aut(X,\cL)$ is discrete. If
659: $(X,\cL^k)$ is balanced, then the choice of basis in $H^0(X,\cL^k)$
660: such that $i_k(\cL)$ is balanced is unique up to the action of
661: $U(N)\times \mathbb{R}^*$.
662: \end{theo}
663: The condition on $\Aut(X,\cL)$, i.e., there are no continuous
664: symmetries, is true for the quintic $Q$.  This theorem then tells us
665: that, if a metric $h$ exists which gives a balanced embedding, it is
666: unique up to scale.
667: 
668: Given a balanced embedding, one defines the {\it balanced metric} on
669: $X$ as the pullback of the Fubini-Study metric (\ref{eq:Ks}):
670: \begin{equation}
671: \label{ }
672: \omega_k=\frac{2\pi}{k}i_k^*(\omega_{FS}),
673: \end{equation}
674: The cohomology class of the Kahler form $[\omega_k]=2\pi c_1(\cL)\in
675: H^2(X)$ is independent of $k$. Using these definitions Donaldson
676: proves that (Theorem 2 in \cite{Donaldson1}):
677: \begin{theo}\label{theo2}
678: Suppose $\Aut(X,\cL)$ is discrete and $(X,\cL^k)$ is balanced for
679: sufficiently large $k$. If the metrics $\omega_k$ converge
680: in the $C^\infty$ norm to some limit $\omega_\infty$ as $k\to \infty$, then
681: $\omega_\infty$ is a Kahler metric in the class
682: $2\pi c_1(\cL)$ with constant scalar curvature.
683: \end{theo}
684: The constant value of the scalar curvature is determined by $c_1(X)$,
685: and in particular for $c_1(X)=0$ the scalar curvature is zero.  Thus,
686: the balanced metrics $\omega_k$, in the large $k$ limit, converge to
687: the Ricci flat metric.
688: 
689: Therefore, if we can find the unique balanced metric for a given
690: $\cL$, it is a good candidate for approximating the Ricci flat metric
691: on $X$.  One may ask where the complex structure and Kahler moduli on
692: which this Ricci flat metric depends, are put in.  The complex
693: structure enters implicitly through the basis for holomorphic sections
694: $s_\alpha$, as we will see in examples below.  As for the Kahler
695: class, recall that this is determined, up to scale, to be $2\pi
696: c_1(\cL)$.  Of course, the Ricci flatness condition is scale
697: invariant, so the overall scale is irrelevant; however the point of
698: this is that if $b^{1,1}>1$, then by appropriately choosing $\cL$ we
699: choose a particular ray in the Kahler cone.  This will not be relevant
700: for our examples here but shows that in principle any Ricci-flat
701: Kahler metric could be approximated in this way.
702: 
703: \subsection{Finding the balanced metric}
704: 
705: In \cite{Donaldson2,Donaldson:numeric} Donaldson proposes a method to
706: determine the hermitian metric $h$ in \eq{Ks}, which will lead to a
707: balanced metric.  He defines the ``T operator'', which given a metric
708: $h$ computes the matrix $H$: 
709: \be\label{eq:defH} H_{\alpha\bbeta} =
710: T(h)_{\alpha\bbeta} \equiv \frac{N}{\vol(X)}\int_X \frac{s_\alpha
711: \bs_\bbeta}{||s||^2_h}\, {d\vol_X}
712: \end{equation} 
713: Now, suppose we find a fixed point of this operator,
714: $$
715: T(h) = h .
716: $$
717: Then, by a $GL(N)$ change of basis $s\rightarrow h^{-1/2}s$,
718: we can bring $h$ to the unit matrix,
719: which will produce the balanced embedding.
720: 
721: The simplest way to find a fixed point of an operator is to iterate
722: it.  If the operator is contracting, this is guaranteed to work.  In
723: our case we have the following theorem \cite{Donaldson1,Sano}:
724: \begin{theo}
725: Suppose that $\Aut(X,L)$ is discrete. If a balanced embedding
726: exists then, for any initial $G_0$ hermitian metric, the sequence $T^r(G_0)$
727: converges to the balanced metric $G$ as $r\to \infty$.
728: \end{theo}
729: Thus the $T$ operator can be used to find
730: approximate Ricci-flat metrics on Calabi-Yau manifolds, and more
731: generally approximate
732: constant scalar curvature Kahler metrics.  
733: In \cite{Donaldson:numeric} Donaldson
734: studies numerically explicit $\cp{1}$ and $K3$ examples.  We will discuss some
735: additional examples below.
736: 
737: \subsection{Balanced metrics and constant scalar curvature}\label{s:prs}
738: 
739: In this subsection we outline the reason why the limit of a family of
740: balanced metrics has constant scalar curvature. This is the content of
741: Theorem~\ref{theo2}. This will be very useful later on, when we
742: generalize the T-operator to vector bundles.
743: 
744: Note that the  function $\rho(\omega)$ is independent of the choice of orthonormal basis, and remains
745: unchanged if we replace $h$ by a constant scalar multiple. Therefore,
746: it is an invariant of the Kahler form. As discussed before, the balanced condition for
747: $(X,L^k)$ is equivalent to the existence of a metric $\omega_k$ such
748: that $\rho(\omega_k)$ is a constant function on $X$. The asymptotic behavior of
749: the ``density of states" function $\rho(\omega_k)$ as $k \to \infty$
750: for fixed $\omega$ has been studied in \cite{Tian:metrics,Catlin,Zelditch:Szego,LuZhiqin}. Note that for any metric
751: \begin{equation}
752: \label{eq-rho}
753: \int_X \rho_k(\omega) = N=\dim H^0(X,L^k)=a_0k^n +a_1k^{n-1}+\cdots,
754: \end{equation}
755: where the  coefficients $a_i$ can be determined using the Riemann-Roch formula.  Note that $a_0$ is just the volume of $X$ and 
756: $$
757: a_1=\frac{1}{2\pi}\int_X S(\omega),
758: $$
759: where $S(\omega)$ is the scalar curvature of $\omega$. We will use the following result (Prop. 6 in \cite{Donaldson1}):
760: \begin{prop}
761: \begin{enumerate}
762:   \item $\rho(\omega)$ has an asymptotic expansion as $k \to \infty$
763:   $$
764:    \rho_k(\omega) \sim A_0(\omega)k^n +A_1(\omega)k^{n-1}+\cdots
765:   $$
766:   where $A_i(\omega)$ are smooth functions on $X$ defined locally by $\omega$. In particular, 
767:   $$
768:   A_0(\omega)=1,\qquad A_1(\omega)=\frac{1}{2\pi}S(\omega).
769:   $$
770:   \item The expansion holds uniformly in the $C^\infty$ norm; in that for any $r,N>0$
771:   $$
772:   \left\| \rho_k(\omega)-\sum_{i=0}^N A_i(\omega)k^{n-i} \right\|_{C^r(X)}\leqslant K_{r,N,\omega}k^{n-N-1}
773:   $$
774:   for some constants $K_{r,N,\omega}$. 
775: \end{enumerate}
776: \end{prop}
777: 
778: Now assume that we are given balanced metrics $\omega_k$ converging to $\omega_\infty$. Then by the previous proposition
779: $$
780: \left\| \rho_k(\omega_k)-k^{n}-\frac{1}{2\pi}S(\omega_k)k^{n-1} \right\|_{C^0(X)}\leqslant ck^{n-2}
781: $$
782: for some constant $c$. Since $\omega_k$ is balanced $\rho_k(\omega_k)$ is constant: $\rho_k(\omega_k)= \dim H^0(X,L^k)/V$, and we can use (\ref{eq-rho}) to find that
783: $$
784: \left\|  \ff 1 V (V k^n + a_1k^{n-1}+\cdots )-k^{n}-\frac{1}{2\pi}S(\omega_k)k^{n-1} \right\|_{C^0(X)}\leqslant ck^{n-2},
785: $$
786: or equivalently
787: $$
788: \left\|  \ff {2\pi} V  a_1-S(\omega_k) \right\|_{C^0(X)}= O(k^{-1})
789: $$
790: Hence $S(\omega_\infty)=S_0=constant$, where $S_0=\frac{1}{V}\int_X S(\omega)$ is the mean curvature.
791: 
792: \section{The hermitian Yang-Mills equations}
793: 
794: We are now ready to generalize the T-operator, which provided an
795: approximation scheme for the constant curvature metric, to a
796: ``generalized T-operator'' which can be used to find a solution of the
797: Yang-Mills equations on a Calabi-Yau manifold $X$.
798: 
799: We briefly recall the argument that a solution of the Yang-Mills equations
800: which preserves ${\cal N}=1$ supersymmetry, must be hermitian Yang-Mills.
801: First, the supersymmetry variation of the gaugino has to vanish,
802: $$
803: \Gamma^{\mu\nu} F^a_{\mu\nu}\epsilon =0,
804: $$
805: where $F^a_{\mu\nu}$ is the Yang-Mills field strength, and
806: $\epsilon$ is the covariantly constant spinor.
807: 
808: Going to complex coordinates $(i,\bi)$ and 
809: rewriting of the Clifford algebra as
810: $$
811: \Gamma_i \rightarrow dz^i; \qquad
812:  \Gamma_{\bi} \rightarrow \omega_{\bi j} ~ {\partial}^j ,
813: $$
814: this is equivalent to
815: $$
816: F_{ij} = F_{\bi\bj} = 0; \qquad
817: \omega^{i\bj} F_{i\bj} = 0 .
818: $$
819: This is the particular case of the hermitian Yang-Mills equations
820: with $\Tr F = 0$.  The general case replaces the last equation with
821: $$
822: \omega^{i\bj} F_{i\bj} = c\cdot{\bf 1} 
823: $$
824: for a constant $c$, determined by the first Chern class.  For convenience
825: we abbreviate this equation below as
826: $$
827: \bigwedge F = c\cdot{\bf 1} .
828: $$
829: 
830: Next we review the relation between solutions of these equations,
831: and holomorphic bundles carrying hermitian-Einstein metrics.
832: In physics, one defines Yang-Mills theory in terms of a connection
833: on a vector bundle with a fixed metric.
834: First, a connection on a vector bundle can be described in terms
835: of a connection one-form by choosing a frame for the
836: bundle, say $e_a(x)$, and defining the covariant derivative as
837: $$
838: D(v^a e_a) 
839:  = (dv^a) e_a + v^a A^b_a e_b .
840: $$
841: In physics, one usually takes the frame to be orthonormal, 
842: $(e_a,e_b) = \delta_{ab} $, and thus
843: \be\label{eq:physmet}
844: (u,v) = (u^a e_a,v^b e_b) = (u^a)^* v^a ,
845: \ee
846: where $*$ is complex conjugation.
847: 
848: The condition that the connection be compatible with the metric,
849: \be\label{eq:metcomp}
850: d(u,v) = (Du,v) + (u,Dv),
851: \ee
852: reduces to requiring the connection one-form to be anti-hermitian,
853: \be\label{eq:antiherm}
854: \Aphys_i = -\Aphys_i^\dag.
855: \ee
856: 
857: In mathematics, one often considers a more general frame, for
858: which the metric is a hermitian matrix,
859: \be\label{eq:genherm}
860: (e_a,e_b) = G_{\ba b}, \qquad G = G^\dag .
861: \ee
862: Decomposing the positive definite hermitian matrix $G$ as
863: \be\label{eq:Ghh}
864: G = h^\dag h,  
865: \ee
866: we see that the math and physics conventions differ
867: by a complex gauge transformation:
868: $ 
869: u = h\,s.
870: $ 
871: This complex gauge transformation  leads to a different form
872: for the connection, according to the standard
873: relation
874: \be\label{eq:compgy}
875: \partial_i + \Amath_i = h (\partial_i + \Aphys_i) h^{-1} .
876: \ee
877: 
878: Now, equations of the form
879: $$
880: F_{\bi\bj} = 0\  \forall\ \bi,\bj
881: $$
882: will be integrability conditions for the covariant derivatives.
883: In particular, this equation has the general solution
884: $$
885: \partial_{\bi} + \Aphys_{\bi} = g^{-1} \bar\partial_{\bi} g ,
886: $$
887: in other words the $\bD$ covariant derivatives are obtained from
888: the derivative $\bar\partial$ by a complex gauge transformation.
889: 
890: Thus, we can use \eq{compgy} to bring the connection to the gauge
891: $\bAmath=0$, at the cost of losing the simple metric \eq{physmet} and
892: \eq{antiherm}.  Actually, the covariant
893: derivative is still compatible with the metric as in \eq{metcomp},
894: we just have a non-trivial fiber metric $G$.
895: The metric compatibility condition becomes
896: $$
897: 0 = \partial(u,v) = ({\bpartial u},v)+(u,D v)
898: $$
899: so
900: $$
901: \partial G_{\ba b} = G_{\ba c} \Amath^c_b
902: $$
903: or equivalently
904: $$
905: \Amath = G^{-1}\partial G .
906: $$
907: 
908: Conversely, if we are given a metric $G$, then we can use
909: the inverse complex gauge transformation to bring the connection
910: back to the unitary form.  
911: This leads to the formula
912: $$
913: \bAphys_{\bi} = h (\bpartial_{\bi} h^{-1}).
914: $$
915: Using \eq{antiherm}, we can get the entire connection, so the metric
916: contains the same information as a connection satisfying 
917: $F^{(0,2)}=F^{(2,0)}=0$.  Thus we can rephrase the final equation on
918: $F^{(1,1)}$, as a condition on the metric.  It is simplest to write
919: this in the ``mathematical'' gauge $\bAmath=0$, in which it is
920: \be \label{e5}
921: c\cdot{\bf 1} = \omega^{i\bj}F_{i\bj}
922:  =\omega^{\bj i}\bpartial_{\bj} \Amath_i
923:  =\omega^{\bj i}\bpartial_{\bj} \left(G^{-1}\partial_i G \right) .
924: \ee
925: A metric $G$ satisfying this equation is a ``hermitian-Einstein'' 
926: metric.  It is simply related to a hermitian Yang-Mills connection
927: as above.
928: 
929: Finally, using the complex gauge transformation above, the standard
930: physical inner product
931: \be\label{eq:physvip}
932: \vev{u|v} \equiv \int_X (u^*)_a v^a
933: \ee
934: is equal to the natural inner product generalizing \eq{defH1},
935: \begin{eqnarray}\label{eq:mathvip}
936: \vev{u|v} &=& \vev{h~ s|h~ t} \\
937:  &=& i\int_X G_{a\bb}~ \bs^\bb~ t^a \,d{\vol_X}.
938: \end{eqnarray}
939: 
940: \subsection{Embedding vector bundles}
941: 
942: We now want to represent the hermitian metric $G_{a\bb}$ in 
943: the same way as we did for line bundles, by 
944: introducing a complete basis of sections.
945: Now an irreducible bundle $E$ with $c_1=0$, and thus of interest 
946: for string compactification, will not have global sections.
947: What we do instead is to make the same construction for 
948: $E(k) \equiv E \otimes \cL^k$, which will have global sections.
949: We can again think of these sections as a basis of polynomials 
950: approximating functions on which to base our numerical scheme.
951: 
952: Thus, consider a rank $r$ vector bundle $E$, and suppose 
953: that $E(k)$ has $N$ global sections.
954: Choosing a local frame as above,
955: a basis for these will be an $N$ by $r$ matrix $z_\alpha^a$.
956: This is defined up to a $GL(N)$ change of basis, and up to a 
957: $GL(r)$ change of frame. After making these identifications, such a
958: matrix $z$ defines a point in the Grassmannian $G(r, N )$  of $r$ planes in $\C^N$.
959: 
960: Given a metric $G_{a\bb}$ on the fibers of $E(k)$,
961: we can define the matrix of inner products
962: $$
963: H_{\alpha\bbeta} = \vev{z_\beta|z_\alpha}
964: $$
965: as above.  Such a metric could be obtained by multiplying a metric
966: $G^{(0)}$ on $E$, by one on $\cL^k$ as defined earlier.  Or, it
967: might simply be an $r\times r$ hermitian matrix of functions (in
968: each local frame) with appropriate transformation properties.
969: 
970: Now there is a natural set of metrics on $E(k)$ generalizing
971: \eq{Ks}, again parameterized by an $N\times N$ matrix,
972: defined by
973: $$
974: (G^{-1})^{a\bb} = g^{\alpha\bbeta} z_\alpha^a (z^\dag)_\bbeta^\bb\, ,
975: $$
976: where the dagger is hermitian conjugation.  Again, the approach will
977: be to find a natural metric in this class which is a good approximation
978: to the hermitian-Einstein metric. This will lead to a hermitian Yang-Mills
979: connection on $E(k)$.  But this is simply related to the hermitian
980: Yang-Mills connection on $E$, because twisting by $\cL^k$ only
981: modifies the trace part of the field strength.
982: 
983: \subsection{Generalized T-operator}
984: 
985: We will now turn to a proposal for a generalized
986: T-operator, which produces the hermitian-Einstein metric on a stable
987: vector bundle. To begin with we use results by Wang about balanced
988: metrics on such bundles \cite{Wang:metricsbundles}.
989: 
990: We consider again a polarized $n$ dimensional manifold $(X,\cL)$ and an
991: irreducible holomorphic vector bundle $E$ of rank $r$ on $X$. Then by
992: Kodaira embedding we know that for k sufficiently large, a basis
993: $z_{\alpha}^a$ of the global sections of $E(k)$ will
994: give rise to an embedding
995: $$
996: \xymatrix{
997: %   E (k)  = i_* U_r   \ar[r]\ar[d]&    U_r    \ar[d]     \\ 
998:         X~       \ar@{^{(}->}[r]^-i       &     G(r, N ) .   \\
999: }
1000: $$
1001: Now Wang proves the following:
1002: \begin{theo}\label{bundl1}
1003: $E$ is Gieseker stable iff there is an integer $k_0$ such 
1004: that for $k > k_0$, the {\em k}th embedding given as above can be moved to a balanced 
1005: place, i.e.,  there is a $g \in SL(N ,\C)$ which is unique up to left translation by $SU(N)$ such that: 
1006: $$ \frac{1}{V} \int_{g\cdot X} {z (z^\dag z)^{-1} z^\dag\ d V }=
1007: \frac{r}{N} I_{N \times N} .
1008: $$ 
1009: \end{theo} 
1010: 
1011: We call the equation above the ``balance equation.''
1012: In the case that $E$ is a line bundle,
1013: this definition reduces to that of a 
1014: balanced  embedding in $\mathbb{P}^{N-1}$. 
1015: 
1016: Now, let $h$ be a hermitian  metric on $\cL$ 
1017: and $H$ be a hermitian  metric on $E $, and fix the K\"ahler 
1018: form on $X $ to be $\omega = \ff i {2\pi} Ric(h)$. 
1019: Let $\vol$ denote the volume of $(X, \omega)$. 
1020: Suppose 
1021: ${S_1 , \ldots , S_N }$ is an orthonormal basis of $H^0(X,E (k))$  with respect to the induced $L^2$ -metric $\langle . \, ,  . \rangle$.
1022: The Szeg\"o kernel $B_k$ is a generalization of the function $\rho(\omega)$ defined in \eq{bk}. It is defined as the fiberwise homomorphism
1023: $$
1024: B_k(x) = \sum_{i=1}^N \langle . , S_i(x)\rangle S_i (x) \colon E_x \to E_x .
1025: $$ 
1026: This expression is independent of the choice of orthonormal basis.
1027: 
1028: Now the local 
1029: form of Theorem \ref{bundl1}  can be stated as follows (Corollary 1.1 of \cite{Wang:metricsbundles}):
1030: \begin{theo}
1031: E is Gieseker stable iff there is an integer $k_0$ such that for any
1032: $k > k_0$, we can find a metric $H^{ (k)}$, which we will call the {\em
1033: balanced metric} on $E (k)$, such that the Szeg\"o kernel satisfies
1034: the equation
1035: $$
1036: B_k (x)  = \frac{\chi(k)}{V~ r} I_E
1037: $$ 
1038: where $I _E$ is the identity bundle morphism 
1039: and $\chi(k)$ is the Hilbert polynomial of 
1040: $E$ with respect to the polarization $\cL$. 
1041: \end{theo}
1042: 
1043: The theorem tells us that if $E$ is Gieseker stable then for large $k$
1044: there is a balanced metric $H^{ (k)}$ on $E (k)$. Hence we will have a
1045: sequence of hermitian metrics $H_k := H^{ (k)}\otimes h^{-k} $ on $E$.
1046: The importance of the balanced metric $H^{ (k)}$ for physical
1047: applications follows from the following theorem:
1048: \begin{theo}
1049: Suppose $E$ is Gieseker stable. If $H_k \to H_\infty$ in the
1050: $C^\infty$ norm as $k \to \infty$, then the metric $H_\infty$ solves
1051: the ``weak hermitian-Einstein equation'',
1052: \begin{equation}
1053: \label{hym}
1054: \frac{i}{2\pi} \bigwedge F_{(E,H_\infty)} + \frac{1}{2}S(\omega) I_E=
1055: \left( \frac{deg(E)}{V r}+\frac{\bar{s}}{2} \right) \cdot I_E
1056: \end{equation}
1057: where $\bigwedge F_{(E ,H_\infty)}$ is the contraction of the
1058: curvature form of E with $\omega$, $S (\omega)$ is the scalar
1059: curvature of $X$ and $\bar{s} := \frac{1}{V} \int_X S
1060: \frac{\omega^n}{n!}$ .  Conversely, suppose there is a hermitian
1061: metric $H_\infty$ solving this equation, then $H_k \to H_{\infty}$ in
1062: $C^r$ norm for any $r$.
1063: \end{theo}
1064: 
1065: To prove (\ref{hym}) one can work along the same lines as in the proof
1066: of Theorem~\ref{theo2}, using Catlin's and Wang's results for the
1067: expansion of the Szeg\"o kernel.
1068: \begin{prop}
1069: 
1070: \begin{enumerate}
1071: \item For fixed hermitian metrics $H$ and $h$ on $E$ and ${\cal
1072: O}_X(1)$ respectively, there is an asymptotic expansion as $k \to \infty$
1073:   $$
1074:   B_k(H,h)\sim A_0(H,h)k^n +A_1(H,h)k^{n-1}+ \cdots,
1075:   $$
1076:   where $A_i(H,h) \in \Gamma(\End E)$ are smooth sections defined locally by $H$. 
1077: In particular,
1078:   $$
1079:   A_0(H,h)=I_E,\;\;A_1(H,h)=\frac{i}{2\pi}\bigwedge F(E,Ric(h)) +\frac{1}{2}S(\omega) \cdot I_E
1080:   $$
1081: \item The expansion holds uniformly in the $C^\infty$ norm; in the sense that for any $r,N>0$
1082:   $$
1083:   \| B_k(H,h) - \sum_{i=0}^N A_i(H,h)k^{n-i}      \|_{C^r}\leqslant K_{r,N,H,h} k^{n-N-1}
1084:   $$
1085:   for some constants $K_{r,N,H,h}$. 
1086: \end{enumerate}
1087: \end{prop}
1088: 
1089: Now we can repeat the steps of the argument outlined in Section~\ref{s:prs}. Under the assumption that $H_k \to H_\infty$ in $C^\infty$ we find that for  $r>0$ 
1090: $$
1091: \| B_k(H_k) - I_E k^{n}-  \frac{i}{2\pi}\bigwedge F(E,Ric(h)) +\frac{1}{2}S(\omega) \cdot I_E  k^{n-1}  \|_{C^r}\leqslant C k^{n-2}
1092: $$
1093: for some fixed constant $C$. By assumption $H^{(k)}$ is balanced, hence $B_k(H_k)=\chi(k)/r V I_E$. This implies that 
1094: $$
1095: \|\frac{i}{2\pi} \bigwedge F_{(E,H_\infty)} + \frac{1}{2}S(\omega) I_E-
1096: \left( \frac{deg(E)}{V r}+\frac{\bar{s}}{2} \right) \cdot I_E\|=O(k^1).
1097: $$
1098: 
1099: \subsubsection{Generalized T-operator }
1100: 
1101: Using the strong analogy between the construction of metrics with
1102: constant Kahler curvature and metrics on stable bundles which obey the
1103: hermitian-Einstein equation, we propose the following generalized
1104: T-operator:
1105: \begin{equation}\label{e4}
1106:  T(G)=\frac{N}{V r}\int_X {z (z^\dag G^{-1} z)^{-1} z^\dag \ d V },
1107: \end{equation}
1108: where as before, $z$ is an $N$ by $r$ matrix of holomorphic sections
1109: of $E$.
1110: 
1111: The relevance of this proposal follows from the following conjecture:
1112: \begin{conj}\label{gT}
1113: If a balanced embedding $i\colon X \hookrightarrow G(r,N)$ exists,
1114: then for every hermitian $N \times N$ matrix $G$, the sequence
1115: $T^r(G)$ converges to a fixed point $G_0$ as $r\to\infty$. Using an
1116: orthonormal basis with respect to $G_0$, the embedding is balanced,
1117: and as outlined above, it provides an approximate solution for the
1118: corresponding hermitian-Einstein equation.
1119: \end{conj}
1120: This conjecture may require additional technical assumptions, such
1121: as the earlier one of $\Aut(X,E)$
1122: being discrete.  We have not attempted to prove it, but would hope
1123: that this can be done along the same lines as \cite{Donaldson1,Sano}.
1124: 
1125: In the following section we will numerically test the conjecture for
1126: several stable vector bundles on $\mathbb{P}^2$, and on the Fermat
1127: quintic in $\P^4$, and find that it works for these cases.
1128: 
1129: \section{Examples}
1130: 
1131: \subsection{Hermite-Einstein metric on the tangent bundle of $\mathbb{P}^n$}
1132: 
1133: Let $\mathbb{P}^n$ be the complex projective space of dimension $n$, and $\{ Z_i \}_{i=0}^{i=n}$ 
1134: its homogeneous coordinates. We will work on the open set $Z_0 \neq 0$  and chose the  local inhomogeneous coordinates
1135: $w_{i}=Z_{i}/Z_{0}$. The Fubini-Study metric on $\P^n$  
1136: $$
1137: g_{i\bar{j}} = \frac{1}{1+\sum_i |w_i|^2} \delta_{i\bar{j}} - \frac{w_{i}\bar{w}_{j}}{(1+\sum_i |w_i|^2)^2}.
1138: $$
1139: is the unique maximally symmetric metric, with its group of Killing symmetries  isomorphic to $U(n+1)$. In addition, this metric is Einstein, that is its Ricci tensor is proportional to the metric itself. Therefore its associated curvature tensor  
1140: obeys the hermitian Yang-Mills  equation. The Donaldson-Uhlenbeck-Yau theorem then implies that the tangent bundle of $\P^n$, $T\mathbb{P}^n$, is a rank $n$ stable bundle on $\mathbb{P}^n$.\footnote{
1141: The stability of $T\mathbb{P}^n$ also has purely algebraic proof.}
1142: It follow from this that  the balanced metric on the bundle $T\mathbb{P}^n$ must be the Fubini-Study metric.
1143: 
1144: To describe the tangent bundle  $T\mathbb{P}^n$ we  use the Euler sequence 
1145: \begin{equation}\label{eulerseq}
1146: 0 \longrightarrow \mathcal{O}_{\mathbb{P}^n} \longrightarrow \mathcal{O}_{\mathbb{P}^n}(1)^{\oplus (n+1)}
1147: \longrightarrow T\mathbb{P}^n \longrightarrow 0.
1148: \end{equation}
1149: Here $ \mathcal{O}_{\mathbb{P}^n}(1)$ denotes the hyperplane line bundle. After twisting the sequence
1150: by $\mathcal{O}_{\mathbb{P}^n}(k)$ and taking the cohomology we find the short exact sequence (SES)
1151: \begin{equation}\label{e1}
1152: 0 \longrightarrow H^0({\mathbb{P}^n},\mathcal{O}_{\mathbb{P}^n}(k)) \longrightarrow H^0({\mathbb{P}^n}, \mathcal{O}_{\mathbb{P}^n}(k+1)^{\oplus (n+1)})
1153: \longrightarrow H^0({\mathbb{P}^n},T\mathbb{P}^n(k)) \longrightarrow 0.
1154: \end{equation}
1155: This gives an explicit description for $H^{0}(\P^n,T\mathbb{P}^n(k))$, which for sufficiently large $k$ gives the embedding
1156: \begin{equation}\label{e2}
1157: \mathbb{P}^n \hookrightarrow \mathrm{G}(n,\, W)
1158: \end{equation}
1159: where $W=H^{0}({\mathbb{P}^n},T\mathbb{P}^n(k))^*$, and $\mathrm{G}(n,\, W)$ is the Grassmannian of $n$-planes in $W$. 
1160: 
1161: Based on (\ref{e1}), we choose to describe the global sections of 
1162: $T\mathbb{P}^n(k)$ by an $n+1$ vector 
1163: $$
1164: (M_0,\ldots,M_n)
1165: $$ 
1166: where $\{M_i\}_{i=1}^n$ are arbitrary monomials of degree $k+1$ in
1167: the homogeneous coordinates $Z_i$, while $M_0$ is any degree $k+1$
1168: monomial which does not contain an $Z_0$.
1169: 
1170: Now we show  how to construct the embedding (\ref{e1}) for any $k\geq 0$. We start by choosing a frame $\{ \hat{e}_i \}_{i=0}^{n}$ for the  vector bundle 
1171: $\mathcal{O}(k+1)^{\oplus (n+1)}$. This amounts to choosing a section for every one the $n+1$ summands. For simplicity we chose the same section in every  summand. The Euler sequence (\ref{eulerseq}) imposes the condition 
1172: $$ 
1173: \sum_{i=0}^n Z_i \hat{e}_i =0.
1174: $$
1175: Locally this gives a frame for $T\mathbb{P}^n$, if we solve for
1176: $$
1177: \hat{e}_0 = -\sum_{i=1}^n \frac{Z_i}{Z_0}\hat{e}_i = -\sum_{i=1}^n w_i \hat{e}_i.
1178: $$
1179: Expanding the global sections of $T\mathbb{P}^n(k)$ in the local frame $\{\hat{e}_i \}_{i=1}^n$ gives an $n \times \dim(W)$ matrix, which is the explicit realization of our embedding \cite{GH}. 
1180: 
1181: To illustrate the procedure consider $T\mathbb{P}^2(0)$. $\mathcal{O}_{\P^2}(1)$ has 3 global sections: $Z_0,Z_1,Z_2$. Choosing $Z_0$ to be the local frame in every summand of $\mathcal{O}_{\P^2}(1)^{\oplus 3}$, and discarding the global section $Z_0$ from the first $\mathcal{O}_{\P^2}(1)$, we find the matrix
1182: \begin{equation}
1183: z=\left(\begin{array}{cccccccc}-w_1^2 & -w_1w_2 & 1 & w_1 & w_2 & 0 & 0 & 0 \\-w_1w_2 & -w_2^2 & 0 & 0 & 0 & 1 & w_1 & w_2\end{array}\right)
1184: \end{equation}
1185: 
1186: For an initial hermitian metric $G_{0}$ on the vector space 
1187: $W=H^{0}(\P^n, T\mathbb{P}^n(k))^*$, our generalized T-operator  (\ref{e4}) gives the iterations 
1188: $$
1189: G_{m+1}=T(G_m)=
1190: \frac{ \dim W}{n\, \mathrm{Vol}(\mathbb{P}^n)}\int_{\mathbb{P}^n} {z (z^\dag G_m^{-1} z)^{-1} z^\dag \ d V }.
1191: $$
1192:  
1193: We tested the converges of the $T$-map starting with $G_{0}=I$ in the case $n=2$ for $k=1,\ldots ,5$. In all cases we converged to a given $G_{\infty}$ for less than 10 iterations, with a precision of 0.1\%. 
1194: 
1195: The balanced metric $H^{(k)}$ on the vector bundle $T\mathbb{P}^n(k) $ induced by $G_{\infty}$ is given by
1196: %we can interpret
1197: %the rectangular matrix $z$ in two different ways:
1198: %First, as an array of global sections as we show in (\ref{interpz}). Second, at
1199: %each point $p\in \mathbb{P}^n$, $z_p$ defines a basis of vectors which spans the
1200: %$n$-plane in $W=H^{0}(T\mathbb{P}^n\otimes \mathcal{O}(k))^*$ which pulls back
1201: %o the fiber $T_p \mathbb{P}^n \otimes \mathcal{O}(k)$ under the universal map
1202: %$$
1203: %U: T\mathbb{P}^n\otimes \mathcal{O}(k)\longrightarrow U_n,
1204: %$$
1205: %given the embedding of $\mathbb{P}^n$ 
1206: %nto the Grassmannian, and the universal bundle $U_n\to \mathrm{G}(n,N)$ defined on it.
1207: %Thus, we can restrict fiberwise the balanced metric $G_{\infty}$, by choosing
1208: %as generators of the fiber $T_p \mathbb{P}^n \otimes \mathcal{O}(k)\hookrightarrow W$
1209: %the matrix $z_p$, what means that at $p$ the metric on the fiber is given by
1210: \begin{equation}
1211: H^{(k)}=(z^\dag G_{\infty}^{-1} z)^{-1} .
1212: \end{equation}
1213: %up to certain local function that we can determine. This defines globally a metric on
1214: %the vector bundle $T\mathbb{P}^n \otimes \mathcal{O}(k)$.  
1215: Let $h$ be Fubini-Study metric on the hyperplane bundle $\mathcal{O}_{\P^n}(1)$, that is the metric with constant scalar curvature. %such that the induced
1216: %curvature 2-form defines the Kahler form used to build up the Liouville's volume form
1217: %$dV$ on $\mathbb{P}^n$ by wedging it $n$ times
1218: Then the metric
1219: $$
1220: H_k := H^{(k)}\otimes h^{-k}=(z^\dag G_{\infty}^{-1} z)^{-1}\otimes h^{-k}
1221: $$
1222: is the balanced metric on $T\mathbb{P}^n$. Our numerical computations show that this is indeed  the Fubini-Study metric  on $T\mathbb{P}^n$, as explained earlier. The numerical  agreement is within 0.5\%. This provides the first non-trivial test of our conjecture.
1223: 
1224: \subsection{A stable rank 3 bundle over $\mathbb{P}^2$}
1225: 
1226: In this section we test our generalized T-operator on a rank 3 vector
1227: bundle $V^*$ over $\mathbb{P}^2$.  We first consider its dual $V$, defined by
1228: four linearly independent global sections $\{m_i\}$ of 
1229: ${\cal O}_{\mathbb{P}^2}(2)$ through the SES
1230: $$
1231: \xymatrix{
1232: 0\ar[r] &V \ar[r] &{\cal O}_{\mathbb{P}^2}^{\oplus 4} \ar[r]^-m &{\cal O}_{\mathbb{P}^2}(2) \ar[r] & 0.
1233: }
1234: $$
1235: This bundle has moduli, which are implicitly determined by the choice
1236: of the sections $\{m_i\}$.  Before choosing these, let us
1237: check stability, which does not depend on the specifics of this choice.
1238: 
1239: To check stability, we have to ensure that neither $V$ nor $\wedge^2V$
1240: have destabilizing line bundles. Using the canonical isomorphism
1241: $$
1242: \wedge^2 V = \det V \otimes V^*
1243: $$
1244: we find  the slopes
1245: $$
1246: \mu(V)=-2/3,\,\,\,\mu(\wedge^2 V)=-4/3.
1247: $$
1248: Since $Pic(\mathbb{P}^2)=\mathbb{Z}$,  all line bundles are of the form ${\cal O}_{\mathbb{P}^2}(p)$ for some $p$.
1249: Hence it is sufficient to show that 
1250: $$
1251: H^0({\mathbb{P}^2},V)=0,\;\;\;H^0({\mathbb{P}^2},\wedge^2 V(1))=0.
1252: $$
1253: The first fact follows from the defining sequence of $V$, if we assume that $\{m_i\}$ are linearly independent. To prove the second statement we use 
1254: $$
1255: H^0({\mathbb{P}^2},\wedge^2 V(1))=H^0({\mathbb{P}^2},V^*(-1))=H^2({\mathbb{P}^2},V(-2))^*.
1256: $$
1257: Again, this statement follows easily from the defining sequence of $V$.
1258: Finally, stability of $V$ implies stability for $V^*$. 
1259: 
1260: We will now compute the hermitian Yang-Mills  connection on $V^*$ using our generalized T-operator. 
1261: First observe that $V^*(k)$ fits into the short exact sequence
1262: \begin{equation}\label{Vdual}
1263: 0\to {\cal O}_{\mathbb{P}^2}(k-2) \to {\cal O}_{\mathbb{P}^2}(k)^{\oplus 4} \to V^*(k)\to 0.
1264: \end{equation}
1265: This leads to another SES
1266: $$0\to 
1267: H^0({\mathbb{P}^2},{\cal O}_{\mathbb{P}^2}(k-2)) \to H^0({\mathbb{P}^2},{\cal O}_{\mathbb{P}^2}(k)^{\oplus 4}) \to H^0({\mathbb{P}^2},V^*(k))\to 0.
1268: $$
1269: We can use this expression for an explicit parameterization of $H^0({\mathbb{P}^2},V^*(k))$.  
1270: 
1271: For concreteness let us choose to be  four global sections $\{m_i\}_{i=1}^4$ defining $V$  to be 
1272: $$
1273: Z_1Z_2,\;Z_0Z_1,\; Z_0Z_2,\;Z_0^2.
1274: $$
1275: Now we choose a frame $\{\hat{e_i}\}$ for ${\cal O}_{\mathbb{P}^2}(k)^{\oplus 4}$. The defining equation (\ref{Vdual}) of $V^*(k)$ imposes the condition $\sum_i m_i \hat{e_i} =0$, and gives a frame for $V^*$. Locally we can solve for $\hat{e}_0$, and working in inhomogeneous coordinates $w_i$ we find that
1276: $$
1277: \hat{e}_0= -\frac{1}{w_2}\hat{e}_1-\frac{1}{w_1}\hat{e}_2-\frac{1}{w_1w_2}\hat{e}_3.
1278: $$
1279: Expanding the global sections of $H^0({\mathbb{P}^2},V^*(k))$ in the frame $\{\hat{e}_i\}_{i=1}^3$ gives a matrix, which is the embedding map. 
1280: 
1281: We studied the convergence of our generalized $T$-operator numerically for $k=2,3$ and 4, and found that convergence was achieved for less than 10 iterations. As before, the  metric on $V^*(k)$ is given by
1282: \begin{equation}
1283: H^{(k)}=(z^\dag G_{\infty}^{-1} z)^{-1} ,
1284: \end{equation}
1285: while the corresponding metric on $V^*$ is
1286: $$
1287: H_k := H^{(k)}\otimes h^{-k}=(z^\dag G_{\infty}^{-1} z)^{-1}\otimes h^{-k},
1288: $$
1289: where $h$ is again the Fubini-Study metric on ${\cal O}_{\mathbb{P}^2}(1)$.
1290: 
1291: Since in this case the balanced metric on $V^*$ is not a priori known,
1292: one needs a different approach, than used in the previous section for
1293: $T\P^2$, to test how close is the approximate balanced metric to
1294: satisfying the hermitian Yang-Mills equation. But this quite easy to
1295: do numerically once the balanced metric $G_{\infty}$ is known, as all
1296: we need to do is to check how close we are to satisfying
1297: Eq.~(\ref{e5}). In all cases considered Eq.~(\ref{e5}) was satisfied
1298: to within 1\% accuracy.
1299: 
1300: \subsection{A rank 3 bundle on the Fermat quintic}
1301: 
1302: In this section we turn to a much more complicated case than our previous examples, that of a stable rank 3 bundle on the Fermat quintic $Q$ in $\P^4$:
1303: \begin{equation}
1304: Q:\qquad  Z_0^5+Z_1^5+\cdots +Z_4^5=0.
1305: \end{equation}
1306: 
1307: Testing our generalized T-operator in this case necessitates knowledge of the Ricci flat metric on the Fermat quintic. For this we use Donaldson's original T-operator  \eq{defH}, which we turn to first.
1308: 
1309: \subsubsection{Ricci flat metric on Fermat quintic}
1310: 
1311: We consider the embedding of $Q$ given by the complete linear system of cubics, $H^{0}(Q, \mathcal{O}_Q (3))$, whose
1312: complex projectivization is isomorphic to $\mathbb{P}^{34}$. The balanced metric will be the restriction of a Fubini-Study metric on $\mathbb{P}^{34}$. An indirect test that this has indeed vanishing Ricci curvature is included in the next section. 
1313: 
1314: In order to do practical calculations with  Donaldson's  T-map, we have to perform the  integrals on $Q$ numerically. We introduce
1315: a discrete approximation to the Calabi-Yau volume form $d\mu_{\Omega}=\Omega\wedge\bar{\Omega}$,
1316: defining it by a weighted set of $M$ points $\{ x_a \}_{a=1}^M\in Q$, with masses $\nu_a$:
1317: \begin{equation}\label{numint}
1318: \int_Q \, \big( \,\,\, \big) \, d\mu_{\Omega} \approx \sum_{a=1}^M\, \big( \,\,\, \big) \, \delta(x-x_a)\nu_a.
1319: \end{equation}
1320: This numerical measure  gives an accurate approximation to the analytical one for large $M$.
1321: In our computations we build $10$ different samples of 100,000 points, which we use independently to 
1322: iterate the T-map until convergence is reached, i.e.,  the sequence $\{ T^{r}(G_0) \}_{r=0}$
1323: obeys
1324: $$
1325: \vert\vert T^{r+1}(G_0) - T^r (G_0) \vert \vert < \epsilon.
1326: $$
1327: In our simulations the fixed point of this discrete version of the $T$-map was  
1328: reached after 15-20 iterations.
1329: Each weighted point set gave rise to a convergent sequence. 
1330: The 10 different hermitian forms $\{ G^{e}_{\infty} \}_{e=1}^{10}$ 
1331: approximating the balanced metric in $\P H^{0}(Q,\mathcal{O}_Q (3))$ agree up to 
1332: \begin{equation}
1333: \label{errordef}
1334: \mathrm{max}\, \Bigg[ 
1335: \frac{\sigma(G^{e}_{\infty})}{\vert \langle G^{e}_{\infty} \rangle \vert} \Bigg]
1336: \approx 0.9 \% ,
1337: \end{equation}
1338: where $\vert \langle G^{e}_{\infty} \rangle \vert$ is the average matrix of the 
1339: ten different outputs and $\sigma(G^{e}_{\infty})$ is the standard deviation matrix.
1340: The ratio $\sigma(G^{e}_{\infty})/ \vert \langle G^{e}_{\infty} \rangle \vert$ is computed entry by entry,
1341: and the maximum is taken over all entries.
1342: We used the average $\langle G^{e}_{\infty} \rangle$ as approximation for the balanced metric on
1343: $H^{0}(Q,\mathcal{O}_Q (3))$.
1344: 
1345: \begin{figure}[h]
1346:       \centering
1347:       \includegraphics[width=\textwidth]{balancedandnonbalanced.eps}
1348:       \caption{The shape of the rational curve for the balanced and non-balanced metrics.} \label{p:bm}
1349: \end{figure}
1350: 
1351: To get a visual picture of the geometry implicit in the construction, 
1352: we consider the rational curve $t\colon\mathbb{P}^1 \hookrightarrow Q$,
1353: defined locally by the parametrization 
1354: \begin{equation}
1355: \label{RatCur}
1356: \left(\begin{array}{ccccc} 1, & -1, &  t, & 0, & -t
1357: \end{array}\right)
1358: \end{equation}
1359: with $t\in \mathbb{C}\cup\infty$. Take the sections $Z_1^3+Z_4^3$ and
1360: $Z_0^3$ from $H^{0}(Q,\mathcal{O}_{Q}(3))$, and consider the function
1361: $s=(Z_1^3+Z_4^3)/Z_0^3=w_1^3+w_4^3 $. In Fig.~\ref{p:bm} we consider
1362: the real function $|s|^{2}_{G}$ restricted to the rational curve
1363: (\ref{RatCur}), where we take the stereographic projection of the
1364: complex $t$-plane and for a given $t$ we plot $|s(t)|^{2}_{G}$ in the
1365: radial direction. For the balanced metric $\langle G^{e}_{\infty}
1366: \rangle$ the deviation from being spherical is small. For comparison
1367: we also show the same plot for the case of a generic non-balanced
1368: hermitian form $G$ with random entries.
1369: 
1370: \subsubsection{Solution of the hermitian  Yang-Mills  equation}
1371: 
1372: In this section we use the generalized T-operator to produce a hermitian  Yang-Mills  connection on a rank three stable vector bundle $V$ on the Fermat quintic $Q$. We also implicitly test that the previously obtained balanced metric on $Q$ indeed has vanishing Ricci curvature.
1373: 
1374: We define the rank three bundle $V$  by the following SES
1375: $$
1376: \xymatrix{
1377: 0\ar[r] &{\cal O}_Q(-1) \ar[r]^-{\beta} & {\cal O}^{\oplus 4}_Q  \ar[r] & V \ar[r] & 0.
1378: }
1379: $$
1380: $\beta$ is given by four generic global sections of ${\cal O}_Q(1)$, which do not intersect on $Q$, hence $V$ is indeed a vector bundle. In addition, the  first Chern class of $V$ is $
1381: c_1(V)=H,
1382: $
1383: hence $V$ is not a simple twist of the tangent bundle of $Q$.
1384: That fact that $V$ is stable was proved in \cite{DRY}.
1385: 
1386: Once again, we use 
1387: $$
1388: \xymatrix{
1389: 0\ar[r] &{\cal O}_Q(k-1) \ar[r] & {\cal O}^{\oplus 4}_Q(k)  \ar[r] & V(k) \ar[r]& 0		,}
1390: $$
1391: and its associated long exact sequence in cohomology 
1392: $$
1393: \xymatrix{
1394: 0\ar[r] &H^0(Q,{\cal O}_Q(k-1)) \ar[r] & H^0(Q,{\cal O}^{\oplus 4}_Q(k))  \ar[r] & H^0(Q,V(k))\ar[r]& 0,
1395: }
1396: $$
1397: to derive a frame for $V$ and an explicit parameterization for the global sections.
1398: We choose $\beta=(Z_0,\ldots, Z_3)$. Using the frame $\{\hat{e}_i\}_{i=0}^4$ for ${\cal O}^{\oplus 4}_Q$, we also get a frame for $V$ with the relation
1399: $$
1400: \hat{e}_0=-\sum_{i=1}^3 w_i \hat{e}_i.
1401: $$
1402: 
1403: In this paper we restrict to the case $k=1$ for which $\dim H^{0}(Q,V(1))=19$. The coordinate matrix 
1404: \begin{equation}
1405: \label{embbedG }
1406: z(w)=\left(\begin{array}{ccccccc}1 \ldots w_4 & 0 & 0 & -w_1^2 & -w_1w_2 & -w_1w_3 & -w_1w_4 \\0 & 1 \ldots w_4 & 0 & -w_1w_2 & -w_2^2 & -w_3w_2 & -w_4w_2 \\0 & 0 & 1 \ldots w_4 & -w_1w_3 & -w_2w_3 & -w_3^2 & -w_4w_3\end{array}\right)
1407: \end{equation}
1408: gives the embedding into the Grassmannian $Q\hookrightarrow G(3,19)$.
1409: 
1410: Using the integration techniques described in the previous section,  we iterate the generalized T-operator. We reach the fixed point after 12-15 iterations for several different samples of weighted points which approximate
1411: the analytical measure, allowing us to estimate the balanced metric for $H^{0}(Q,V(1))$ with an error
1412: of 1.1\%.\footnote{We estimate the errors using (\ref{errordef}).} 
1413: 
1414: The metric on $V(1)$ is given by
1415: \begin{equation}\label{HEmet}
1416: H=(z^\dag G_{\infty}^{-1} z)^{-1} ,
1417: \end{equation}
1418: To test the accuracy of this metric  we  evaluate the right hand side of the hermitian  Yang-Mills equations (\ref{e5}).
1419: We find  the mean to be
1420: $$
1421: \langle \omega^{i\bj}F_{i\bj} \rangle=\frac{1}{\mathrm{Vol}(Q)}\int_Q \big( \omega^{i\bj}F_{i\bj} \big) d\mu_{\Omega} \approx 1.31\times \mathbf{I}_{3\times 3}
1422: $$
1423: with $\mathbf{I}_{3\times 3}$ the $3\times 3$ identity matrix. In our conventions the theoretical value of the constant is 4/3. The standard deviation of the individual matrix elements is
1424: $$
1425: \sigma\big( \omega^{i\bj}F_{i\bj} \big) =\mathrm{max}\left[\sqrt{\frac{1}{\mathrm{Vol}(Q)}\int_Q \Big( \omega^{i\bj}F_{i\bj} - \langle \omega^{i\bj}F_{i\bj} \rangle \Big)^2 d\mu_{\Omega}}\ \right] \approx 0.15,
1426: $$
1427: where the  square-root and the square are performed entry by entry.
1428: Therefore, $\omega^{i\bj}F_{i\bj}$ is a global constant on $Q$ times the identity, within an error of  $ 0.15/1.31\approx 11 \%$. This implies that the hermitian Yang-Mills equation (\ref{e5}) is satisfied with this  accuracy. 
1429: 
1430: Testing the hermitian Yang-Mills equation provides an implicit test of Ricci flatness, since it is precisely the Ricci flat metric that is needed in the hermitian Yang-Mills equation. If we had gotten this metric wrong, then we would have had no chance of satisfying the hermitian Yang-Mills equation.
1431: 
1432: \begin{figure}[h]
1433:       \centering
1434:       \includegraphics[width=0.35\textwidth, angle=105]{wavefunction.eps}
1435:       \caption{The probability density 
1436:                     in the radial direction on the rational curve.}\label{p:wf}
1437: \end{figure}
1438: 
1439: Finally, to visualize the construction,  
1440: in Fig.~\ref{p:wf}  we took the rational curve defined in (\ref{RatCur}), and we plotted
1441: the function $|\Psi|_G^2$, where
1442: $$
1443: \Psi =(w_1 w_4)\hat{e}_1 + (w_2 w_4)\hat{e}_2 + (w_3 w_4)\hat{e}_3 
1444: $$
1445: and $G$ is the balanced metric we obtained.
1446: If we interpret $\Psi$ as a wave-function, then Fig.~\ref{p:wf} exhibits the probability density $\langle \Psi \vert \Psi \rangle $ in the radial direction, restricted to the rational curve (\ref{RatCur}).
1447: 
1448: \subsubsection*{Acknowledgements}
1449: 
1450: We would like to thank B. Shiffman and S. Zelditch for valuable
1451: advice at the beginning of this project.
1452: This research was supported in part by the DOE grant DE-FG02-96ER40949.
1453: 
1454: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1455: 
1456: \input p.bbl
1457: 
1458: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1459: \end{document}
1460: 
1461: \def\p{\partial}
1462: \begin{equation}
1463: \p A\cdot G^{-1}\cdot \left(\mathbf{1} - A^\dag (AG^{-1}A^\dag)\cdot (AG^{-1})
1464: \right)\cdot \bar{\p}A^\dag\cdot (AG^{-1}A^\dag)^{-1}
1465: \end{equation}
1466: