1: \documentclass[12pt]{article}
2: %\usepackage[notref, notcite]{showkeys}
3: \usepackage{amsmath,amssymb,theorem,graphicx,color}
4:
5: \setlength{\textwidth}{165mm}
6: \setlength{\topmargin}{-.51cm}
7: \setlength{\oddsidemargin}{0mm}
8: \setlength{\evensidemargin}{0pt}
9: \setlength{\textheight}{218mm}
10: \usepackage[all]{xy}
11:
12: %%%%%% to help proofreading
13: \usepackage{color}
14: \definecolor{dblue}{rgb}{0,0,.7}
15: \definecolor{indigo}{RGB}{50,0,105}
16:
17: \usepackage[dvips, hyperindex=true,
18: colorlinks=true, linkcolor=dblue,
19: citecolor=indigo, pagecolor=dblue,
20: bookmarksopen=false, urlcolor=dblue,
21: linktocpage]{hyperref}
22:
23: \def\P{{\mathbb P}} \def\O{\mathcal{O}}
24: \def\C{{\mathbb C}}
25: \def\ff#1#2{{\textstyle\dfrac{#1}{#2}}}
26: \def\CY{Calabi-Yau} \def\c#1{\mathcal{#1}}
27: \def\R{{\mathbb R}}
28: %%%%%%%%%% Some commands %%%%%%%%%%%%%%
29: \newcommand{\be}{\begin{equation}}
30: \newcommand{\ee}{\end{equation}}
31: \newcommand{\eq}[1]{Eq.~(\ref{#1})}
32: %
33: \def\H{\operatorname{H}}
34: \newcommand{\Aphys}{{A^{(phys)}}}
35: \newcommand{\bAphys}{{{\bar A}^{(phys)}}}
36: \newcommand{\Amath}{{A^{(math)}}}
37: \newcommand{\bAmath}{{{\bar A}^{(math)}}}
38: %
39: \newcommand{\bbc}{{\mathbb C}}
40: \newcommand{\bbz}{{\mathbb Z}}
41: \newcommand{\bbf}{{\mathbb F}}
42: \newcommand{\bbr}{{\mathbb R}}
43: \newcommand{\rk}{{\text{ rk} }}
44: \newcommand{\Aut}{{\operatorname{Aut}}}
45: \newcommand{\vol}{{\text{vol} }}
46: \newcommand{\Tr}{{\text{ Tr}~ }}
47: \renewcommand{\Re}{{\text{ Re}~ }}
48: \newcommand{\End}{{\text{ End}~ }}
49: \newcommand{\Coh}{{\text{ Coh} }}
50: \newcommand{\effective}{{\text{effective} }}
51: \newcommand{\ch}{{\text{ch} }}
52: \newcommand{\sgn}{{\text{sgn}\,}}
53: \newcommand{\KC}{{\it KC}}
54: \newcommand{\tH}{{\tilde H}}
55: \newcommand\ba{\bar{a}}
56: \newcommand\bb{{\bar{b}}}
57: \newcommand\bc{\bar{c}}
58: \newcommand\bd{\bar{d}}
59: \newcommand\bz{\bar{z}}
60: \newcommand\bZ{\bar{Z}}
61: \newcommand\bW{\bar{W}}
62: \newcommand\bD{\bar{D}}
63: \newcommand\bA{\bar{A}}
64: \newcommand\bB{\bar{B}}
65: \newcommand\bR{\bar{R}}
66: \newcommand\bS{\bar{S}}
67: \newcommand\bT{\bar{T}}
68: \newcommand\bU{\bar{U}}
69: \newcommand\bPi{\bar{\Pi}}
70: \newcommand\bphi{\bar{\phi}}
71: \newcommand\balpha{{\bar{\alpha}}}
72: \newcommand\bbeta{{\bar{\beta}}}
73: \newcommand\bOmega{\bar{\Omega}}
74: \newcommand\bpartial{\bar{\partial}}
75: \newcommand\bj{{\bar{j}}}
76: \newcommand\bi{{\bar{i}}}
77: \newcommand\bk{{\bar{k}}}
78: \newcommand\bm{{\bar{m}}}
79: \newcommand\btheta{\bar{\theta}}
80: \newcommand\bpsi{\bar{\psi}}
81: \newcommand\bF{\bar{F}}
82: \newcommand\bs{{\bar{s}}}
83: \newcommand\bt{\bar{t}}
84: \newcommand\bv{\bar{v}}
85: \newcommand\bx{\bar{x}}
86: \newcommand\by{\bar{y}}
87: \newcommand\btau{\bar{\tau}}
88: \newcommand\bC{\bar{C}}
89: \newcommand\bX{\bar{X}}
90: \newcommand\bY{\bar{Y}}
91: \newcommand{\bra}[1]{{\langle}#1|}
92: \newcommand{\ket}[1]{|#1\rangle}
93: \newcommand{\bbra}[1]{{\langle\langle}#1|}
94: \newcommand{\kket}[1]{|#1\rangle\rangle}
95: \newcommand{\vev}[1]{\langle{#1}\rangle}
96: %
97: \newcommand{\nch}{{\text{ch}}}
98: \newcommand{\td}{{\text{td}\; }}
99: \newcommand{\pt}{{\text{pt} }}
100: %
101: \newcommand{\cp}[1]{{\mathbb P}^{#1}}
102: \newcommand{\op}[1]{\operatorname{#1}}
103: \newcommand{\ocp}[1][]{{\mathcal
104: O}_{\cp{1}}{#1}}
105: \newcommand{\ocpn}[2][]{{\mathcal
106: O}_{\cp{#1}}{#2}}
107: \newcommand{\oy}[1][]{{\mathcal O}_B{#1}}
108: \newcommand{\oz}[1][]{{\mathcal
109: O}_{B^{'}}{#1}}
110: \newcommand{\ox}[1][]{{\mathcal O}_X{#1}}
111: \newcommand{\om}{{\mathcal O}_M}
112: \newcommand{\MW}{{\mathbb M}{\mathbb W}}
113: \newcommand{\mw}[1]{[#1]}
114: \newcommand{\Pic}{\op{Pic}}
115: \newcommand{\FM}{\boldsymbol{F}{\boldsymbol{M
116: }}}
117: \newcommand{\fm}{\boldsymbol{f}\boldsymbol{m}
118: }
119: \newcommand{\T}{\boldsymbol{T}}
120: \newcommand{\ct}{\boldsymbol{t}}
121: \newcommand{\D}{\boldsymbol{D}}
122: \newcommand{\num}[1]{{\#}\text{\bfseries #1}}
123: \newcommand{\rhom}{R^{\bullet}{\mathcal H}om}
124: \newcommand{\cH}{{\mathcal H}}
125: \newcommand{\cL}{{\mathcal L}}
126: \newcommand{\cV}{{\mathcal V}}
127: \newcommand{\cW}{{\mathcal W}}
128: \newcommand{\heck}{\boldsymbol{H}\boldsymbol{
129: e}\boldsymbol{c}
130: \boldsymbol{k}\boldsymbol{e}}
131: \newcommand{\hup}[2]{\heck^{+}_{(#1)}(#2)}
132: \newcommand{\hdown}[2]{\heck^{-}_{(#1)}(#2)}
133: \newcommand{\homsh}[3]{{\mathcal
134: H}om_{#1}(#2,#3)}
135: \newcommand\Hom{{\rm Hom}}
136: \newcommand\Ext{{\rm Ext}}
137: \newcommand{\p}[1]{{\mathbb P}(#1)}
138: \newcommand{\bl}[2]{\operatorname{Bl}_{#1}#2}
139: \newcommand{\ses}[3]{\xymatrix{
140: 0 \ar[r] & {#1} \ar[r] & {#2} \ar[r] & {#3}
141: \ar[r] & 0 \\ }}
142:
143: %%%%%%%%% Theorems and the like %%%%%%%%%
144: \newtheorem{theo}{Theorem}[section]
145: \newtheorem{lem}[theo]{Lemma}
146: \newtheorem{cor}[theo]{Corollary}
147: \newtheorem{prop}[theo]{Proposition}
148: \newtheorem{defi}[theo]{Definition}
149: \newtheorem{conj}[theo]{Conjecture}
150: {\theorembodyfont{\rmfamily}
151: \newtheorem{remark}[theo]{Remark}}
152:
153: % Put preprint number in top-right.
154: \def\pplogo{\vbox{\kern-\headheight\kern
155: -29pt
156: \halign{##&##\hfil\cr&{\ppnumber}\cr\rule{0pt
157: }{2.5ex}&\ppdate\cr}}}
158: \makeatletter
159: \def\ps@firstpage{\ps@empty
160: \def\@oddhead{\hss\pplogo}%
161: \let\@evenhead\@oddhead}
162: \def\maketitle{\par
163: \begingroup
164: \def\thefootnote{\fnsymbol{footnote}}
165: \def\@makefnmark{\hbox{$^{\@thefnmark}$\hss}
166: }
167: \if@twocolumn
168: \twocolumn[\@maketitle]
169: \else \newpage
170: \global\@topnum\z@ \@maketitle
171: \fi\thispagestyle{firstpage}\@thanks
172: \endgroup
173: \setcounter{footnote}{0}
174: \let\maketitle\relax
175: \let\@maketitle\relax
176: \gdef\@thanks{}\gdef\@author{}\gdef\@title{}
177: \let\thanks\relax}
178: \makeatother
179:
180: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
181: %%%
182:
183: \begin{document}
184: \setcounter{page}0
185: \def\ppnumber{\vbox{\baselineskip14pt
186: \hbox{hep-th/0612075}}}
187: \def\ppdate{RUNHETC-2006-32} \date{}
188:
189: \title{\bf \LARGE
190: Numerical Calabi-Yau metrics
191: \\[10mm]}
192: \author{\bf
193: {Michael R. Douglas$^\&$,
194: Robert L.~Karp,
195: Sergio Lukic}\\
196: {\bf and Ren\'e Reinbacher} \\
197: [10mm]
198: \normalsize Department of Physics, Rutgers
199: University \\
200: \normalsize Piscataway, NJ 08854-8019 USA
201: \\ [10mm]
202: \normalsize $^\&$ I.H.E.S., Le Bois-Marie,
203: Bures-sur-Yvette, 91440 France}
204:
205: {\hfuzz=10cm\maketitle}
206:
207: \vskip 1cm
208:
209: \begin{abstract}
210: \normalsize
211: \noindent
212: We develop numerical methods for approximating Ricci flat metrics on
213: Calabi-Yau hypersurfaces in projective spaces. Our approach is based
214: on finding balanced metrics, and builds on recent theoretical work by
215: Donaldson. We illustrate our methods in detail for a one parameter
216: family of quintics. We also suggest several ways to extend our
217: results.
218: \end{abstract}
219:
220: \vfil\break
221:
222: \tableofcontents
223:
224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225: %%%%%%
226: \section{Introduction} \label{s:intro}
227:
228: Calabi-Yau manifolds were proven to admit Ricci flat metrics in
229: \cite{Yau}. Explicit expressions for such metrics would have many
230: applications in mathematics and in string compactification; however it
231: is widely thought that for compact Calabi-Yau manifolds
232: no closed form expression exists,
233: except in trivial cases (tori and orbifolds).
234: Thus it is useful to develop methods for constructing and working with
235: approximate Ricci-flat metrics; and for the solutions of related
236: equations, such as hermitian Yang-Mills, and the other equations of
237: string compactification.
238:
239: One such approximating method is finding the balanced metric on an
240: embedding of the manifold into complex projective space,
241: given by the sections of a holomorphic line bundle.
242: This method was
243: suggested by Yau in the early 90's, and was proven to work in a
244: fundamental paper by Tian \cite{Tian:metrics}. Since then many people
245: have worked on the problem of balanced metrics in various contexts
246: \cite{Bourguignon:Yau,Luo:Toper,Zelditch:Szego,Donaldson1,Donaldson2},
247: just to name a few. A recent result of Donaldson
248: \cite{Donaldson:numeric} shows that this approximation scheme is both
249: mathematically elegant and relatively easy to implement
250: numerically. It can also be directly generalized to find approximate
251: hermitian Yang-Mills connections on vector bundles \cite{en:Mike1},
252: which can in turn be used to compute metrics on moduli spaces, and the
253: kinetic terms in $N=1$ string compactifications \cite{en:Mike3}.
254:
255: In this work we develop numerical methods for approximating Ricci flat
256: metrics on Calabi-Yau hypersurfaces based on these ideas. This also
257: supplies a detailed analysis of the numerical methods used in
258: \cite{en:Mike1}. We study the effectiveness of our approach in the
259: example of a one parameter family of quintics in $\C\P^4$. As we
260: review in section \ref{s:T}, we work with a space of approximating metrics
261: parameterized by an $N\times N$ hermitian matrix; the balanced metric
262: is then the fixed point of the so-called ``T map'' on this space,
263: defined in \eq{T}.
264:
265: The main computational problem in implementing the T map numerically
266: is the evaluation of a large number of integrals on the manifold.
267: More precisely, given a Calabi-Yau $n$-fold
268: $X$, with its corresponding holomorphic $n$-form
269: $\Omega \in
270: \Omega^{n,0}(X)=\Lambda^{n}(T^*X)^{1,0}$, and
271: volume form $d\mu_{\Omega} = \Omega\wedge
272: \overline{\Omega}$,
273: one needs to compute integrals of the type
274: \begin{equation}\label{e1}
275: \int_{X}f \; d\mu_{\Omega},
276: \end{equation}
277: where $f\colon X\to \C$ is a smooth complex
278: valued (but not holomorphic) function.
279: Consequently, the heart of the paper (section 3) will
280: be devoted to developing a numerical
281: approximation scheme to efficiently
282: and accurately compute these integrals.
283:
284: A second technical point, which is very valuable in simplifying these
285: computations, is to take advantage of the discrete symmetries of the
286: manifold. We discuss this in section \ref{s:sym}.
287:
288: Our explicit numerical results appear in section \ref{s:NR}, where
289: we also provide a general discussion of the
290: efficiency and accuracy of the algorithm, comparisons with alternatives,
291: and suggestions for future work.
292:
293: Before we begin, let us briefly set out the problem.
294: Denote the Ricci flat metric on $X$ (which is unique
295: given a complex structure
296: and K\"ahler class) as $g_{RF}$. We want to propose a set of
297: approximating metrics $g_h$ parameterized by parameters $h$, and
298: give a numerical procedure to
299: find the ``best'' approximation to $g_{RF}$ within this set.
300:
301: The criteria that a best approximation should satisfy include
302: \begin{enumerate}
303: \item {\bf Accuracy}: we want to minimize the error
304: $\epsilon=d(g_h,g_{RF})$, where $d$ is some measure of the
305: distance between the approximate and true metrics.
306: A simple and natural choice for $\epsilon$
307: in the present context is to consider the function
308: \begin{equation}\label{define-eta}
309: \eta_h = \frac{\det \omega_h}{\Omega\wedge\bar\Omega}
310: \end{equation}
311: on $X$, where $\omega_h$ is the K\"ahler form for $g_h$.
312: For a Ricci flat metric, this will be the constant function. We then take \footnote{
313: We consider only compact varieties, hence both the minimum and
314: the maximum are attained.}
315: \begin{equation}\label{define-error}
316: \epsilon = 1-\frac{{\rm min}_{x\in X}~\eta_h(x)}{{\rm max}_{x\in X}~\eta_h(x)} .
317: \end{equation}
318: Of course, one could use other norms, such as
319: $||\eta_h-\frac{1}{{\rm vol}~X}\int\eta_h||_p$, or curvature integrals.
320: \item {\bf Control}: we want an explicit bound on the error,
321: \begin{equation}\label{control-eta}
322: \epsilon(g_h,g_{RF}) < \epsilon_{max},
323: \end{equation}
324: depending on the parameters of the problem.
325: \item {\bf Systematic improvement}: we would like to have a control
326: parameter $k$, such that by increasing $k$, we can bring the error estimate
327: $\epsilon_{max}$ down to any desired accuracy.
328: \item {\bf Mathematical naturalness}.
329: Our experience with string theory (and more
330: generally in mathematics and physics)
331: has been that in exploratory work such as this,
332: rather than trying to incorporate all known aspects of a problem and
333: find a ``best'' solution, we can learn far more
334: by studying a well chosen simplification in depth. This favors a scheme
335: in which one makes the smallest possible number of arbitrary or
336: {\it ad hoc} choices not inherent in the original statement of the problem.
337: \end{enumerate}
338: Of course the approximation
339: should be efficiently computable as well. We will comment
340: on these various aspects as they arise.
341:
342: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
343: %%%%%%%%%%%%
344: \section{T-map} \label{s:T}
345:
346: In this section we review the construction
347: of balanced metrics, which in turn lead to a
348: convenient approximation of the Ricci flat
349: metric on Calabi-Yau threefolds. Our
350: presentation is based on \cite{en:Mike1},
351: which in turn was deeply inspired by
352: \cite{Donaldson:numeric}. We refer to these
353: papers, and the references therein, for more
354: details.
355:
356: We start with a holomorphic line bundle $\cL$
357: on a K\"ahler manifold $X$,
358: with $N$ global sections. This fact is
359: usually phrased as $\H^{0}(X,\,
360: \mathcal{L})=\C^N$.
361: Let $\{s_\alpha\}_{\alpha=1}^N$ be a basis of
362: this vector space, and consider the map
363: $$
364: i\colon X\longrightarrow \cp{N-1}
365: \qquad i(Z_0,\ldots,Z_n) =
366: (s_1(Z),s_2(Z),\ldots,s_N(Z)).
367: $$
368: The geometric picture is that each point in
369: our original manifold $X$
370: (with coordinates $Z_i$)
371: corresponds to a point in $\bbc^{N}$
372: parameterized by the sections
373: $s_\alpha$. Since choosing a different frame
374: for $\cL$ would produce
375: an overall rescaling $s_\alpha\rightarrow
376: \lambda s_\alpha$, the overall
377: scale is undetermined. Granting that
378: $s_1(Z),s_2(Z),\ldots,s_N(Z)$
379: do not vanish simultaneously, this gives us a
380: map into $\P^{N-1}$.
381:
382: We want this map to be an
383: embedding, {\it i.e.} that distinct points on $X$
384: map to distinct points on $\mathbb{P}^{N-1}$, and that tangent vectors
385: are also separated. In general, we can
386: appeal to the Kodaira embedding
387: theorem, which asserts that for $\cL$ ample
388: this will be true for all powers $\cL^k$,
389: starting with some $k_0$. As an example, for
390: non-singular quintics in $\P^4$, $\mathcal{O}_{X}(k)$
391: is both ample and very ample for all $k\ge
392: 1$.
393:
394: Next we consider the $N^2$-parameter family of K\"ahler
395: potentials on $\cp{N-1}$,
396: \be\label{Ks}
397: K_{h} = \log \left(\sum_{\alpha,\bbeta}
398: h^{\alpha\bbeta}
399: s_\alpha \bs_\bbeta \right)\equiv \log
400: ||s||^2_h\, ,
401: \ee
402: parametrized by the $N$ by $N$ hermitian matrix $h$. These
403: give rise to a family of K\"ahler metrics on $X$
404: by restriction, and
405: we will seek a ``best'' approximation to the Ricci flat metric
406: within this space of approximating metrics.
407: Note that it is the
408: inverse of $h$, $h^{\alpha\bbeta}$, that
409: appears in \eqref{Ks}. The reason for this will become clear
410: shortly.
411:
412: Mathematically, the simplest interpretation
413: of \eqref{Ks} is that it
414: defines a hermitian metric on the line bundle
415: $\cL $. This is a sesquilinear map
416: from $\bar\cL\otimes \cL$ to smooth functions
417: $C^\infty(X)$, here defined by
418: $$
419: (s,s') = e^{-K_{h}}\cdot\bar s\cdot s' =
420: \frac{\bar s\cdot s'}{\sum_{\alpha,\bbeta}
421: h^{\alpha\bbeta}
422: s_\alpha \bs_\bbeta} .
423: $$
424: Notice that a change of frame, which acts on
425: our sections
426: as $s_\alpha\rightarrow \lambda s_\alpha$,
427: cancels out in this expression.
428:
429: This metric allows us to define an inner
430: product between the global sections:
431: \begin{equation}\label{T1}
432: \vev{s_\beta|s_\alpha} = \int_X
433: \frac{s_{\alpha}\bs_{\bbeta}}{||s||^2_h} \;
434: d{\vol_X}.
435: \end{equation}
436: Note that this inner product depends on $h$
437: in a nonlinear way, since $h$ appears in the
438: denominator, and $||s||_h$ involves the
439: inverse of $h$.
440: Here $d\vol_X$ is a volume form on $X$, which
441: has to be chosen.
442:
443: One choice which makes sense for any $X$ is
444: $d{\vol_X}=\det\omega_h$, the standard volume form $\sqrt{\det g_h}$.
445: For $X$ a Calabi-Yau variety, we can instead use
446: $d{\vol_X}=\nu= \Omega\wedge \bar{\Omega}$. The latter is
447: significantly simpler, mainly
448: because it is $h$-independent. Our numerical procedure will use a discrete
449: approximation to this volume form.
450:
451: The explicit form of the expression in
452: \eqref{T1} suggests that one studies the map
453: \begin{equation}\label{T}
454: T(h)_{\alpha\bbeta} = \frac{N}{\vol(X)}\int_X
455: \frac{s_\alpha
456: \bs_\bbeta}{||s||^2_h}\, {d\vol_X},
457: \end{equation}
458: dubbed the ``T-map'', which acts on the
459: space of hermitian matrices in a non-linear
460: way.
461:
462: A fixed point of this map, $T(h) = h $,
463: the pair $(h,s_{\alpha})$ is called a {\em balanced} embedding, and the
464: metric on $X$
465: associated to the corresponding K\"ahler
466: potential \eqref{Ks} is called the balanced
467: metric.\footnote{In a more precise nomenclature, which we will
468: not use here, the term ``balanced embedding'' is reserved for the
469: definition with $d{\vol_X}=\det\omega_h$, while
470: the alternate definition with $d{\vol_X}=\nu$
471: is referred to as the ``$\nu$-balanced embedding.''
472: Also, there are several other equivalent ways of
473: defining the notion of balanced metric, which
474: are closely linked to the notion of stability
475: in Mumford's sense, but this utilitarian
476: definition will suffice for our purposes. For
477: reviews of this topic we recommend
478: \cite{Thomas:GITrev,Mabuchi}.}
479: As it turns
480: out, the balanced metric is unique,
481: up to $U(N)$ transformations of the basis of sections and
482: rescaling, provided that the manifold in
483: question has no continuous symmetries. This
484: is certainly the case for a \CY\ variety, and
485: in particular our quintics.
486:
487: It turns out that the T-map is contracting, so
488: the simplest way to find a fixed point of the
489: T-map is to iterate it.
490: We have the following
491: \begin{theo} (see, e.g. ,\cite{Sano,Donaldson1};
492: and \cite{Bourguignon:Yau,Donaldson:numeric} for the $\nu$-balanced case)
493: Suppose that $\Aut(X,L)$ is discrete. If a
494: balanced embedding
495: exists, then, for any initial hermitian
496: matrix $G_0$, as
497: $r\to \infty$ the sequence $T^r(G_0)$
498: converges to a fixed point.
499: \end{theo}
500:
501: The importance of balanced metrics stems from
502: a theorem that goes back to Tian
503: \cite{Tian:metrics} and Zelditch
504: \cite{Zelditch:Szego}.
505: Let us consider the sequence of balanced metrics associated to
506: the bundles ${\mathcal L}^k$ (defined in terms
507: of the Fubini-Study metric $\omega_k^{FS}$ from \eq{Ks}):
508: \begin{equation}
509: \omega_k=\frac{1}{k}\, i_k^*(\omega_k^{FS}) .
510: \end{equation}
511: The rescaling is made so that
512: the cohomology class of the K\"ahler form
513: $[\omega_k]= c_1(\cL)\in
514: \H^2(X,\mathbb{Z})$ is independent of $k$.
515: With this definition one has
516: \begin{theo}\label{theo2}
517: Suppose $\Aut(X,\cL)$ is discrete and
518: $(X,\cL^k)$ is balanced for
519: sufficiently large $k$. If the metrics
520: $\omega_k$ converge
521: in the $C^\infty$ norm to some limit
522: $\omega_\infty$ as $k\to \infty$, then
523: $\omega_\infty$ is a K\"ahler metric in the
524: class
525: $ c_1(\cL)$ with {\em constant} scalar curvature.
526: \end{theo}
527:
528: The constant value of the scalar curvature is
529: determined by $c_1(X)$.
530: In particular, for $c_1(X)=0$ the scalar
531: curvature is zero. Thus,
532: the balanced metrics $\omega_k$, in the large
533: $k$ limit, converge to
534: the Ricci flat metric. Furthermore, in this
535: case the $1/k$ convergence is enhanced to a
536: $1/k^2$ convergence. We will see this
537: explicitly in Section~\ref{s:NR}.
538:
539: One may ask where the complex
540: structure and K\"ahler moduli enter in this setup.
541: The complex structure enters implicitly through the basis
542: of holomorphic sections
543: $s_\alpha$. As for the K\"ahler
544: class, this is $c_1(\cL)$. Of course, the
545: Ricci flatness condition is scale
546: invariant, so the overall scale is
547: irrelevant; however the point of
548: this is that if $h^{1,1}>1$, then by
549: appropriately choosing $\cL$ we
550: choose a particular ray in the K\"ahler cone.
551:
552: Therefore, if we can find the unique balanced metric for a given
553: $\cL$, we have a candidate approximation scheme. Let us evaluate it
554: by our criteria. One great advantage is that we have a control
555: parameter $k$, which is easy to implement and is mathematically
556: natural. The balanced metric is natural in many respects,
557: which should make it possible to get error bounds like
558: \eq{control-eta}, though this has not yet been done.
559:
560: On the other hand, the balanced metric is not in general the
561: most accurate approximation within this class of metrics
562: to $g_{RF}$. As discussed in
563: \cite{Donaldson:numeric}, one can find other series of metrics which
564: converge to $g_{RF}$ faster than $1/k^2$.
565:
566: As an illustration, once we choose our measure of accuracy
567: $\epsilon$, say \eq{define-error}, we can propose a simple
568: scheme which is guaranteed to be the most accurate possible. It is a
569: two-step procedure in which we take the balanced metric as the
570: starting point for a numerical search in the space of parameters
571: $h$ for the metric which minimizes
572: $\epsilon$. As we discuss a bit later,
573: standard numerical optimization routines will work for this purpose
574: if the starting point is close enough to the actual minimum.
575: This would not be the most efficient possible approach; one could
576: improve the efficiency by using information about the linearized
577: problem, as discussed in \cite{Donaldson:numeric}.
578:
579: For the applications we have in mind, for example being able to detect
580: anomalously large or small numbers in observables (perhaps having to
581: do with singularities), control, mathematical naturalness and ease of
582: programming are more important than accuracy, and thus we stick to the
583: balanced metric in our present work.
584:
585: As another simple physical illustration, suppose that by using the
586: techniques of \cite{en:Mike3} we could use these results to get
587: canonically normalized fields and physical Yukawa couplings in
588: quasi-realistic compactifications. At this point we would probably be
589: much happier to have results for quark and lepton masses in a variety
590: of models which were guaranteed accurate to within a factor of $2$,
591: than to have ``probable'' $10\%$ accuracy in one model. Of course
592: this is a rather long term goal, but the point should be clear.
593:
594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595: %%%%%%%%%%%%
596: \section{Numerical integration on \CY\
597: varieties}
598:
599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
600: %%%%%%%%%%%%
601: \subsection{Basic setup}\label{s:nm}
602:
603: It is clear from the outset that analytic evaluation of the
604: integrals appearing in the T-map \eqref{T} is not possible. On the
605: other hand, if the integrands are smooth and relatively slowly varying
606: functions, it will be possible to evaluate the integrals using Monte
607: Carlo methods. This is clear for the sections themselves. Since $h$
608: is positive definite, the denominator in \eqref{T} is strictly
609: positive, mitigating (though not eliminating) the possibility of
610: numerical blow-ups.
611:
612: Let $X$ be a compact Calabi-Yau
613: $n$-fold,\footnote{Although most of what we
614: present generalizes to varieties other than
615: \CY , we restrict attention to these spaces
616: due to their importance in string theory. }
617: with its corresponding holomorphic $n$-form
618: $\Omega \in \Lambda^{n,0}(X)$. The volume
619: form $\Omega\wedge \overline{\Omega}$
620: determines a natural measure $d\mu_{\Omega}$
621: on $X$ in the sense that
622: \[
623: \int_{X}f \,d\mu_{\Omega}=\int_{X}f
624: \;\Omega\wedge \overline{\Omega}.
625: \]
626: From now on we will not distinguish between a
627: top form and the associated measure.
628:
629: We can use $d\mu_{\Omega}$ to measure
630: volumes. For an open set $\mathcal{U}\subset
631: X$ the indicator or characteristic function
632: $\mathbf{1}_{\mathcal{U}}$ is defined by
633: \begin{displaymath}
634: \mathbf{1}_{\mathcal{U}}(x) =
635: \left\{ \begin{array}{lll}
636: 1 && \text{if $x \in \mathcal{U}$}\\
637: 0 && \text{if $x \notin \mathcal{U}$}.
638: \end{array}
639: \right.
640: \end{displaymath}
641: The measure of $\mathcal{U}$ is its volume
642: \begin{equation}\nonumber
643: \mu_{\Omega}(\mathcal{U})=\int_{X}
644: \mathbf{1}_{\mathcal{U}}\,d\mu_{\Omega}
645: =\vol(\mathcal{U}).
646: \end{equation}
647:
648: To do a Monte Carlo integration,
649: one would ideally like to produce samples of
650: points on $X$ which are uniformly
651: distributed according to the measure
652: $d\mu_{\Omega}$. This means that for every
653: sample of points $\{ q_i\in X
654: \}_{i=1}^{N_p}$,
655: the expected number of points within each
656: open subset $\mathcal{U}\subset X$ is
657: \[
658: \sum_{i=1}^{N_{p}} \mathbf{1}_{\mathcal{U}}(q_i) =
659: N_p\frac{\mu_{\Omega}(\mathcal{U})}{\mu_{
660: \Omega}(X)}.
661: \]
662: Using this, we can estimate integrals as finite sums:
663: \begin{equation}\label{e3}
664: \int_{X}f d\mu_{\Omega}\approx
665: \mu_{\Omega}(X)\frac{ 1
666: }{N_p}\,\sum_{i=1}^{N_p} f(q_i),
667: \end{equation}
668: The statistical error of such an
669: approximation is of order
670: ${1}/{\sqrt{N_p}}$ times a quantity
671: proportional to the mean of the $f(q_i)$'s.
672: \cite{NR}.
673: % \sigma(f) \approx \sqrt{\sum_i \big( E(f)-\vol(X)\times
674: %f(q_i)\big)^2},
675:
676: In practice, producing samples of points which are distributed
677: according to the measure $\mu_{\Omega}$ is not so easy.
678: One way to overcome this problem is by producing
679: points which are uniformly distributed
680: according to another auxiliary measure, say
681: $d\mu_A$. Let us assume that $d\mu_A$ is
682: associated to the global top form $A$.
683: The ratio $\Omega\wedge \overline{\Omega}/ A$ is a
684: global function on $X$, which we call the {\em mass function} $m_A$.
685: At a point $x$ it is defined to be the ratio of the two top forms
686: evaluated at $x:$
687: \begin{equation}\label{e33}
688: m_A(x) = \frac{\Omega\wedge
689: \overline{\Omega}(x)}{A(x)}.
690: \end{equation}
691: In general this function is neither constant nor holomorphic.
692:
693: While one could use this information to generate a sample distributed
694: according to $d\mu_\Omega$ ({\it e.g.}, by rejection sampling or MCMC),
695: it is simplest to explicitly put the mass function into the integrand.
696: Thus,
697: given a sample of points distributed according to $d\mu_A$,
698: and the mass function, we can estimate \eq{e1} as
699: \begin{equation}\label{e2}
700: \int_{X}f d\mu_{\Omega} = \int_{X}f
701: \,\frac{d\mu_{\Omega}}{d \mu_{A}} \,d\mu_{A}
702: \approx \frac{\mu_{\Omega}(X)}{\sum
703: m_j}\sum_{i=1}^{N_p} f(q_i) m(q_i),
704: \end{equation}
705: The presence of the mass function increases the statistical error.
706: On the other hand, the generic values of our mass function are
707: order one, and this is a very mild penalty.
708:
709: Rather than regarding the Monte Carlo as a way to estimate the
710: original T-map, an alternate point of view is to regard the right-hand
711: side of \eq{e2} as defining a new measure $\nu$ and a new T-map,
712: leading to a new $\nu$-balanced metric which approximates the desired
713: $\Omega\wedge\bar\Omega$-balanced metric. An advantage of this point
714: of view is that in \cite{Donaldson:numeric} it is shown that (under a
715: very mild hypothesis on $\nu$) the new T-map is contracting, and the
716: new $\nu$-balanced metric is unique. Thus, numerical pathologies will
717: not enter at this stage, provided that we use the {\em same} sample of
718: points throughout the computation of the balanced metric. This is also
719: advantageous for efficiency reasons, so we always do this. One can
720: then repeat the computation with different samples to estimate the
721: statistical error.
722:
723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
724: %%%%%%%%%%%%
725: \subsection{Generating the sample}
726:
727: We now discuss how to efficiently generate points according to a known
728: simple distribution. In this paper we restrict to the case of $X$
729: a degree $d$ hypersurface in $\P^{n+1}$. For definiteness let $X$
730: be defined as the zero locus of the degree $d$ homogenous polynomial
731: $f$. The case of a complete intersection is a straightforward
732: extension. Our main interest will be in $d=n+2$, but we can be more
733: general for the time being.
734:
735: First, it is easy to generate random points distributed according to
736: the Fubini-Study measure (for any $h$) in the ambient $\P^{n+1}$. We
737: simply generate uniformly distributed points on $S^{2n+3}$, a standard
738: numerical problem, and then mod out the overall phase.
739:
740: Using this distribution, one approach to generating points on $X$
741: would be to keep only the points that lie sufficiently close to $X$,
742: in other words satisfy the defining equation of $X$ with a given
743: precision, and then use a root finding method (say Newton's method) to
744: ``flow'' down to $X$. In essence this is a rejection-type algorithm.
745: We implemented this strategy, but it has some
746: problems. First, it is hard to control the emerging distribution on
747: $X$ (this depends on details of the root finding method). Second, it
748: is an order of magnitude slower than the second method we are about to
749: describe.
750:
751: The approach we use starts by taking a pair of independently
752: chosen random points $(X,Y)\in \P^{n+1}\!\times\! \P^{n+1}$, which define
753: a random line in $\P^{n+1}$. By Bezout's theorem, a generic complex
754: line in $\P^{n+1}$ intersects $X$ in precisely $d$ points, and we take
755: these $d$ points with equal weight. Repeating this process $M$ times
756: generates some random distribution of $N_p=dM$ points.
757:
758: One advantage of this approach is that finding all $d$ roots of $f(z)=0$
759: numerically is not much harder than finding one root. But the main
760: advantage, as we show in Section~\ref{s:ec}
761: using results by Shiffman and Zelditch on
762: zeroes of random sections, is that that the resulting points are
763: distributed precisely according to the Fubini-Study measure restricted
764: to $X$. The mass function \eqref{e33} is then computable quite efficiently.
765:
766: A possible disadvantage for some applications is that the resulting
767: sample will have correlations between the points in each $d$-fold subset.
768: For our purpose of Monte Carlo integration, this is not a problem, as
769: \eqref{e2} is the expectation value of a function of a single random
770: variable, and does not see these correlations. If one were considering
771: functions of several independent random variables, one would probably
772: want to further randomize the sequence (say by permuting points
773: between subsets) to remove these correlations.
774:
775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
776: %%%%%%%%%%%%
777: \subsection{Expected values of
778: currents}\label{s:ec}
779:
780: Let us start with a smooth compact algebraic
781: variety $X$, and an ample line bundle $\c L$
782: on $X$. As reviewed in Section~\ref{s:T},
783: this means that $\c L^k$ defines an embedding
784: $i_k$ into projective space for any $k\geq
785: k_0$, for some positive integer $k_0:$
786: \begin{equation}\label{e4}
787: i_k \colon X \longrightarrow\P \H^{0}(X,\,
788: \mathcal{L}^k)^{\ast}.
789: \end{equation}
790: The idea is to consider {\em random} global
791: sections of $\c L^k$, distributed uniformly
792: according to a natural measure, and look at
793: the expected value of the zero locus that
794: they cut out in $X$. For this it is
795: convenient to use the Poincare dual
796: formulation, where the divisor associated to
797: a section becomes a form, and ask what is the
798: expected value of the random forms. Shiffman
799: and Zelditch answer this question, and the
800: more general one when we intersect $l$ such
801: divisors, in full generality using the
802: language of currents. This section is a brief
803: review of some aspects of their work
804: \cite{Shiffman:Zelditch1,Shiffman:Zelditch2}.
805: For brevity we adapt their results to fit our
806: needs, rather than reproducing them verbatim.
807:
808: The space of global sections
809: $\Gamma=\H^{0}(X,\, \mathcal{L}^k)$ is a
810: complex vector space of dimension $d_k$. If
811: we choose a basis for it, then it
812: automatically defines a hermitian inner
813: product, with respect to which the basis in
814: question is orthonormal. Conversely, given a
815: hermitian inner product $\langle \cdot, \cdot
816: \rangle$ on $\Gamma$, there is an orthonormal
817: basis $\c B=\{s_1,\ldots,s_{d_k}\}$ on
818: $\Gamma$. Now given $s\in\Gamma$, we can
819: expand it in the basis $\c B$, and the inner
820: product induces a complex Gaussian
821: probability measure on $\Gamma$:
822: \begin{equation}
823: d\gamma(s)=\frac{1}{\pi^m}\,e^{-||c||^2}d^{
824: d_k}c\,,\qquad \txt{where
825: $s=\sum_{j=1}^{d_k}c_js_j$ and
826: $||c||^2=\sum_{j=1}^{d_k} |c_j|^2$ }\,.
827: \end{equation}
828: Given a metric $h$ on the line bundle $\c L$,
829: as explained in \eqref{T1}, $h$ defines a
830: hermitian metric on $\Gamma$. This is the
831: inner product that we are going to use on
832: $\Gamma=\H^{0}(X,\, \mathcal{L}^k)$
833: throughout this section.
834:
835: Given a random variable $Y$ on the
836: probability space $(\Gamma, d \gamma)$,
837: the expected value of $Y$ in the
838: probability measure $d \gamma$ is
839: \begin{equation}\label{e66}
840: E(Y) =\int_{\Gamma}Y d \gamma.
841: \end{equation}
842:
843: We can think of the probability space
844: $(\Gamma, d \gamma)$ in a slightly different
845: way. Consider the unit sphere
846: \[
847: \c S \Gamma= \c S \H^0(X, \c L^{k})=\{s\in
848: \H^0(X, \c L^{k})\colon\langle s, s \rangle
849: =1\}.
850: \]
851: The Gaussian probability measure on $\Gamma$
852: restricts to the {\em uniform} measure $d
853: \mu$ on $\c S \Gamma$. The
854: expected value of $Y|_{\c S\Gamma}$ is
855: $E(Y) =\int_{\c S\Gamma}Y d\mu $. On the
856: other hand, the uniform measure on the sphere
857: $\c S \Gamma$ descends to the Fubini-Study
858: measure on the projectivization $\P \Gamma$.
859: This alternative view will be very useful later on.
860:
861: If we choose a section $s\in \Gamma= \H^0(X,
862: \c L^{k})$, then there is a divisor $Z_s$
863: associated to it, which, roughly speaking, is
864: the zeros of $s$ minus the poles of $s$.
865: Since we work with $\c L^{k}$ very ample,
866: $Z_s$ consists of only the zero locus of $s$.
867: Given the probability measure $d \gamma$ on
868: $\Gamma$, we can choose $s$ randomly, and
869: ask what is the expected value of the random
870: variable $Z$ (defined by $s\mapsto Z_s$).
871: This same question can be asked in an
872: equivalent form using Poincare duality. The
873: Poincare dual of $Z_s$ is a $(1,1)$ form
874: $T_s$, and it is more convenient to work with forms in this
875: context than to work with divisors. In general $T_s$ is not a
876: $C^\infty$-form on $X$, but it can be given
877: an explicit expression using the notion of
878: currents, i.e., distribution valued forms.
879:
880: Currents are defined as it is customary in the
881: theory of distributions.\footnote{For a nice
882: introduction to distributions and currents
883: in algebraic geometry the reader can consult
884: \cite{GH}, Chapter 0 resp. Chapter 3.}
885: Let $\Omega^{p,q}_0(X)$ be the space of
886: compactly supported $C^{\infty}$
887: $(p,q)$-forms on $X$, and for now we assume
888: that $\dim X=n$. The space of
889: $(p,q)$-\emph{currents} is the distributional
890: dual of $\Omega_0^{n-p,n-q}(X)$:
891: $\mathcal{D}^{\, p,q}(X) =
892: \Omega_0^{n-p,n-q}(X)^\prime$.
893: An element of $\mathcal{D}^{\, p,q}(X)$ is a
894: linear functional on $ \Omega_0^{n-p,n-q}(X)$
895: which continuous in the $C^{\infty}$ norm.
896:
897: The usefulness of currents in our context
898: stems from the fact that Poincare dual
899: $T_{Y}$ of an algebraic subvariety $Y$,
900: defined by
901: $$
902: \int_{X}T_{Y}\wedge \alpha =
903: \int_{Y}\iota^{\ast}(\alpha) ,
904: \qquad\mbox{for any $\alpha\in \Omega_0^{\dim
905: Y,\dim Y}(X)$},
906: $$
907: oftentimes has an explicit form in terms of
908: currents ($\iota\colon Y\hookrightarrow X$ is the embedding).
909: Let us focus on the case when $Y$
910: is a hypersurface, or more generally the zero
911: divisor of a section $s\in \Gamma= \H^0(X,
912: \c L^{k})$. In this case the current is given
913: by the \emph{Poincare-Lelong} formula:
914: \begin{equation}\nonumber
915: T_{s} =
916: \frac{i}{\pi}\,\partial\bar{\partial}\,\log
917: \langle s, s \rangle\,\, \in \mathcal{D}^{\,
918: 1,1}(X).
919: \end{equation}
920: $T_s$ is also known as the zero current of
921: ${s}$. Thus, the {Poincare-Lelong} formula
922: induces a map
923: $$
924: T\colon\P \H^{0}(X,\, \c L^k) \longrightarrow
925: \mathcal{D}^{\, 1,1}(X),\qquad s\mapsto T_s.
926: $$
927:
928: As discussed earlier, the Fubini-Study
929: measure makes $\P \H^{0}(X,\, \c L^k)$ into a
930: probability space, and we can also view $T$
931: as a random variable. Since the currents form
932: a linear space, we can inquire about their
933: expected value in this probability measure
934: $$
935: E(T) = \int_{\P \H^{0}(X,\, \c L^k)}T_s\;
936: d\mu_{FS}(s),
937: $$
938: As it happens oftentimes in the theory of
939: distributions, although $T_s$ is not a
940: $C^{\infty}$ form, $E(T)$ is, and we have the
941: following proposition:
942:
943: \begin{prop} (\cite[Lemma
944: 3.1]{Shiffman:Zelditch1}) \label{p:1}
945: With $X$ and $\c L$ as above, for $k$
946: sufficiently large so that Kodaira's map
947: $i_k$ associated to $\H^{0}(X,\, \c L^k)$, as
948: defined in \eqref{e4}, is an embedding, the
949: expected value of the random variable $T$
950: representing the zero current is
951: $$E (T_s) = \frac{1}{k}\, i_k^{\ast} \;
952: \omega_k^{FS},$$
953: where $\omega_k^{FS}$ is the Fubini-Study
954: 2-form on $\P\H^{0}(X,\, \c L^k)$, and
955: $i_k^{\ast}$ is pullback of forms (in this
956: case restriction).
957: \end{prop}
958:
959: This result generalizes to the case when we
960: intersect several divisors, and this will be
961: the case of main interest to us. Since
962: intersection of subvarieties is Poincare dual
963: to the wedge product, there is an obvious
964: guess how Prop.~\ref{p:1} should generalize.
965: Let $s_1,\ldots,s_m$ be sections of
966: $\H^{0}(X,\, \c L^k)$, and consider the zero
967: current $T^{(m)}_{s_1,\ldots,s_m} $
968: associated to the set
969: \[
970: Z_{s_1,\ldots,s_m} = \{x\in X\colon
971: s_1(x)=\cdots = s_m(x)=0\}.
972: \]
973: It is obvious that if we consider any $m$
974: linear combinations of $s_1,\ldots,s_m$
975: (which are themselves linearly independent),
976: then they determine the same zero set
977: $Z_{s_1,\ldots,s_m}$. Hence
978: $Z_{s_1,\ldots,s_m}$ is really associated to
979: the $m$-plane spanned by $s_1,\ldots,s_m$ in
980: $\H^{0}(X,\, \c L^k)$. Therefore the
981: probability space is the Grassmannian of
982: $m$-dimensional subspaces of $\H^0(X,\c
983: L^{k})$, with its natural Haar measure
984: $d\mu_{Haar}$ (a generalization of the
985: Fubini-Study measure). So we can ask what is
986: the expected value of this zero current,
987: computed with $d\mu_{Haar}$.
988:
989: \begin{prop} (\cite[Lemma
990: 4.3]{Shiffman:Zelditch1}) \label{p:2}
991: In the notation of Prop. \ref{p:1}, the
992: expected value of the zero current
993: $T^{(m)}$, associated to the simultaneous
994: vanishing of $m$ random sections of
995: $\H^{0}(X,\, \c L^k)$, distributed according
996: the Haar measure of the corresponding
997: Grassmannian, is
998: \begin{equation}\label{e7}
999: E(T^{(m)}) = k^{m-1}\,\big( i_k^{\ast} \;
1000: \omega_k^{FS}\big)^m
1001: \end{equation}
1002: \end{prop}
1003:
1004: Note in particular, that unlike in Prop.
1005: \ref{p:1}, the distribution is according to
1006: the Haar measure, but the final result still
1007: involves the Fubini-Study form on
1008: $\H^{0}(X,\, \c L^k)$. This is in fact
1009: natural, given that an $m$-tuple of sections,
1010: each distributed according to Fubini-Study on
1011: $\H^{0}(X,\, \c L^k)$, is the same thing as
1012: one $m$-plane, distributed according to Haar
1013: on the Grassmannian.
1014:
1015: As usual, besides having expected values,
1016: random variables also have variance. The zero
1017: current $T^{(m)}$ is no different in this
1018: respect. Its variance has been computed
1019: recently in \cite[Theorem
1020: 1.1]{Shiffman:Zelditch2}. In particular, it
1021: was shown that the ratio of the variance and
1022: the expected value goes to zero as $k$ is
1023: increased
1024: $$\frac{\left[ \operatorname{Var}\big(T^{(m)}
1025: \big) \right]^{1/2}}{E\big(T^{(m)} \big)}
1026: \sim k^{-\frac{ m}{2} -\frac {1}{4}}.$$
1027:
1028: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1029: %%%%%%%%%%%%
1030: \subsection{The \CY\ case}
1031:
1032: In this section we put all the pieces
1033: together, and explicitly show how to build
1034: the numerical measure
1035: $\{ q_i\in X,\, m(q_i) \}_{i=1}^{N_p}$
1036: introduced in Section~\ref{s:nm}. We will
1037: focus on a smooth Calabi-Yau hypersurface $X$
1038: in $\P^{n+1}$ of degree $n+2$. Let $X$ be
1039: given by the zero locus of the degree $n+2$
1040: homogeneous polynomial $f$, and let $(Z_{0},
1041: Z_{1},\ldots Z_{n+1})$ be the homogeneous
1042: coordinates on $\P ^{n+1}$. We denote the
1043: embedding by
1044: \[
1045: i \colon X=\mathcal{Z}(f)\hookrightarrow\P
1046: ^{n+1}.
1047: \]
1048:
1049: Our approach is to generate random points on
1050: $X$ using random lines on $\P ^{n+1}$, by
1051: looking at the intersection of these random
1052: lines with $X$. We can view a random line as
1053: the intersection of $n$ random hyperplanes.
1054: This allows us to compute the expected value
1055: of the corresponding zero current using the
1056: techniques of Section~\ref{s:ec}.
1057:
1058: For computational purposes, designing an
1059: algorithm to generate points on $X$ in such a
1060: fashion is straightforward. To generate a
1061: random line on $\P ^{n+1}$ we generate two
1062: random points, which lie on the unit sphere
1063: $\mathrm{S}^{2n+3}\subset\C^{n+2}$, and are
1064: distributed uniformly on this sphere.
1065: For instance, to generate random points
1066: uniformly on $\mathrm{S}^{2n+3}$ we can start
1067: with the unit cube in $\R^{2n+4}$, i.e.,
1068: $[-1,\,1]^{2n+4}\subset \R^{2n+4}$.
1069: Using a good quality random number
1070: generator
1071: %\footnote{It is well known in
1072: %computational circles that random
1073: %number generators built into languages are
1074: %usually low quality.}
1075: we generate an uniform distribution of points
1076: in $[-1,\,1]^{2n+4}$. Now take {\em only}
1077: those points which
1078: fall within the unit disk
1079: $\mathrm{D}^{2n+4}$, and then project them
1080: radially
1081: to the boundary $\partial \mathrm{D}^{2n+4}
1082: =\mathrm{S}^{2n+3}$.
1083:
1084: The intersection of the random line with $X$
1085: can be computed by restricting the defining
1086: polynomial $f$ to the line.
1087: As a result, computing the common zero locus
1088: reduces to solving for the roots of a
1089: polynomial of degree $n+2$ in one variable.
1090: We find numerically the $n+2$ roots using
1091: the Durand-Kerner algorithm \cite{Durand,Kerner},
1092: which is a refinement of the multidimensional
1093: Newton's method applied to a polynomial.
1094: This whole approach turns out to be
1095: very efficient in practice, in that one can
1096: generate a million points on a quintic in a
1097: matter of seconds.
1098:
1099: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1100: %%%%%%%%%%%%
1101: \subsubsection{The expected zero current}
1102:
1103: We chose to work with the hyperplane line
1104: bundle $\c L=\O_X(1)$ on $X$. $\O_X(1)$ is
1105: ample, and its global sections are in one to
1106: one correspondence with the homogeneous
1107: coordinates of the ambient $\P ^{n+1}$
1108: (restricted to $X$). The associated Kodaira
1109: embedding is precisely the defining one:
1110: \[
1111: i_1 \colon X\hookrightarrow\P
1112: ^{n+1}=\P\H^{0}(X,\, \O_X(1))^*.
1113: \]
1114: If we take $n$ sections of $\c L=\O_X(1)$,
1115: and look at their common zero locus, then by
1116: Bezout's theorem this is generically $n+2$
1117: points (degenerations might occur).
1118: Therefore, considering random $n$-tuples of
1119: sections will give random $n+2$-tuples of
1120: points on $X$. But now we can tell how these
1121: points are distributed, provided that the
1122: sections were distributed according to the
1123: Fubini-Study measure on $\P\H^{0}(X,\,
1124: \O_X(1))=\P ^{n+1}$. Using Prop.~\ref{p:2} we
1125: know that the expected value of the zero
1126: current associated to the $n+2$ points of
1127: intersection is $\big( i_1^{\ast} \;
1128: \omega^{FS}_{\P ^{n+1}}\big)^n$. This is an
1129: $(n,n)$ form on $X$, and plugging it into
1130: \eqref{e33} we obtain the mass formula
1131: \begin{equation}\label{f2}
1132: m(x)=\frac{\Omega\wedge
1133: \overline{\Omega}}{\big( i_1^{\ast} \;
1134: \omega^{FS}_{\P ^{n+1}}\big)^n }(x).
1135: \end{equation}
1136:
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138: %%%%%%%%%%%%
1139: \subsubsection{The numerical mass}\label{s:nM}
1140:
1141: Let us look at the two differential forms
1142: involved in \eq{f2}. For this we first
1143: choose affine coordinates
1144: $w_{a}=Z_{a}/Z_{0}$, $i=1,2,\ldots,n+1$ on
1145: $\P^{n+1}$. The Fubini-Study 2-form on
1146: $\P^{n+1}$ is
1147: \begin{equation}\label{e6}
1148: \omega^{\P ^{n+1}}_{FS} = \Bigg[
1149: \frac{\sum \mathrm{d}w_{a}\wedge
1150: \mathrm{d}\bar{w}_{a}}{1 + \sum
1151: w_{a}\bar{w}_{a}} -
1152: \frac{\big(
1153: \sum\bar{w}_{a}\mathrm{d}w_{a}\big)\wedge
1154: \big( \sum w_{a}\mathrm{d}\bar{w}_{a}\big)}
1155: {\big(1 + \sum w_{a}\bar{w}_{a} \big)^{2}}
1156: \Bigg].
1157: \end{equation}
1158: The pullback $\big( i_1^{\ast} \;
1159: \omega^{FS}_{\P ^{n+1}}\big)^n
1160: \in\Omega_0^{n,n}(X,\mathbb{Z})$ is a top
1161: form on $X$. Let $x_1,\ldots,x_n$ be local
1162: coordinates on $X$. Then
1163: \begin{equation}\label{f1}
1164: \big( i_1^{\ast} \; \omega^{FS}_{\P
1165: ^{n+1}}\big)_{i\bar\jmath}=
1166: \frac{\partial w_a}{\partial x_i} \; \big(
1167: \omega^{FS}_{\P ^{n+1}}\big)_{a\bar b} \;
1168: \overline{\frac{\partial w_{\bar b}}{\partial
1169: x_{\bar\jmath}}},
1170: \end{equation}
1171: and $\big( i_1^{\ast} \; \omega^{FS}_{\P
1172: ^{n+1}}\big)^n$ is proportional to the
1173: determinant of this matrix. For obvious
1174: reasons we need to evaluate this determinant.
1175: Let us outline how this can be done, paying
1176: attention to some of the numerical aspects.
1177:
1178: The idea is to choose local coordinates on
1179: $X$ that are convenient to work with. Let us
1180: start with the point $P$ on $X$ with
1181: homogenous coordinates $Z_i$. To minimize the
1182: numerical error we go to the affine patch
1183: where $|Z_i|$ is maximal.
1184: Without loss of generality let us assume that
1185: this happens for $i=0$. The affine coordinates are
1186: $w_{a}=Z_{a}/Z_{0}$.
1187:
1188: Let $p$ be the affine form of $f$, i.e.,
1189: $p(w)= f(1,w_{1}, w_{2},\ldots, w_{n+1})$.
1190: This equation determines one of the $w_{a}$-s
1191: in term of the others, as an implicit
1192: function. Let us assume for the sake of this
1193: presentation that $\partial p/ \partial
1194: w_{n+1}(P)\neq 0$. The implicit function
1195: theorem then tell us that in an open
1196: neighborhood of $P$ $w_{n+1}$ is a function
1197: of the remaining variables:
1198: $w_{n+1}=w_{n+1}(w_{1}, w_{2},\ldots,
1199: w_{n})$. This allows us to choose the
1200: coordinates $w_{1},\ldots, w_{n}$
1201: to be the local coordinates $x_1,\ldots,x_n$
1202: on $X$.
1203:
1204: This choice of coordinates is quite
1205: advantageous for computing \eqref{f1}. All we
1206: need is to compute $\partial w_{n+1}/\partial
1207: x_i$, as $\partial w_j/\partial x_i=\delta_{i
1208: j}$. This can be done algebraically, without
1209: explicitly solving the $p=0$ equation.
1210: Namely, using the fact that
1211: \[
1212: p( w_{1},\ldots, w_{n},w_{n+1}(w_{1},\ldots,
1213: w_{n}))\equiv 0
1214: \]
1215: is the identically zero function, its
1216: derivative with respect to any $w_{i}$
1217: vanishes identically, for $i=1,\ldots, n$. As
1218: a result we have that
1219: \begin{equation}\label{df}
1220: \frac{\partial w_{n+1}}{\partial w_i}(P)=
1221: -\frac{\partial p}{\partial
1222: w_i}(P)/\frac{\partial p}{\partial
1223: w_{n+1}}(P).
1224: \end{equation}
1225: For numerical stability one should always
1226: solve for the variable for which
1227: $|{\partial p}/{\partial w_i}(P)|$ is the largest.
1228:
1229: The second differential form entering
1230: \eqref{f2} is $\Omega\wedge
1231: \overline{\Omega}$. The holomorphic $n$-form
1232: $\Omega\in \Omega^{n,0}(X,\mathbb{C})$ can be
1233: represented using the Poincare residue map
1234: \cite[Section 1.1]{GH}
1235: \begin{equation}\label{omega}
1236: \Omega = (-1)^{i-1}
1237: \frac{\mathrm{d}w_{1}\wedge
1238: \mathrm{d}w_{2}\ldots \wedge
1239: \widehat{\mathrm{d}w_{i}}\wedge\ldots \wedge
1240: \mathrm{d}w_{n}}{\partial p(w)/\partial
1241: w_{i}}.
1242: \end{equation}
1243: where $\widehat{\mathrm{d}w_{i}}$ means the
1244: omission of $\mathrm{d}w_{i}$ in the wedge
1245: product.
1246:
1247: These explicit expressions allow us to perform integrals
1248: numerically on elliptic curves, $K3$ surfaces and
1249: more interestingly, quintic 3-folds.
1250:
1251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1252: %%%%%%%%%%%%
1253: \subsubsection{Symmetries}\label{s:sym}
1254:
1255: Suppose our Calabi-Yau $X$ is preserved as a complex manifold
1256: by the action of a discrete group $\Gamma$. Then a Ricci flat
1257: metric whose K\"ahler class $\omega$ is preserved by $\Gamma$ will
1258: also be $\Gamma$-invariant, because it is unique. As we will see in this section,
1259: the same statement applies to the balanced metrics as well.
1260:
1261: A general hermitian $N$ by $N$ matrix has $N^2$
1262: independent real coefficients. On the other hand, if the Calabi-Yau
1263: $X$ has discrete symmetries, then we expect to find symmetry
1264: relations between the matrix elements of $T(h)$.
1265: Taking advantage of these relations can drastically reduce the size of
1266: the problem.
1267: In this section, we argue that these symmetry relations are respected
1268: by the balanced metric and the T-map, and explain how we used them
1269: in the examples of Section~\ref{s:NR}.
1270:
1271: Next, let us review the symmetries of
1272: $X$ defined as a
1273: hypersurface in $\mathbb{P}^{n+1}$ by the
1274: degree $n+2$ homogenous polynomial
1275: \be\label{defX}
1276: f=\sum_{i=0}^{n+1}Z_{i}^{{n+2}}-(n+2)\psi
1277: \prod_{i=0}^{n+1}Z_{i}.
1278: \ee
1279: Here $\psi$ controls the complex structure of the hypersurface. Using the fact that $X$
1280: is
1281: Calabi-Yau, the symmetry group is
1282: finite.
1283: To find the symmetries of $X$ we consider two natural group actions
1284: on $\mathbb{P}^{n+1}$, and impose conditions such that these group actions descend to $X$.
1285:
1286: We start with the abelian group
1287: $$
1288: \bigoplus_{i=0}^{n+1}
1289: \mathbb{Z}_{p} \subset GL(n+2)
1290: $$
1291: that acts by independently rescaling the $n+2$ homogenous coordinates $Z_{i} \mapsto { \alpha_{i}}
1292: Z_{{i}}$, where $\alpha_{i}$ are $p$th roots of unity. Since the projective coordinates are defined only up to overall rescaling, we have to mod out by the diagonal action ${\triangle\mathbb{Z}_{p}} $ and find the group
1293: $$
1294: \bigoplus_{i=0}^{n+1}
1295: \mathbb{Z}_{p}/{\triangle\mathbb{Z}_{p}} \subset \mathbb{P}GL(n+2)
1296: $$
1297: acting on $\mathbb{P}^{{n+1}}$.
1298: In order for this group to descend to $X$,
1299: it must leave the defining equation \eqref{defX}
1300: invariant. In the Fermat case, that is
1301: $\psi=0$, we set $p=n+2$ and find that the
1302: Calabi-Yau is invariant under
1303: \be\label{abelian}
1304: \bigoplus_{i=0}^{n+1}
1305: \mathbb{Z}_{n+2}/{\triangle\mathbb{Z}_{n+2}}\cong
1306: \left( \mathbb{Z}_{n+2}\right)^{n+1}.
1307: \ee
1308: For non-vanishing $\psi$, the $\alpha_{i}$
1309: have to obey the additional constraint
1310: $
1311: \prod_{i=0}^{n+1}\alpha_{i}=1.
1312: $
1313: This shows that the symmetry group is a subgroup of
1314: $\mathbb{Z}_{n+2}/{\triangle\mathbb{Z}_{n+2}}\cong
1315: \left( \mathbb{Z}_{n+2}\right)^{n+1}$, given by the kernel of the
1316: product map $(\alpha_0,\ldots,\alpha_{n+1})\mapsto \prod_{i=0}^{n+1}\alpha_{i}$.
1317: We call this group $Ab_{n+2}$, and it is clear that there is an isomorphism
1318: $Ab_{n+2}\cong \left( \mathbb{Z}_{n+2}\right)^n$.
1319: For example, in the case of the torus defined by our
1320: cubic in $\mathbb{P}^{2}$,
1321: $Ab_{3}\cong\mathbb{Z}_{3}$. At the
1322: Fermat point this group is enhanced to
1323: $\mathbb{Z}_{3}^{2}$.
1324:
1325: The second symmetry group we consider is the
1326: symmetric group on $n+2 $ elements $\mathbb{S}_{{n+2}}$.
1327: This group acts by
1328: permuting the coordinates of
1329: $\mathbb{P}^{n+1}$. Since \eq{defX} is
1330: invariant under permutations, $\mathbb{S}_{n+2}$ is a symmetry
1331: of $X$ as well.
1332:
1333: To see how these actions on the coordinates
1334: of $\mathbb{P}^{n+1}$ induce an action on
1335: $\{s_{{\alpha}}\}$, the global sections of
1336: the line bundle ${\cal L}^{k}$ on $X$ defining the
1337: embedding in $\mathbb{P}^{N-1}$, we can use some simple algebraic geometry.
1338: The fact that $X$ is given by a hyperplane in
1339: $\mathbb{P}^{n+1}$ gives a natural way to
1340: parameterize the global sections of ${\cal
1341: L}^{k}=\O_X(k)$. We start with the short exact sequence (SES) defining $X$:
1342: \[
1343: \xymatrix@1{0 \ar[r] & \O_{\P^{n+1}}(-n-2) \ar[r]^-{\cdot f} & \O_{\P^{n+1}} \ar[r] & \O_X \ar[r] & 0}
1344: \]
1345: Tensoring with $\O_{\P^{n+1}}(k)$, and using the fact that $\H^1(\mathbb{P}^{n+1},\mathcal{O}_{\mathbb{P}^
1346: {n+1}}(k-n-2))=0$, for $k>0$, we get another SES:
1347: \be\label{Mike1}
1348: \xymatrix@1{0 \ar[r] &
1349: \H^{0}(\mathbb{P}^{n+1},\mathcal{O}_{\mathbb{P}^
1350: {n+1}}(k-n-2)) \ar[r]^-{\cdot f} &
1351: \H^{0}(\mathbb{P}^{n+1},\mathcal{O}_{\mathbb{P}^
1352: {n+1}}(k)) \ar[r] & \H^{0}(X,{\cal L}^{k}) \ar[r] & 0
1353: }
1354: \ee
1355: which shows that the global sections of
1356: $\H^{0}(X,{\cal L}^{k})$ can be parameterized
1357: by degree $k$ monomials in $n+2$
1358: variables modulo the ideal generated by $f$. Therefore
1359: the sections inherit an obvious group action.
1360:
1361: In particular, we also find that
1362: \[
1363: N=\dim \H^{0}(X,{\cal L}^{k})=
1364: \binom{n + k+1}{k}-\binom{k-1}{k-n-2}.
1365: \]
1366: In addition, note that the map $i_k\colon
1367: X\hookrightarrow{\mathbb{P}}^ {N-1}$
1368: factorizes
1369: \be\label{factor}
1370: \xymatrix{*++{X} \ar@{^{(}->}[r]^-i
1371: \ar@/_2pc/[rr]^-{i_k} & *++{\P^{n+1}} \ar@{^{(}->}[r]^-v & *++{\P^{N-1}} }
1372: \ee
1373: The second embedding, $v\colon\P^{n+1}
1374: \hookrightarrow{\mathbb{P}}^ {N-1}$, is the (Veronese) map associated to
1375: the incomplete linear system on $ \P^{n+1}$ induced by the complete
1376: linear system $|{\cal L}^{k}|$ on $X$.
1377:
1378: We will now consider the consequences of these actions on the $T$-map
1379: and the sequence of hermitian matrices $\{T^{l}(h)\}_{l=1,2,\ldots }$.
1380: First we consider the action of $Ab_{n+2}$. We assume that $T^{0}(h)$
1381: is invariant under the group action. This is a choice we can always
1382: make. Since $Ab_{n+2}$ is an abelian group, its irreducible
1383: representations are one dimensional, and can be labeled by the
1384: characters. Each section $s_{\alpha}$ transforms under a character
1385: $\chi_{{\alpha}}$. The operator $T$ is defined in terms of the
1386: sections, and the $Ab_{n+2}$ will force some of the
1387: $T(h)_{\alpha\bbeta}$ matrix elements to be zero. To better understand
1388: this we look at a toy example: the integral of an odd function $a $ on
1389: $\mathbb{R}$. The group $G$ in question is $\mathbb{Z}_2$, and acts on
1390: $\mathbb{R}$ by $x\mapsto -x$. $\mathbb{Z}_2$ has only one nontrivial
1391: representation, and being odd, $a $ transform in this irrep. Now we
1392: have that
1393: $$
1394: \int_{-\infty}^{+\infty}a(t)\,dt=
1395: \int_{+\infty}^{-\infty}a(-x)\,d(-x)=-\int_{
1396: -\infty}^{+\infty}a(x)\,dx,
1397: $$
1398: where we used a change of variable $t=-x$. This implies that
1399: $\int_{-\infty}^{+\infty}a(x)dx=0$.
1400:
1401: More generally, in $\mathbb{R}^n$ for a function $a$, a group $G$, and an element $g \in G$, we can do the change of coordinates $t=g\cdot x$ and then
1402: \begin{equation}\label{c1}
1403: \int_{X} a(t)\, dV(t) = \int_{g\cdot
1404: X} a(g\cdot x)g^{{*}}\, dV(x)=
1405: \int_{X} a(g\cdot x) \, dV(x)\,.
1406: \end{equation}
1407: Here we assumed the measure to be $G$-invariant.
1408:
1409: Applying \eq{c1} for $G= Ab_{n+2}$,
1410: and using the fact that $s_\alpha$ transform as a character of
1411: $G= Ab_{n+2}$, it gives that
1412: \begin{equation}
1413: T(h)_{\alpha\bbeta}
1414: =\frac{N}{\vol(X)}\int_X \frac{
1415: \chi_{{\alpha}}(u)s_\alpha
1416: \overline{\chi_{{{\bbeta}}}(u)} \bs_\bbeta}{||s||^2_h} \, d\mu_\Omega
1417: =\chi_{{\alpha}}(u)\overline{\chi_{\bar{\beta}}(u)}\, T(h)_{
1418: \alpha\bbeta}.
1419: \end{equation}
1420: We used the fact that $X$ is
1421: invariant under that action of $Ab_{n+2}$,
1422: and so is the measure $\Omega\wedge\bar{\Omega}$, and the
1423: denominator $||s||^2_h$. (The latter follows by induction from the
1424: initial choice of $ T^{0}(h) $ being invariant.) In particular, if
1425: $\chi_{{\alpha}}(u)\overline{\chi_{\bar{\beta}}(u)}\neq 1$, for any
1426: $u\in Ab_{n+2}$,
1427: then the corresponding $T(h)_{\alpha\bbeta}$ has
1428: to vanish.
1429: In our numerical routine we impose this
1430: vanishing condition on all the matrices
1431: $T^{l}(h)$.
1432:
1433: A similar argument applies for $G= \mathbb{S}_{n+2}$.
1434: Since $\mathbb{S}_{n+2}$ is not
1435: abelian, and hence its generic irreducible
1436: representations are not one dimensional, this
1437: constraint does not result in vanishing rules, but rather sets a priori
1438: independent coefficients of $T(h)$ equal to each other. To see how
1439: $\mathbb{S}_{n+2}$ acts, recall from \eq{Mike1} that
1440: $$ \H^{0}(X,{\cal L})\cong
1441: \H^{0}(\mathbb{P}^{n+1},\mathcal{O}_{\mathbb{P}^
1442: {n+1}}(1))=\C^{n+2}$$
1443: is the fundamental representation of $\mathbb{S}_{n+2}$, call it $F$. Then
1444: $\H^{0}(\mathbb{P}^{n+1},\mathcal{O}_{\mathbb{P}^
1445: {n+1}}(k))$ is the $k$th symmetric tensor power of $F$,
1446: $Sym^k\,F$, and by \eqref{Mike1} $\H^{0}(X,{\cal L}^{k})$
1447: is a quotient of this. Now we can return to \eq{c1}.
1448: Once again,
1449: we choose $T^{0}(h)$ to be invariant under $\mathbb{S}_{n+2}$, and then
1450: induction and \eq{c1} tell us which matrix elements of $T(h)$
1451: equal each other.
1452:
1453: Therefore, imposing the symmetries of both
1454: finite groups, the number of independent components of
1455: $T^{l}(h)$ (for any $l$) reduces significantly. To illustrate this we consider
1456: $k=12$ on the quintic in $\P^4$, i.e., $n=3$ (this was the largest $k$ we computed).
1457: In this case $N=1490$. This means that $T(h)$ is a hermitian matrix with
1458: 2,220,100 components. Taking into account the $Ab_5$ and $\mathbb{S}_5$
1459: relations, one is reduced to computing 9800 components. This simplification
1460: speaks for itself.
1461:
1462: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1463: %%%%%%%%%%%%
1464: \section{Numerical results}\label{s:NR}
1465:
1466: In this section we present our explicit numerical results.
1467: The main object that we compute is the balanced metric
1468: associated to the embedding of the quintic threefold
1469: defined in \eq{defX}. For definiteness we chose to work with $\psi=0.1$, but also tested other values of $\psi$. We also considered the case of elliptic curves ($n=1$) and K3 surfaces ($n=2$). In all these cases we obtained results similar to the ones to be presented here.
1470:
1471: To find the balanced metric we study the associated $N\times N$ matrix
1472: $h_k$ for several values of $k$, from $k=1$ to $k=12$.
1473: We use $h_k$ to construct the associated
1474: K\"ahler form $\omega_k$ on $X$, and check how well it approximates
1475: the Ricci flat metric. We do this in several ways.
1476:
1477: First, one can study the function defined in \eq{define-eta}
1478: $$
1479: \eta_k = \frac{\det\,\omega_k}{\Omega\wedge\bar{\Omega}}\colon X\longrightarrow \R.
1480: $$
1481: For a good approximation $g_k$ to the Ricci flat metric
1482: $g_{RF}$ the function $\eta_k$ is almost constant.
1483: We study the behavior of $\eta_k$ statistically, by summing over
1484: all the regions of $X$, and also locally paying attention to certain
1485: special regions of the threefold.
1486:
1487: Second, one can compute the Ricci tensor of $\omega_k$.
1488: To check pointwise how close to zero the Ricci tensor is, we need a diffeomorphism invariant
1489: quantity. We chose to work with the Ricci scalar. We also perform this
1490: analysis for several values of $k$, and show how the Ricci
1491: scalars decrease pointwise with $k$.
1492:
1493: Before presenting the results let us comment on the errors coming from
1494: Monte Carlo integration. We estimate them by computing the balanced
1495: metrics associated to different samples of points, and then looking at the mean and variance
1496: of each individual matrix element.
1497: Ideally, one would like to produce samples of points with minimal induced error.
1498: Constructions that reduce the standard deviation of the integrals
1499: are refinements to the theory of numerical
1500: integration presented here.
1501: Markov Chain Monte Carlo techniques, construction
1502: of lattices on Calabi-Yau varieties, and of quasi-random points on such manifolds are
1503: different approaches that one could consider.
1504:
1505: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1506: \subsection{Approximating volumes v.s. \CY\ volume}
1507:
1508: Here, we consider the way the function
1509: $$
1510: \vert \eta_{k}-1_X\vert\colon X\longrightarrow \R_+,\qquad x\mapsto \vert\eta_{k}(x)-1 \vert
1511: $$
1512: behaves on $X$. As argued earlier, we expect $\vert \eta_{k}-1_X\vert$ to approach the constant zero function.
1513: One can study the deviation of
1514: $\vert \eta_{k}-1_X\vert$ from the zero function
1515: by computing the integral
1516: \begin{equation}\label{Aeta}
1517: \sigma_k = \int_{X} \vert \eta_{k}-1_X\vert\, d\mu_{\Omega}\,.
1518: \end{equation}
1519: We compute this integral by our Monte Carlo method, which introduces an error, and this error can be estimated by
1520: \begin{equation}\label{SIGeta}
1521: \delta \sigma_k = \frac{1}{\sqrt{N_p}}\left(
1522: \int_{X} \left(\vert \eta_{k}-1_X\vert-\sigma\right)^{2}\, d\mu_{\Omega}
1523: \right)^{1/2},
1524: \end{equation}
1525: where $N_p$ is the number of points used to perform the Monte Carlo
1526: integration in \eqref{Aeta}.
1527:
1528: \begin{figure}[h]
1529: \begin{center}
1530: \begin{tabular}{cc}
1531: \includegraphics[angle=0,width=3in]{Eta.ps}&
1532: \includegraphics[angle=0,width=3in]{RicciScalar.ps}
1533: \end{tabular}
1534: \end{center}
1535: \caption{$\sigma_k$ and Ricci scalars.} \label{f:eta}
1536: \end{figure}
1537:
1538: In Fig. \ref{f:eta} we plot the values $\sigma_k$ defined in \eqref{Aeta}
1539: for $k=3,\ldots, 12$. The error bars for each value are the
1540: corresponding standard deviations \eqref{SIGeta}.
1541: We also see how the errors decrease, along with $\sigma_k$, for higher and higher $k$.
1542: The fit in Fig. \ref{f:eta} is a curve of type
1543: $$
1544: \sigma_k = \frac{\alpha}{k^2} +\frac{\beta}{k^3} +\mathrm{O}\left(\frac{1}{k^4}\right),
1545: $$
1546: as we expect from the theory.
1547:
1548: We can also study the local behavior of $\eta_k$
1549: by restricting it to a subspace. Given our quintic 3-fold, we consider the rational curve defined by
1550: \begin{equation}
1551: \label{ratcuv}
1552: ( Z_0 = z_0,\, Z_1 = -z_0,\,Z_2 = z_1,\,Z_3 =
1553: 0,\,Z_4 = -z_1 ),
1554: \end{equation}
1555: where $Z_{i}$ are homogeneous
1556: coordinates on $\mathbb{P}^4$, while
1557: $(z_{0},\, z_1)$ are homogeneous coordinates
1558: for $\mathbb{P}^1$.
1559: This rational curve lies on every quintic
1560: defined by \eq{defX}.
1561:
1562: \begin{figure}[h]
1563: \begin{center}
1564: \begin{tabular}{cccc}
1565: \includegraphics[width=1.2in]{1.eps}&
1566: %\includegraphics[width=1.2in]{2.eps}&
1567: \includegraphics[width=1.2in]{3.eps}&
1568: \includegraphics[width=1.2in]{4.eps}&
1569: \includegraphics[width=1.2in]{5.eps}\\ \\
1570: %\includegraphics[width=1.2in]{6.eps}&
1571: \includegraphics[width=1.2in]{7.eps}&
1572: %\includegraphics[width=1.2in]{8.eps}\\ \\
1573: \includegraphics[width=1.2in]{9.eps}&
1574: %\includegraphics[width=1.2in]{10.eps}&
1575: % \includegraphics[width=\textwidth , bb= 20 20 575 575]{/home/karp/p/ss/11.jpg}&
1576: \includegraphics[width=1.2in]{11.eps}&
1577: \includegraphics[width=1.2in]{12.eps}
1578: \end{tabular}
1579: \end{center}
1580: \caption{The values of $\eta$ on the rational curve, for
1581: $k=1$, 3, 4, 5, 7, 9, 11 and 12.}
1582: \label{plots}
1583: \end{figure}
1584:
1585: In Fig. \ref{plots} we plot the values of function $\eta_{k}$ restricted to
1586: the rational curve defined above for 12
1587: different values of $k$,
1588: ranging between $1$ and $12$.
1589: More concretely, given the embedding \eqref{ratcuv}, we choose
1590: the local coordinate system on $\mathbb{P}^1$
1591: defined by
1592: $t=z_1/z_0$, and take the stereographic
1593: projection of
1594: the $t$-plane. Using spherical coordinates $(\theta,\, \phi)$
1595: on
1596: $\P^1\equiv S^2$ we embed it into $ \mathbb{R}^3$,
1597: by the parameterization
1598: $$
1599: z_0 = \sin\theta\,\cos\phi,\qquad
1600: z_1 = \sin\theta\,\sin\phi + i\,\cos\theta.
1601: $$
1602: In the radial direction of $\mathbb{R}^3$ we
1603: plot the function $\eta_{k}$. As expected, $\eta_k$ approaches
1604: the constant function 1 as $k$ increases.
1605:
1606: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1607: \subsection{Ricci scalars}
1608:
1609: Next we discuss how we compute the
1610: Ricci scalar on $X$.
1611: The K\"ahler potential on $X$ is given by the
1612: restriction of the K\"ahler potential
1613: \eqref{Ks}, which is associated to a balanced metric. The metric and Ricci curvature on
1614: $X$ are given by
1615: \begin{equation}\label{r4}
1616: g_h=\partial\bar{\partial} K_{h}
1617: ,\qquad
1618: Ric =\partial\bar{\partial} \det{g_h}.
1619: \end{equation}
1620:
1621: To compute these quantities we
1622: need the first and second derivatives of the sections
1623: $s_{\alpha}$ with respect to the
1624: coordinates of $X$.
1625: We can compute these derivatives algebraically,
1626: using the ideas outlined in Section~\ref{s:nM}.
1627: To get the actual value of
1628: $g$ and $Ric$, at a specific point $x \in X$, we
1629: evaluate our numerical expressions at
1630: this point. Let us stress that this way $g$ and $Ric$
1631: are evaluated algebraically, without numerical derivatives.
1632:
1633: Let us now discuss the effect of the parameter $k$ on the Ricci scalar $R$. The theoretical estimate predicts a vanishing of $R$ for large $k$ as
1634: \be\label{vanR}
1635: ||R_k|| < \frac{\gamma}{k} + \frac{\delta}{k^2}+\cdots,
1636: \ee
1637: where $\gamma $ and $\delta$ are constants, in any $C^r$ norm. We can see this pointwise, after a short statistical analysis. Let us pick $100$ randomly chosen points $P_i$ on the 3-fold, and compute the associated sets of Ricci scalars $R_{k}$, for $k \in [3,\dots, 10]$. To obtain a normalized set $\tilde{R}_{k}$ of Ricci scalars we rescale all $R_{k} $ by $1/R_{3}$, and we do this for every point. This normalization leads to a more meaningful comparison between the different points.
1638:
1639: We are only interested in points where $\tilde{R}_{k}$ shows a generic behaviour. For this we compute the expectation value of $|\tilde{R}_{10}|$ ($k=10$ gives a good accuracy)
1640: $$
1641: \langle |\tilde{R}_{10}| \rangle = \frac{1}{100}\sum_{i=1}^{100} |\tilde{R}_{{10}}(P_i)|=0.48.
1642: $$
1643: We consider a point $P$ generic if it lies at a distance of order one from the mean $\langle |\tilde{R}_{10}| \rangle$, that is
1644: $$
1645: |\tilde{R}_{{10}}(P)-\langle |\tilde{R}_{10}| \rangle |<0.5.
1646: $$
1647: We find that $95$ points out of $100$ obey this condition. We use these $95$ points for our statistical check of \eq{vanR}.
1648: In Fig.~\ref{f:eta} we plot $ \langle |\tilde{R}_{k}| \rangle$ as a function of $k$. We do a least square fit for both $\gamma $ and $\delta$, and obtain very good agreement.
1649:
1650: \begin{figure}[h]
1651: \begin{center}
1652: \begin{tabular}{cc}
1653: \includegraphics[angle=0,width=3in]{histo.eps}&
1654: \includegraphics[angle=0,width=3in]{Scatter.ps}
1655: \end{tabular}
1656: \end{center}
1657: \caption{Histogram of Ricci scalars, and scattered plot of mass versus the Ricci scalars.}
1658: \label{f:Histogram}
1659: \end{figure}
1660:
1661: We can obtain a visual picture of how the Ricci scalars decrease with $k$ by plotting them on a histogram. We use the same $100$ points from above. To keep the picture simple we plot only six $k$ values: 3, 4, 5, 7, 9 and 12. Unlike earlier, here we do not normalize the Ricci scalars pointwise, instead we reorder them decreasingly. This ordering is done for no other reason but to enhance the visual clarity. Accordingly, in the first graph of Fig.~\ref{f:Histogram} we plotted the value of the Ricci scalar for every point (the $y$ axis is the absolute value of Ricci scalar, while on the $x$ axis we have the points from 1 to 100). We did this for the $k$ values indicated above, and used different colors to distinguish them (e.g., yellow corresponds to $k=3$, while $k=12$ is blue). It is evident from this graph that by going from $k=3$ to $k=12$ the Ricci scalars decrease by an order of magnitude, in line with the theoretical expectation.
1662:
1663: One observes that for any $k$ there are a few anomalously large Ricci scalars. To understand this we can do a scattered plot of the mass of that point versus the Ricci scalar in question (we present this for $k=12$), using $1000$ random points. This is the second graph in Fig.~\ref{f:Histogram}. The picture shows a correlation between large values of the Ricci scalar and large values of the mass (large in a logarithmic sense). In other words, in regions where our point generator needs large correction, via the mass, the balanced metric is a less accurate approximation of the Ricci flat metric, compared to points with smaller mass. This fact is then amplified by the formula for the Ricci tensor in \eq{r4}, where the logarithmic scale is also supplied.
1664:
1665: Finally let us note that $1/\eta_1$ is
1666: %up to a global constant
1667: precisely the mass function \eqref{e33}. This is because the $k=1$ balanced matrix is proportional to the identity, a consequence of the discrete symmetries present for our quintic. This leads to an alternative interpretation for the first graph ($k=1$) in Fig.~\ref{plots}, as depicting the inverse masses of the points on that rational curve.
1668:
1669: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1670: \subsection{Discussion}
1671:
1672: We will discuss further applications of these results elsewhere; here we
1673: discuss the advantages and limitations of this approach compared to
1674: others, for example position space methods \cite{Headrick:Wiseman}.
1675:
1676: The runtime of a computation of the balanced metric can be approximated
1677: as
1678: $$
1679: T = N_{it} \times N_p \times S^2 ,
1680: $$
1681: where $S$ is the number of independent sections
1682: (taking into account discrete symmetry),
1683: $N_p$ is the number of points used in the Monte Carlo integration,
1684: and $N_{it}$ is the number of iterations of the T-map
1685: required for convergence. Since convergence is exponential,
1686: this leads to a rough scaling with the accuracy as
1687: $$
1688: T \sim \frac{\log \epsilon}{\epsilon^2} S^2 .
1689: $$
1690:
1691: The value of $S$ required for a given accuracy depends on the
1692: symmetries and dimension. For the balanced metrics, we expect
1693: to need $k\sim 1/\sqrt{\epsilon}$; as discussed in section 2
1694: this could probably be improved by choosing a different scheme
1695: if accuracy were paramount. For hypersurfaces in $n$ complex dimensions,
1696: we then have $S \sim N \sim k^{n+1}$, leading to a rough overall scaling
1697: of $T \sim 1/\epsilon^{n+3}$. This might be compared with a (naive)
1698: $T \sim 1/\epsilon^{2n}$ for position space methods, so the two appear
1699: generally competitive. However, along with the other
1700: advantages we mentioned, we believe the approach we are discussing is
1701: far easier to program, and requires relatively little effort to adapt
1702: to different manifolds, and related problems such as hermitian Yang-Mills.
1703:
1704: Since the sections $s_\alpha$ of $\mathcal{O}_{X}(k)$ are degree $k$
1705: polynomials, this basis is a simple type of Fourier or momentum space
1706: basis. Very roughly speaking, a degree $k$ basis should be able to
1707: represent arbitrary structures on length scales down to $1/k$. They
1708: are particularly well suited for approximating smooth functions, as
1709: the Fourier coefficients of such a function fall off faster than any
1710: power of $k$ (see the appendix of \cite{Donaldson:numeric} for more
1711: precise statements). This is advantageous as the Ricci flat metric is
1712: smooth, suggesting that other approximation schemes could do better
1713: than $\epsilon\sim 1/k^2$.
1714:
1715: On the other hand, in some limits (say a conifold limit) the metric
1716: can develop structure on small scales, which might not be well
1717: represented by a fixed $k$ basis. This is also a problem for position
1718: space methods with a fixed lattice; there one deals with it by multi-scale
1719: methods, for example allowing the lattice spacing and structure to vary
1720: over the manifold. This is very powerful but also very intricate to
1721: program. In the present context, rather than increase $k$, one might
1722: look for analogous simplifications; either a multi-scale method which uses
1723: different $k$ in different regions (or even some sort of wavelet-inspired
1724: method). Or, since we have many explicit expressions for Ricci flat
1725: metrics near singularities, it might be useful to develop
1726: a way to patch these solutions into the global approximate solutions
1727: we discussed.
1728:
1729: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1730: \subsubsection*{On the computer code}
1731:
1732: Our numerics is based on code that has been written entirely in C++. Our experience shows that these computations must be done in a compiled language, rather than an interpreted one. We have made extensive use of the following Boost libraries: uBlas, random, bind and thread. These libraries are on par with Fortran code, due to implementation techniques using expression templates and template metaprograms. The computations were done on an Athlon 64 4800+ dual core machine, with 4GB memory. The computational time ranges from minutes, for low $k$, to hours, and eventually 2 days (for $k=12$).
1733:
1734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1735: \subsubsection*{Acknowledgments}
1736:
1737: This research was supported in part by the
1738: DOE grant DE-FG02-96ER40949.
1739:
1740: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1741: %%%%%%%%%%%%%%%%
1742:
1743: %\bibliographystyle{/home/karp/lat/physics}
1744: \bibliographystyle{/home/karp/lat/utcaps}
1745: %\bibliography{/home/karp/lat/BIB}
1746: \input p.bbl
1747:
1748: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1749: %%%%%%%%%%%%%%%%
1750: \end{document}
1751: