hep-th0507090/poly.tex
1: \documentclass[12pt,titlepage]{utarticle}
2: \usepackage{amssymb,color}
3: %\usepackage{dcpic,pictexwd}
4: %\usepackage{graphicx}
5: %\usepackage[rflt]{floatflt}
6: \usepackage{hyperref}
7: 
8: \numberwithin{equation}{section}
9: 
10: \def\defeq{\buildrel\rm def\over=}
11: 
12: \def\eg{\mathfrak{g}}
13: \def\del{\partial}
14: \def\delb{{\bar{\partial}}}
15: \def\wdg{{\wedge}}                              % wedge product
16: \newcommand{\vvev}[1]{{\langle\kern-.5ex\langle #1\rangle\kern-.5ex\rangle}}
17: \newcommand{\px}[1]{{\partial_{#1}}}
18: \newcommand{\qx}[1]{{\partial^{#1}}}
19: \newcommand{\pxpy}[2]{\frac{\partial{#1}}{\partial{#2}}}
20: 
21: \newcommand{\rep}[1]{{{\bf {#1}}}}      % representation
22: \newcommand{\tr}[1]{{\mbox{tr}\left({#1}\right)}}          % trace
23: 
24: \newcommand{\mf}[1]{{{\bf {#1}}}}
25: \def\Z{\mathbb{Z}}
26: \def\C{\mathbb{C}}
27: \def\R{\mathbb{R}}
28: \def\Q{\mathbb{Q}}
29: \def\I{\mathbf{I}}
30: %\def\BT{\mathbb{T}}
31: \def\BT{U(1)}
32: \def\MB{\mathbf{B}}
33: \def\ME{{\mathbf{E}}}
34: \def\Sp{{\bf S}^{\bf 1}_p}
35: \newcommand{\MR}[1]{{\mathbb{R}^{#1}}}            % Real numbers
36: \newcommand{\MC}[1]{{\mathbb{C}^{#1}}}            % Complex numbers
37: \newcommand{\MS}[1]{{{\bf S}^{#1}}}               % Circle, sphere,...
38: \newcommand{\MT}[1]{{{\bf T}^{#1}}}               % Torus
39: \renewcommand{\CP}[1]{{\mathbb{C}{\mathbf{P}^{#1}}}}        % CP
40: \renewcommand{\arraystretch}{1.25}
41: 
42: \def\Hil{\mathcal{H}}             % Hilbert space
43: \def\mcO{\mathcal{O}}
44: \def\mcA{\mathcal{A}}
45: \def\mcM{\mathcal{M}}
46: \def\Ad{\mathrm{Ad}}
47: \def\Tr{\mathrm{Tr}}
48: %\def\tr{\mathrm{tr}}
49: \def\Ot{\widehat{\Omega}}
50: \def\N{\mathcal{N}}
51: \def\nn{\nonumber}
52: \def\ie{\textit{i.e.,\ }}
53: \def\<{\langle}
54: \def\>{\rangle}
55: \def\P{{\Phi}}
56: \def\Pb{{\bar{\Phi}}}
57: \def\Db{{\bar{D}}}
58: \def\Wb{{\bar{W}}}
59: \def\Ab{{\bar{A}}}
60: \def\Jb{{\bar{\mathbf{J}}}}
61: \def\Ib{{\bar{\mathbf{I}}}}
62: \def\Wav{{W_{\mbox{av}}}}
63: \def\ib{{\bar{\imath}}}
64: \def\jb{{\bar{\jmath}}}
65: \def\kb{{\bar{k}}}
66: \def\lb{{\bar{l}}}
67: \def\nb{{\bar{n}}}
68: \def\a{{\alpha}}
69: %\def\b{{\beta}}
70: %\def\g{{\gamma}}
71: \def\d{{\delta}}
72: \def\e{{\epsilon}}
73: \def\p{{\phi}}
74: %\def\u{{\mu}}
75: %\def\v{{\nu}}
76: %\def\s{{\sigma}}
77: %\def\S{{\Sigma}}
78: %\def\t{{\tau}}
79: %%\def\h{{\eta}}
80: %\def\w{{\omega}}
81: %\def\x{{\xi}}
82: %\def\z{{\zeta}}
83: %\def\th{{\theta}}
84: %\def\Th{{\Theta}}
85: \def\L{{\mathcal{L}}}
86: %\def\inf{{\infty}}
87: \def\vev#1{\langle #1\rangle}
88: 
89: \begin{document}
90: \preprint{UTTG--07--05\\
91: \texttt{hep-th/0507090}\\}
92: 
93: \title{Random Polynomials and the Friendly Landscape}
94: 
95: \author{Jacques Distler and Uday Varadarajan}
96: 
97: \oneaddress{Theory Group, Physics Department\\
98:              University of Texas at Austin\\
99:              Austin, TX 78712\\ {~}\\
100:              \email{distler@golem.ph.utexas.edu}
101:              \email{udayv@physics.utexas.edu}}
102: 
103: \Abstract{  
104:   In hep-th/0501082, a field theoretic ``toy model'' for the Landscape
105:   was proposed. We show that the considerations of that paper carry
106:   through to realistic effective Lagrangians, such as those that
107:   emerge out of string theory. Extracting the physics of the large
108:   number of metastable vacua that ensue requires somewhat more
109:   sophisticated algebro-geometric techniques, which we review.}
110: 
111: \maketitle
112: \newpage
113: 
114: \section{Introduction and Summary}\label{sec:intro}
115: 
116: One of the striking observations of KKLT \cite{Kachru:2003aw} was that
117: the existence of a large number of long-lived metastable vacua in
118: certain compactifications of string theory (dubbed the ``landscape''
119: \cite{Susskind:2003kw}) provides a concrete instance in which the
120: anthropic principle \cite{Weinberg:1987dv, Brown:1987dd,
121:   Weinberg:1988cp, Martel:1997vi, Garriga:1999bf, Weinberg:2000qm,
122:   Bousso:2000xa} might be realized in Nature.  Unfortunately, many
123: discussions of these ideas get bogged down because it is hard to
124: disentangle the intricacies of the string theoretic constructions from
125: the ``anthropic'' questions one would like to address.
126: 
127: The paper of Arkani-Hamed, Dimopoulos and Kachru
128: \cite{Arkani-Hamed:2005yv} was, therefore, very useful in clearing
129: away the string-theoretic underbrush and presenting a simple,
130: tractable field-theoretic model with a large number of vacua in which
131: anthropic questions could be addressed (see also
132: \cite{Dienes:2004pi}).
133: 
134: One of the concepts to emerge from their investigation was the notion
135: of a ``friendly landscape.''  In general, all of the couplings,
136: $c_{a}$ of the theory will vary between the different vacua. However,
137: there is a qualitative difference between those couplings which
138: ``scan'' (those whose standard deviation is much larger than their
139: mean value) and those couplings which ``don't scan'' (those which are
140: sharply-peaked about their mean value). In making anthropic arguments,
141: one usually considers the situation in which one coupling is allowed to
142: vary, while the others are held fixed. If all the couplings vary
143: appreciably, the anthropic bounds are much weaker, or go away
144: entirely. So it was very useful for the authors of
145: \cite{Arkani-Hamed:2005yv} to identify a \emph{mechanism} by which the
146: couplings one would like to ``tune'' anthropically \emph{scan},
147: whereas the remaining couplings are sharply peaked.
148: 
149: Their model had three basic features:
150: \begin{itemize}
151: \item[1)]A large number, $N$, of scalar fields, $\phi_{i}$.
152: \item[2)]A decoupled  form for the scalar potential,
153:   $V(\phi)=\sum_{i}V_{i}(\phi_{i})$.
154: \item[3)]A decoupled form for the $\phi$-dependence of the observable
155:   couplings of the model, $c_{a}(\phi)=\sum_{i}c_{ai}(\phi_{i})$.
156: \end{itemize}
157: The first has a fairly natural realization in string theory. Many
158: compactification of string theory have hundreds of moduli which, when
159: the physics which lifts the vacuum degeneracy is included \cite{
160:   Giddings:2001yu, Kachru:2002he, Acharya:2002kv, Acharya:2003gb,
161:   Giryavets:2003vd, Denef:2004dm, Denef:2005mm,
162:   Balasubramanian:2004uy, Balasubramanian:2005zx, Conlon:2005ki,
163:   DeWolfe:2005uu}, form natural candidates for the $\phi_{i}$. We can,
164: quite plausibly, take ``$N$'' to be the number of (complex) moduli.
165: However, 2,3) are rather unnatural from this point of view and,
166: indeed, it's hard to imagine that such a decoupled structure might
167: emerge from string theory.
168: 
169: In the present work, we would like to overcome this drawback and
170: present what we hope is a realistic version of the scenario of
171: \cite{Arkani-Hamed:2005yv}. Our approach will be to start with,
172: essentially, the most general $\N=1$ supersymmetric effective field
173: theory with a large number of ``moduli'' chiral multiplets,
174: $\Phi^{i}$. We will then see what conditions must be imposed in order
175: to realize a friendly landscape.
176: 
177: In \S\ref{sec:sectors}, we will discuss the theory of $N$ chiral
178: multiplets coupled to $\N=1$ supergravity. Whereas any such theory
179: must be cut off at a scale $M_{c}< M_{p}$, we will see that, at
180: large-$N$, we will will need to impose the stronger condition $M_{c}<
181: M_{p}/\sqrt{N}$, in order to have a sensible effective field theory.
182: Moreover, we will find, in \S\ref{subsec:local}, that the couplings
183: among the moduli chiral multiplets must be suitably small; we will
184: summarize the condition that we must impose on these couplings by
185: saying that they are \emph{generically small}. We then turn to a
186: polynomial truncation of the superpotential of the model. While not
187: indispensable, such a truncation makes the analysis of the vacuum
188: structure amenable to perturbation theory. As usual, for this to be a
189: \emph{reliable} guide to the vacuum structure, the super-renormalizable
190: terms in the superpotential must have coefficients governed by a mass
191: scale $M_r \ll M_{c}$ (\S\ref{subsec:VacStruct}). In \S\ref{subsec:Flux},
192: we will see how these considerations mesh with the most popular arena
193: for landscape considerations --- F-theory vacua with fluxes. 
194: 
195: Next (\S\ref{subsec:R}), we impose the discrete R-symmetry found by
196: \cite{Arkani-Hamed:2005yv} to lead to a friendly landscape for the
197: cosmological constant, and use the $GL(N,\BC)$ symmetry of field
198: redefinitions to simplify our problem (\S\ref{subsec:Fermat}).
199: 
200: In \S\ref{sec:Solve}, we lay out the algebro-geometric techniques used
201: to determine the vacuum structure and extract information about the
202: distribution of values for various physical quantities among the
203: $2^{N}$ vacua of the theory. In \S\ref{sec:Stats}, we use these
204: techniques to compute certain ``holomorphic moments'' which
205: characterize the distribution of values of the cosmological constant.
206: In \S\ref{sec:Scanning}, we discuss the physics that ensues when one
207: assumes that the couplings are chosen from some (unspecified)
208: probability distribution, generalizing the considerations of
209: \cite{Arkani-Hamed:2005yv}. We apply our analysis both to the
210: superpotential and to other holomorphic couplings.
211: 
212: Finally, in \S\ref{sec:Generalizations}, we discuss some
213: generalizations of our techniques and future directions.
214: 
215: \section{General Features of SUSY Landscape Sectors}\label{sec:sectors} 
216: 
217: As a field theoretic model for the landscape sector, we could start
218: with an arbitrary SUSY field theory of $N$ chiral superfields and $N'$
219: vector superfields, all coupled to $\N=1$ SUGRA. For simplicity, we
220: will restrict ourselves to the case $N'=0$ and only briefly touch on
221: generalizations to gauged hidden sectors in this work. Thus, the
222: vacuum structure of the model can be determined using the
223: two-derivative effective action of the SUSY non-linear $\sigma$-model
224: describing the $N$ chiral fields at energies below a cutoff scale,
225: $M_c$.
226: 
227: Clearly, for any sort of effective field theory to be valid, we must
228: take $M_{c}<M_{p}$. However, as noted in \cite{Arkani-Hamed:2005yv},
229: when one has a large number, $N\gg 1$, of fields, radiative stability
230: of Newton's constant requires
231: \begin{equation}
232:    \frac{M_{c}^{2}}{M_{p}^{2}}< \frac{1}{N}
233: \end{equation}
234: To see this, note that the action for $\N=1$ supergravity
235: interacting with $N$ chiral fields is, in superspace notation (we
236: use $M_p$ for the reduced Planck mass and the conventions of
237: \cite{Gates:1983nr}),
238: \begin{equation}
239:   \label{eq:SUGRAaction}
240:   S =  -3M_p^2 \int d^8 z E^{-1} e^{- \frac{1}{3M_p^2}K(\P,\Pb)}
241:     + \int d^6 z \phi^3 W(\P) + \mathrm{h.c.}
242: \end{equation}
243: where $E^{-1}$ is the superdeterminant of the vielbein, $z$ is a
244: superspace coordinate, and $\phi$ is a compensator superfield. The
245: Einstein-Hilbert term comes from the leading ($\P$-independent) piece
246: of the first term in \eqref{eq:SUGRAaction}. At one-loop, this
247: receives a quadratically-divergent contribution,
248: \begin{equation}
249:  \left. \Delta K_W^{(1)}\right|_{\P=\Pb=0} \sim
250:     \left(\frac{M_c^2}{16\pi^2} \right) g^{i \jb} \del_i \delb_\jb K
251:     \sim N \left(\frac{M_c^2}{16\pi^2} \right) 
252: \end{equation}
253: where $g_{i\overline{\jmath}}= \del_{i}\delb_{\jb} K$ is the K\"ahler
254: metric of the $\sigma$-model.  The enhancement which comes from having
255: $N$ fields running around the loop requires us to set the cutoff,
256: $M_{c}$ to be parametrically smaller than $M_{p}$.
257: 
258: In $\N=1$ supergravity, the chiral multiplets parameterize a K\"ahler
259: manifold, whose K\"ahler form, $\omega = \tfrac{i}{2}
260: \partial\overline{\partial} K$. The superpotential, $W(\P)$ transforms
261: as a section of a line bundle, $\CL$, whose first Chern class,
262: \begin{equation}
263: c_1(\L) = -\frac{1}{\pi M_{p}^{2}} \omega = \frac{1}{2\pi i M_p^2} \del \delb K.
264: \end{equation}
265: The fiber metric on $\CL$ is
266: \begin{equation}
267: h(\P,\Pb)=e^{K(\P,\Pb)/M_p^2}
268: \end{equation}
269: and the connection is of type (1,0), $D_i = \del_i + \del_i K/M_p^2$.
270: Under K\"ahler transformations, $W(\P)\to e^{f(\P)/M_{p}^{2}} W(\P)$,
271: \begin{equation}
272:   \label{eq:Kpot}
273: K(\P^i,\Pb^\ib) \rightarrow K(\P^i,\Pb^\ib) + f(\P^i) +
274: \bar{f}(\Pb^\ib),
275: \end{equation}
276: The supersymmetric vacua of this model are the critical points of the
277: superpotential with respect to the Chern connection, i.e. the points
278: at which $D_i W \in \Gamma(T^*X \otimes \L)$ intersect the zero
279: section of $T^*X \otimes \L$,
280: \begin{equation}
281:   \label{eq:SUSYvac}
282: D_i W = \del_i W + (\del_i K) W /M_p^2=0.
283: \end{equation}
284: More generally, all the vacua of this model, supersymmetric or not,
285: are critical points of its scalar potential,
286: \begin{equation}
287:   \label{eq:scalarpot}
288:   V = e^{K/M_p^2} \left( g^{i \jb}  D_i W \Db_\jb \Wb - 3 |W|^2/M_p^2 \right).
289: \end{equation}
290: Since we will be interested in vacua in which $|W|\sim M_{r}^{3} <
291: M_{c}^{3} \ll M_{p}^{3}$, we will always be safe in neglecting the
292: connection term in \eqref{eq:SUSYvac} and hunting for ordinary
293: critical points, $\partial_{i}W=0$.
294: 
295: In fact, to justify a perturbative analysis of the effective
296: Lagrangian \eqref{eq:SUGRAaction}, we require that $W(\P)$ have a
297: convergent Taylor expansion in a polydisk, $|\P_{i}|< M_{c}$. Just as
298: the radiative stability of Newton's constant imposed constraints on
299: the cutoff scale, $M_{c}$ of our effective Lagrangian, radiative
300: corrections to the K\"ahler metric imposes, at large-N, further
301: constraints on the coefficients.
302: 
303: \subsection{Quantum Corrections to the K\"{a}hler
304:   Potential and Large $N$ Scaling}\label{subsec:local}
305: 
306: Before embarking upon a study of the vacua of these models, we should
307: study the radiative stability of these models in the limit of large
308: $N$. As our theory is an effective field theory, we expect that the
309: characteristic radius of curvature for $X$ is given by the cutoff
310: scale, $M_c$. Thus, in a small enough neighborhood of a smooth point
311: $p \in X$, we can choose local coordinates $\P^i$ such that
312: $\P^i(p)=0$ and the K\"{a}hler potential takes the form (modulo
313: K\"{a}hler transformations),
314: \begin{equation}
315:   \label{eq:Kpower}
316:   K (\P, \Pb) = g_{i \jb} \P_i \Pb_\jb + \Real \sum_{\I_n, \Ib_\nb}
317:   \frac{1}{n \nb M_c^{n+\nb-2}} K_{\I_n \Ib_\nb} \P^{i_1}
318:   \cdots \P^{i_n} \Pb^{\ib_1} \cdots \Pb^{\ib_\nb} 
319: \end{equation}
320: where the $K_{\I_n \Ib_\nb}$ are dimensionless and symmetric in the
321: multi-indices $\I_n$ and $\Ib_\nb$ and $g_{i \jb}$ is the K\"{a}hler
322: metric at $p$. Further, as the superpotential $W$ is holomorphic, it
323: can be locally expanded as a power series in the holomorphic
324: coordinates $\P^i$ about the origin $\P^i(p)=0$,
325: \begin{equation}
326:   \label{eq:Wexp}
327:   W(\P^i) = A_0 + \sum_{n, \I_n} \frac{A_{\I_n}}{n} 
328:   \P^{i_1} \cdots \P^{i_n} = M_c^3 W_0 + \sum_{n, \I_n}
329:   \frac{W_{\I_n}}{n M_c^{n-3} } \P^{i_1} \cdots \P^{i_n},
330: \end{equation}
331: where the $A_{\I_n}$ are symmetric in the multi-indices $\I_n$ and
332: have mass dimension $3-n$ while the $W_{\I_n} = M_c^{n-3} A_{\I_n}$
333: are dimensionless.
334: 
335: While the superpotential is holomorphic and not renormalized, the
336: K\"{a}hler potential is renormalized. It is important to check the
337: radiative stability of its assumed form given the form of $W$ in the
338: large $N$ limit.  More precisely, we compute the effective K\"{a}hler
339: potential (in the Wilsonian sense) at a scale $M_{c'}$ lower than
340: $M_c$ by integrating out the modes in a shell of momenta between $M_c$
341: and $M_{c'}$, and require that the corrections are not parametrically
342: larger in $N$ than the bare values.  As we will see, this requirement
343: will restrict the asymptotic growth of both the $A_{\I_n}$ and
344: $K_{\I_n \Ib_\nb}$ parametrically in $N$.
345: 
346: Let us first consider the one-loop corrections to the K\"{a}hler
347: potential coming from the superpotential $W$,
348: \begin{equation}
349: \begin{split}
350:   \label{eq:WcorrK}
351:   \Delta K_W^{(1)} & \sim \left(\frac{1}{8\pi^2}
352:     \log{\frac{M_c^2}{M_{c'}^2}} \right) g^{i
353:     \ib} g^{j \jb} \del_i \del_j W \delb_\ib \delb_\jb \Wb  \\
354:   & \sim \left(\frac{1}{8\pi^2} \log{\frac{M_c^2}{M_{c'}^2}} \right)
355:   \sum_{\I_n, \Ib_\nb} (n+1)(\nb+1)g^{i \ib} g^{j \jb} W_{ij\I_n}
356:   \Wb_{\ib \jb \Ib_\nb} \P^{i_1} \cdots \P^{i_n} \P^{\ib_1} \cdots
357:   \P^{\ib_\nb},
358: \end{split}
359: \end{equation}
360: where $g^{i \jb}$ denotes the uncorrected inverse K\"{a}hler metric at
361: the point $p$, the origin in field space.
362: %By comparing the correction
363: %terms with the corresponding bare terms in (\ref{eq:Kpower}) we see
364: %that we must have that for $n,\nb>3$,
365: %\begin{equation}
366: %  \label{eq:KWcomp}
367: %  \left(\frac{1}{8\pi^2} \log{\frac{M_c^2}{M^2}} \right)
368: %  (n-1)(\nb-1)g^{i \ib} g^{j \jb} W_{ij\I_{n-2}} 
369: %  \Wb_{\ib \jb \Ib_{\nb-2}} \lesssim \frac{1}{(n-2) (\nb-2) }
370: %  K_{\I_{n-2} \Ib_{\nb-2}}. 
371: %\end{equation}
372: %In particular, we see that the natural mass scale of the $A_{\I_n}$
373: %for $n>3$ must indeed be at least as large as $M_c$. [NO WE DON'T]
374: For $n=\nb=1$, we get a one loop correction to the K\"{a}hler metric
375: at the origin $p$ arising from the dimensionless superpotential
376: couplings $A_{ijk}$,
377: \begin{equation}
378:   \label{eq:gcorr}
379:   \Delta g^{(1)}_{i\jb} \sim \left(\frac{1}{2\pi^2}
380:   \log{\frac{M_c^2}{M_{c'}^2}} \right) g^{k \kb} g^{l \lb} A_{ikl}
381:   \Ab_{\jb \kb \lb}. 
382: \end{equation}
383: Radiative stability requires that $\Delta g^{(1)}_{i \jb}$ is
384: ``small'' compared to $g_{i \jb}$, that the coordinate-invariant norms
385: of tangent vectors to $X$ at $p$ under $\Delta g^{(1)}_{i \jb}$ are
386: smaller that those under $g_{i \jb}$, or roughly,
387: \begin{equation}
388:   \label{eq:glessthan}
389:   g^{j \jb} \Delta g^{(1)}_{i\jb} \sim \left(\frac{1}{2\pi^2}
390:   \log{\frac{M_c^2}{M_{c'}^2}}
391:   \right) g^{j \jb} g^{k \kb} g^{l \lb} A_{ikl} \Ab_{\jb \kb \lb}
392:   \lesssim  g^{j \jb} g_{i \jb} = \delta^j_i.
393: \end{equation}
394: We can re-phrase this requirement in precise, coordinate invariant
395: terms by taking traces and determinants of both sides,
396: \begin{equation}
397:   \label{eq:KboundsW}
398:   \det{\left[g^{j \jb} \Delta g^{(1)}_{i\jb} \right]}
399: \lesssim 1, ~~~~ \Tr{\left[ g^{j \jb} \Delta g^{(1)}_{i\jb} \right]}
400: \lesssim N.
401: \end{equation}
402: As promised, this gives us parametric bounds on the growth of the
403: dimensionless superpotential couplings $A_{ijk}$,
404: \begin{align}
405:   \label{eq:AIIIbounds}
406:  g^{i \ib} g^{j \jb} g^{k \kb} A_{ijk} \Ab_{\ib \jb \kb}
407: & \sim \mcO(N), \\
408:   \det{}_{ij}{\left[ g^{j \jb} g^{k \kb} g^{l \lb} A_{ikl} \Ab_{\jb \kb
409:   \lb}\right]} & \sim \mcO(1).
410: \end{align}
411: Since $A_{ijk}$ is a tensor, the interpretation of these bounds on the
412: size of its components is, of course, a coordinate dependent
413: question. Now, in a K\"{a}hler manifold, $d\omega = 0$ implies that it
414: is always possible (see \cite{GrifHar} p.107) to choose local
415: holomorphic coordinates about any smooth point $p$ such that
416: $\P^i(p)=0$ and
417: \begin{equation}
418: \del_i \delb_\jb K (\P, \Pb) = g_{i \jb}(\P,\Pb) = \delta_{i \jb} +
419: \mbox{terms of order $\ge 2$ in the $\P, \Pb$}.
420: \end{equation}
421: In these coordinates the above bounds simplify to,
422: \begin{equation}
423:   \label{eq:AIIIsimp}
424:   \sum_{i,j,k} |A_{ijk}|^2 \sim \mcO(N), ~~~~ 
425: \det{}_{ij}{\left[\sum_{k,l} A_{ikl} \Ab_{jkl} \right]} \sim \mcO(1).
426: \end{equation}
427: Thus, in these special coordinates, we see that in order to satisfy
428: these bounds, the generic components of $A_{ijk}$ must have
429: parametrically suppressed magnitudes at large $N$,
430: \begin{equation}
431: |A_{ijk}| \sim \mcO(N^{-1})
432: \end{equation}
433: though $\mcO(N)$ of them can still be as large as $\mcO(1)$ in
434: that limit. Therefore, we will refer to a tensor $A_{ijk}$ obeying
435: (\ref{eq:AIIIbounds}) as being {\em generically small}. 
436: %We occasionally find it useful to consider, in the
437: %following, the two extreme cases where either {\em all} the $|A_{ijk}|
438: %\sim \mcO(N^{-1})$ (the small component case) or where all but
439: %$\mcO(N)$ of the components vanish (the sparse case).\footnote{ Of
440: %  course, the sparse case is highly non-generic, as a $GL_N(\C)$
441: %  transformation can at best make $N^2$ entries of the tensor sparse.
442: %  We discuss this issue further in section \ref{subsec:Fermat}).}
443: %Further, note that the special coordinates we introduced are ambiguous
444: %up to unitary coordinate transformations, and that these unitary
445: %transformations will generally transform the sparse case to the small
446: %component case, which is the more generic situation.
447: 
448: We can similarly obtain bounds on the coefficients $A_{\I_n}$ and
449: $K_{\I_n \Ib_\nb}$ by considering the leading loop correction to the
450: quadratic term in the K\"{a}hler potential coming from the
451: corresponding higher order terms. One finds a slew of conditions
452: similar in form to \eqref{eq:AIIIsimp}. For instance,
453: \begin{subequations}\label{eq:KboundsII}
454: \begin{align}
455:   M_c^{2n-2} g^{i_1 \ib_1} \cdots g^{i_{n+2} \ib_{n+2}} A_{i_1 \cdots
456:     i_{n+2}} \Ab_{\ib_1 \cdots \ib_{n+2}} & \sim \mcO(N), \\
457:   \det{}_{ij}{\left[M_c^{2n-2} g^{j \jb } g^{i_1 \ib_1} \cdots
458:       g^{i_{n+1} \ib_{n+1}} A_{i_1 \cdots i_{n+1} i} \Ab_{\ib_1 \cdots
459:         \ib_{n+1} \jb} \right]} & \sim  \mcO(1), \\
460:   g^{i_1 \ib_1} \cdots g^{i_{n} \ib_{n}} K_{i_1 \cdots i_{n} \ib_1
461:     \cdots \ib_{n}} & \sim \mcO(N), \\
462:   \det{}_{ij}{\left[g^{j \jb } g^{i_1 \ib_1} \cdots g^{i_{n-1}
463:         \ib_{n-1}} K_{i_1 \cdots i_{n-1} i \ib_1 \cdots \ib_{n-1} \jb}
464:     \right]} & \sim \mcO(1).
465: \end{align}
466: \end{subequations}
467: In particular, this implies that $\mcO(N)$ of the coefficients
468: $A_{\I_n}$ could be $\mcO(1)$, while the generic coefficient must be
469: small,
470: \begin{equation}
471:   \label{eq:Anbounds}
472: |A_{i_1 \cdots i_n}| \sim M_c^{-(n-3)} \mcO(N^{-(n-1)/2}). ~~~~~ (n \ge 3)
473: \end{equation}
474: Thus, just as with $A_{ijk}$, we will refer to any tensor obeying bounds of
475: the form \eqref{eq:KboundsII} as {\em generically small}.  In general,
476: these conditions, as well as a very large number of others
477: corresponding to loop diagrams involving multiple vertices, constrain
478: all the coefficients of the higher order terms in the K\"{a}hler
479: potential and superpotential to be generically small at large
480: $N$.\footnote{We note that there is an additional constraint on the
481:   size of the coefficients coming from the requirement that the power
482:   series we have been writing down actually {\em absolutely converge}
483:   in a region of size $\sim M_c$. These constraints, while stronger
484:   than those coming from radiative stability,  only constrain the
485:   {\it asymptotics} of the $A_{\I_n}$ for large $n$, rather than
486:   constraining the terms at some particular order in $n$.}
487: 
488: \subsection{SUSY Vacua of the Renormalizable Wess-Zumino Model
489:   at Large $N$} \label{subsec:VacStruct}
490: 
491: We now wish to study the vacuum structure of this effective field
492: theory. As discussed in the previous subsection, this amounts to
493: studying the critical points of the superpotential, $\partial_{i}W=0$.
494: For large $N$, we expect that the number of critical points of $W$
495: within a polydisk of radius $M_{c}$ to scale exponentially with $N$.
496: 
497: Since we assume that the power series expansion for $W$ is absolutely
498: convergent in the polydisk, a field theorist might reasonably take the
499: approach of {\it truncating} the power series at some finite order and
500: looking for the critical points of the resulting polynomial. Of
501: course, we cannot hope that a truncation to any finite order
502: polynomial can be an accurate guide to the critical points in the
503: entire polydisk of radius $M_{c}$. At best, we will hope to obtain the
504: critical points inside some much smaller polydisk, of radius $M_{r}\ll
505: M_{c}$. Generically, however, we would not expect to find {\it any}
506: critical points inside this smaller polydisk. The criterion for
507: finding (trustworthy) critical points within this smaller polydisk for
508: small $N$ is well-known: we require that the coefficients of the linear
509: and quadratic terms in \eqref{eq:Wexp} be $A_{\I_1}\sim\CO(M_{r}^{2})$
510: and $A_{\I_2}\sim\CO(M_{r})$, respectively, where
511: \begin{equation}
512:  \frac{M_{r}}{M_{c}} = \epsilon \ll 1
513: \end{equation}
514: As long as the quartic and higher terms in the superpotential are not
515: anomalously large, they give negligible corrections to the critical
516: points determined by truncating $W$ to cubic order\footnote{As long as
517:   we are away from the discriminant locus to be discussed below}.
518: 
519: At large $N$, this is not quite sufficient. Even assuming that the
520: higher $A_{\I_n}$ are {\it generically small}, in the sense of the
521: previous subsection, we still need to require
522: \begin{equation}\label{eq:epsilonbound}
523:   \epsilon < \frac{1}{\sqrt{N}}
524: \end{equation}
525: in order for these higher-order corrections to be negligible. To see
526: this, we can look at the invariant quantity,
527: \begin{equation}
528:   \label{eq:dWbound}
529: (M_{c}^{-4}g^{i\overline{\jmath}}\partial_{i}W
530: \partial_{\overline{\jmath}}\overline{W})|_{\P=\P_{*}}, 
531: \end{equation}
532: where $\P_{*}$ is the critical point derived from the cubic
533: approximation of $W$. Using the generic smallness of the
534: $|A_{\I_n}|\sim M_{c}^{-(n-3)} N^{-(n-1)/2}$ and $|\P_{i*}|\sim M_{r}
535: = \epsilon M_{c}$, we see that the corrections grow parametrically
536: with $N$, unless \eqref{eq:epsilonbound} is satisfied.
537: In what follows, we will keep $\epsilon$ as a free parameter,
538: cognizant of the fact that it must be sufficiently small for the story
539: to work.
540: 
541: In non-supersymmetric theories, ensuring that under radiative
542: corrections the coefficients of the super-renormalizable terms in the
543: potential {\it remain }much smaller than the cutoff is called the {\it
544:   hierarchy problem}. In a supersymmetric theory, there are no
545: perturbative corrections to the superpotential. Nonetheless, the fact
546: that the coefficients of the super-renormalizable terms in the
547: superpotential are of order $M_{r}$, rather than $M_{c}$, means that
548: the point about which we are expanding is, in some sense, ``special.''
549: That will be further brought home in \S\ref{subsec:R}, where we will
550: assume that the point $\Phi=0$ will be a point where there is an
551: unbroken $\BZ_{4}$ R-symmetry.
552: 
553: \subsection{Distributions of Models and Flux Compactifications of IIB}
554: \label{subsec:Flux}
555: 
556: Since F-theory compactifications with flux are the prime motivating
557: example of ``landscape'' models with a large number of vacua, let us
558: pause to consider how such models fit in with our general
559: considerations, as developed so far. Any {\it given} set of fluxes
560: satisfying the requisite tadpole cancellation conditions in an
561: F-theory or IIB Orientifold compactification gives rise to a
562: Gukov-Vafa-Witten \cite{Gukov:1999ya} superpotential for the complex
563: structure moduli.  Since there are, in general, many solutions to the
564: tadpole cancellation conditions, we have {\it an ensemble of theories}
565: with different superpotentials, labeled by the possible fluxes.
566: 
567: In the IIB case, the elements of the ensemble are
568: labeled by all choices of integer fluxes through the $2b^{2,1} + 2$
569: three cycles $\Sigma_a$ of the Calabi-Yau 3-fold $\mcM$,
570: \begin{equation}
571:   \label{eq:Fluxes}
572: \frac{1}{(2 \pi)^2 \alpha'} \int_{\Sigma_a} F = N_a \in \Z, ~~~~
573: \frac{1}{(2 \pi)^2 \alpha'} \int_{\Sigma_a} H = M_a \in \Z.
574: \end{equation}
575: compatible with the tadpole cancellation condition for induced
576: D3-brane charge on $\mcM$, 
577: \begin{equation}
578:   \label{eq:tadpole}
579:   Q_3 = \frac{1}{(2\pi)^4 \alpha'^2}\int_\mcM F \wedge H.
580: \end{equation}
581: where $Q_3$ includes the contribution of space-filling, mobile
582: D3-branes, D7-branes, and O3-planes. Now, if $Q_3 \sim b_{2,1} \sim
583: \mcO(N)$ then tadpole cancellation requires that each of the integer
584: fluxes can then be at most $N_a \sim M_a \sim \mcO(N^{1/2})$. Each
585: choice of fluxes gives rise to a Gukov-Vafa-Witten superpotential for
586: the IIB axio-dilaton $\tau$ and the complex structure moduli $z_i =
587: \int_{A_i} \Omega$ of $\mcM$,
588: \begin{equation}
589:   \label{eq:GVW}
590:   W = \int_\mcM G \wedge \Omega.
591: \end{equation}
592: where $G=F-\tau H$ and the $A_i$ and $B^i$ form a symplectic basis of
593: 3-cycles $\{ \Sigma_a \} = \{ A_i, B^i \}$ in $\mcM$ with respect to
594: the intersection pairing on $H_3(\mcM, \Z)$. In particular, we see
595: that the tadpole cancellation condition places bounds on the growth of
596: $W$ on $N$, 
597: \begin{equation}
598:   \label{eq:GVWbound}
599:   W \supset \sum_i \int_{B^i} G \int_{A_i} \Omega \sim \sum_i (N^i - \tau
600:   M^i) z_i \sim \mcO(N).
601: \end{equation}
602: which is certainly consistent with the condition that the couplings of
603: the moduli are indeed generically small. However, it is not at all
604: clear that there is any point in the moduli space about which the
605: cubic truncation gives a good approximation to the locations of its
606: critical points -- that is, we do not expect the renormalizable
607: approximation to be a useful guide to the vacuum structure.
608: 
609: Further, notice that the superpotential is odd under $G \rightarrow -
610: G$, while the tadpole cancellation condition is even, so if $G$
611: satisfies the tadpole constraint, then so does $-G$. This means that
612: if we find a supersymmetric vacuum $z_i^*$, a critical point of
613: (\ref{eq:GVW}) with a given flux $G$, and a corresponding value $W^*$
614: of the superpotential, then $z_i^*$ also corresponds to a vacuum of
615: the model with flux $-G$ and superpotential $-W^*$.  Thus, even though
616: the average value of $W$ at the supersymmetric vacua for any fixed
617: choice of the fluxes may not vanish, the {\em ensemble} average of $W$
618: certainly does vanish.
619: 
620: Motivated by this example, we will be interested in situations where
621: we are actually given a distribution or ensemble of models. That is,
622: we will assume that physics at high energies $\sim M_p$ can be
623: understood as providing a distribution of coefficients $A_{I_n}$ for
624: the effective low-energy $\lesssim M_c$ models described above.
625: Indeed, in the IIB flux vacua, the physics which determines the values
626: of the quantized fluxes is not even field theoretic in nature - it
627: likely involves high energy string/brane dynamics, topology change,
628: etc. In particular, the scanning of the cosmological constant in these
629: models is explained by high energy physics. Certainly, we could accept
630: such a high energy explanation and restrict our consideration to
631: distributions of coefficients which are symmetric under $A_{I_n}
632: \rightarrow -A_{I_n}$ and share this property of flux vacua.  However,
633: we will focus on situations where there may be a low energy
634: explanation for this scanning.\footnote{Some high energy input,
635:   however, may be inevitable.  For example, one might worry about the
636:   origin of the small parameter $\epsilon = M_r/M_c$ that we required
637:   in order to make the cubic approximation. This appears to be a
638:   tuning of $\mcO(N)$ relevant couplings in the model. Without some
639:   further explanation, one might worry that, together, these
640:   represents a fine-tuning of order $\epsilon^N$, which would wipe out
641:   whatever ``advantage'' we gained in having $2^{N}$ vacua. This need
642:   not be the case if the smallness of each of the $\mcO(N)$ couplings
643:   has a common explanation. Perhaps there is an approximate symmetry,
644:   broken weakly by the effects that generate the superpotential, which
645:   guarantees that these couplings are small. Indeed, such a symmetry
646:   would be reflected in symmetries of the resulting probability
647:   distribution for the couplings. For example, if the coefficients of
648:   the relevant operators were all selected from the {\em same}
649:   distribution, we would indeed only require a single fine tuning of
650:   the distribution. }  One way to achieve this
651: \cite{Arkani-Hamed:2005yv} is through the imposition of an R-symmetry,
652: which we discuss presently.
653: 
654: \subsection{R-Symmetry, The Renormalizable Wess-Zumino Model, and
655:   $\Lambda$}\label{subsec:R}
656: 
657: The general renormalizable Wess-Zumino model of $N$ interacting chiral
658: superfields $\P^i$ is described by a quadratic K\"{a}hler potential,
659: \begin{equation}
660:   \label{eq:KpotWZ}
661: K(\P, \Pb) = g_{i \jb} \P^i \Pb^\jb,
662: \end{equation}
663: and a cubic superpotential,
664: \begin{equation}
665:   \label{eq:superpot}
666: W(\P^i) = A + A_i \P^i + \frac{1}{2} A_{ij} \P^i \P^j + \frac{1}{3}
667: A_{ijk} \P^i \P^j \P^k.
668: \end{equation}
669: Radiative stability of the scaling of the K\"{a}hler potential at
670: large $N$ restrict us to the case that generically $|A_{ijk}| \sim
671: N^{-1}$ with $\mcO(N)$ terms which are $\mcO(1)$. The supersymmetric
672: vacua for this theory are points where,
673: \begin{equation}
674:   \label{eq:SUSYvacWZ}
675: \del_i W = A_i + A_{ij} \P^j + A_{ijk} \P^j \P^k=0,
676: \end{equation}
677: a set of $N$ {\em complex algebraic} equations in the $\P^i$. In
678: particular, these polynomials determine an ideal\footnote{ See section
679:   \ref{sec:Solve} for a basic review of the commutative algebra
680:   language used here, and \cite{CLO, CLO2} for a more detailed
681:   introduction.} in the ring of polynomials in the $\P^i$, $\< \del_i
682: W \> \subset \C[\P^1, \ldots, \P^N ]$. Bezout's theorem (see Chapter
683: 3, Theorem 5.5 of \cite{CLO}) guarantees that for a \emph{generic}
684: choice of the coefficients of $W$,\footnote{By generic, we mean that a
685:   certain polynomial in the coefficients known as a resultant is
686:   non-vanishing -- see \S\ref{subsubsec:generic}.}  these $N$ simultaneous
687: quadratic equations have $2^N$ roots.  From an algebraic point of
688: view, this translates into the fact that the quotient ring,
689: \begin{equation}
690:   \label{eq:quotring}
691:   \C[\P^1, \ldots, \P^N] / \< \del_i W \> \cong \C^{2^N}
692: \end{equation}
693: is a $2^N$-dimensional vector space over $\C$ (see Chapter 3, Theorem
694: 6.2 of \cite{CLO}) generated by the images of the monomials $\P^{i_1}
695: \cdots \P^{i_r}$ with $i_1 < \cdots < i_r \le N$. Explicitly, the
696: quotient ring should be understood as the ``polynomial functions'' on
697: the algebraic variety cut out by the equations generating the ideal,
698: which in this case are the functions on a set of $2^N$ points, i.e. a
699: $2^N$ dimensional vector space.
700: 
701: Further, we saw that the critical points are generically contained in
702: a neighborhood $U$ of radius $\sim M_{r}$ about the origin. Note that if
703: we assume that the constant term $A$ in $W$ is also of order $\lesssim
704: N M_{r}^3$, the superpotential at each of the critical points in $U$ is
705: roughly of the order of\footnote{As a test, we note that the
706:   prototypical example of a landscape superpotential, the
707:   Gukov-Vafa-Witten superpotential for the complex structure moduli of
708:   a IIB orientifold model indeed scales this way - see
709:   (\ref{eq:GVWbound}).  Again, in a appropriate basis, the nonzero
710:   coefficients in the Gukov-Vafa-Witten superpotential can be $\sim
711:   \CO(1)$, but, they are, indeed, very sparse.}
712: \begin{equation}
713:   \label{eq:sizeW}
714: W \sim M_{r}^3 \times\mcO(N).
715: \end{equation}
716: This gives a supersymmetric contribution to the vacuum energy
717: \begin{equation}
718:   \label{eq:vacenergy}
719:    \Lambda = - 3 \frac{|W|^2}{M_p^2}\left(1 + \mcO(M_{r}^2/M_p^2)
720:    \right)  \sim - N \epsilon^{2} M_r^4,
721: \end{equation}
722: where $\epsilon = M_{r}/M_{c}\lesssim 1/\sqrt{N}$ and
723: $M_{c}/M_{p}\lesssim 1/\sqrt{N}$.  Note that all corrections coming
724: from the inclusion of the Chern Connection terms and higher order
725: terms in the K\"{a}hler potential and superpotential are
726: parametrically suppressed at large $N$. 
727: 
728: When supersymmetry is broken, we get a positive contribution,
729: $\Lambda_{S}$ to the vacuum energy. The hope is that the distribution
730: of values for $W$ among the $2^{N}$ vacua will be such that
731: \eqref{eq:vacenergy} can very nearly cancel
732: $\Lambda_{S}$.\footnote{See \cite{ Douglas:2004qg,
733:     Arkani-Hamed:2004fb, Arkani-Hamed:2004yi, Susskind:2004uv,
734:     Kallosh:2004yh, Dine:2004is, Dine:2004ct, Dine:2005iw} for
735:   discussions regarding the scale of supersymmetry breaking in
736:   landscape models.} To achieve this \cite{Arkani-Hamed:2005yv}, one
737: wants a distribution of values of $W$, such that the standard
738: deviation is large, compared to the mean value, $\vvev{W}$.
739: 
740: As the authors of \cite{Arkani-Hamed:2005yv} noted, this is easy to
741: arrange. If the superpotential is odd under $\P^i \rightarrow - \P^i$,
742: then supersymmetric vacua come in pairs ${\P^i}^*$ and $-{\P^i}^*$
743: with opposite values of $W$ and we would have $\vvev{W}=0$. We can
744: enforce this by imposing a $\Z_4$ R-symmetry $\P^i(y,\theta)
745: \rightarrow -\P^i(y,i\theta)$, under which the superpotential must
746: have charge 2 and therefore is an odd polynomial,
747: \begin{equation}
748:   \label{eq:superpotR}
749: W(\P^i) = \sum_{i=1}^N A_i \P^i + \frac{1}{3} \sum_{i,j,k=1}^N A_{ijk}
750: \P^i \P^j \P^k.
751: \end{equation}
752: Such an R-symmetry is clearly non-generic. In particular, the origin
753: in these coordinates must be a special point for such a symmetry to
754: hold, as an expansion of the superpotential at any nearby point in
755: field space certainly would not exhibit the same R-symmetry.  That is,
756: the R-symmetry is spontaneously broken by a vev for the $\P^i$.  In
757: fact, given an arbitrary SUSY non-linear sigma model, there is no
758: reason to believe that its superpotential will generically ever have a
759: point in the target space where $W$ has such a symmetry.  Thus, the
760: imposition of an R-symmetry means that we are restricting our
761: consideration to very special SUSY non-linear sigma models expanded
762: locally about a special point in their target spaces. This is the
763: price that one must pay for a low-energy explanation for the scanning
764: of the vacuum energy. For further discussion regarding this issue, see
765: \cite{DeWolfe:2004ns, Dine:2004dk, Dine:2005gz}. Of course, R-symmetry
766: is phenomenologically desirable for many other reasons as well (see
767: the discussion in \cite{Arkani-Hamed:2005yv}).
768: 
769: The algebraic consequences of the R-symmetry and their geometric
770: interpretations will be of use in our discussion of the statistics of
771: the vacua of this model.  Note that the conditions for unbroken SUSY
772: are actually equations for the scalars $\p^i$ in the chiral
773: superfields $\P^i$ and the R-symmetry acts as a $\Z_2$ parity on these
774: scalars, $\p^i \rightarrow -\p^i$. In particular, the SUSY vacua of
775: this model are determined by $N$ quadratic equations,
776: \begin{equation}
777:   \label{eq:SUSYvacR}
778: \del_i W = A_i + \sum_{j,k=1}^N A_{ijk} \p^j \p^k = 0
779: \end{equation}
780: which are invariant under the $\Z_2$ parity. As a result, these
781: quadratics actually determine an ideal in the ring of $\Z_2$
782: invariants constructed from polynomials in the $\p^i$,
783: \begin{equation}
784: \<\del_i W\> = \< A_i + A_{ijk} \p^j \p^k \> \in \C [\p^1, \ldots,
785: \p^N]^{\Z_2}.
786: \end{equation}
787: We can interpret the ring of invariants $\C[\p^1, \ldots, \p^N
788: ]^{\Z_2}$ as the ``polynomial functions'' on the orbifold $\C^N/\Z_2$.
789: Then, in analogy with the general case, we can consider the quotient
790: ring,
791: \begin{equation}
792:   \label{eq:quotringR}
793:   \C[\p^1, \ldots, \p^N]^{\Z_2} / \< \del_i W \> \cong \C^{2^{N-1}},
794: \end{equation}
795: which is a $2^{N-1}$-dimensional vector space over $\C$ generated by
796: the images of the even monomials $\P^{i_1} \cdots \P^{i_{2r}}$ with
797: $i_1 < \cdots < i_{2r} \le N$. Geometrically, this quotient ring can
798: be interpreted as the polynomial functions on the variety cut out by
799: the $\del_i W$ in the orbifold, which consists of the $2^{N-1}$ images
800: of the $2^{N}$ critical points of $W$ in $\C^N/\Z_2$. 
801: %As we will see
802: %in Section \ref{sec:Stats}, working in this finite dimensional
803: %quotient ring will allow us to easily obtain interesting statistical
804: %features of the vacua of these models. 
805: 
806: \subsection{Fermat Form of the Cubic}\label{subsec:Fermat}
807: 
808: We are interested in the properties of distributions of models with
809: superpotentials respecting the R-symmetry (\ref{eq:superpotR}) and
810: quadratic K\"{a}hler potentials (\ref{eq:KpotWZ}). At first glance, one
811: might expect to describe such a distribution of models as an arbitrary
812: probability distribution on the space of all superpotential and
813: K\"{a}hler couplings, $f(A_i,A_{ijk},g_{i \jb})$. However, we should
814: not distinguish between models which differ by field
815: redefinitions. Of course, arbitrary field redefinitions will
816: not preserve the form of the Wess-Zumino models we are considering.
817: However, it is easy to see that the field redefinitions which do
818: are of the form $\p^i \rightarrow G^i_m \p^m$, where $G^i_m \in GL_N
819: (\C)$. In particular, they act on the space of couplings as,
820: \begin{equation}
821:   \label{eq:GLNac}
822: A_i \rightarrow A_m G^m_i, ~~~~ A_{ijk} \rightarrow A_{mnp} G^m_i
823: G^n_j G^p_k, ~~~~ g_{i \jb} \rightarrow g_{m \nb} G^m_i
824: \bar{G}^\nb_\jb.
825: \end{equation}
826: If we posit some probability distribution on the space of {\it
827:   couplings}, this distribution gets averaged over $GL_{N}(\C)$ orbits
828: to produce a probability distribution on the space of {\it theories}.
829: That is awkward to deal with. If possible, it is much better to fix
830: the redundancy by choosing a gauge for the $GL_N(\C)$ field
831: redefinitions. Since $GL_N(\C)$ is $N^2$ dimensional, such a gauge
832: condition should involve precisely $N^2$ independent, complex
833: algebraic conditions.  Now, in working with algebraic equations, it is
834: often convenient to redefine our variables in order to make the
835: coefficients of the terms of highest degree as simple as possible. If
836: $W$ is a generic polynomial of degree $d$, we can use the $GL_N(\C)$
837: symmetry to enforce $N^2$ conditions on $A_{\I_d}$,
838: \begin{equation}
839:   \label{eq:GaugeCond}
840: A_{i \cdots i} = 1, ~~~~ A_{ij \cdots j} = A_{jij \cdots j} = \cdots =
841: A_{j \cdots j i} = 0,~~ \mbox{for $i \neq j$}.
842: \end{equation}
843: We will refer to a polynomial satisfying these conditions as being in
844: Fermat form, 
845: \begin{equation}
846:   \label{eq:FermatForm}
847:   W = \frac{1}{d} \left( (\p^1)^d + \cdots + (\p^N)^d \right) +
848:   (\mbox{terms of degree $\le d-1$ in each $\p^i$}).
849: \end{equation}
850: Note that if $W$ is {\em odd} (so it has a $\Z_2$ R-symmetry as
851: above), then the Fermat form is somewhat stronger, as the additional
852: terms in (\ref{eq:FermatForm}) actually have degree at most $(d-2)$ in
853: any $\p^i$. In particular, for the case of interest here, $d=3$, with
854: the R-symmetry above, the superpotential takes the form,
855: \begin{equation}
856:   \label{eq:FermatCubic}
857:   W(\p^i) = \sum_{i=1}^N \left( \tfrac{1}{3} (\p^i)^3 - a_i \p^i \right) 
858: - \sum_{i < j < k}^N b_{ijk} \p^i \p^j \p^k,
859: \end{equation}
860: where $a_i$ has mass dimension 2 and is generically $M_{r}^2\times
861: \mcO(1)$ and the $b_{ijk}$ are symmetric, traceless (so $b_{iii} =
862: b_{ijj} = b_{jij} = b_{jji}=0$), dimensionless and generically
863: $|b_{ijk}| \sim \mcO(N^{-1})$ with at most $\mcO(N)$ of them
864: $\mcO(1)$. In going to Fermat form, it is important to note that we
865: cannot also simultaneously set $g_{i \jb} = \delta_{i \jb}$. Further,
866: the genericity of $W$ is important here --- not all polynomials can be
867: put into Fermat form. For example, the following model with
868: spontaneous breaking of supersymmetry via the O'Raifeartaigh
869: Mechanism,
870: \begin{equation}
871:   \label{eq:ORaif}
872:   W = Z (X^2 - a) + Y X^2
873: \end{equation}
874: cannot be put into this form. In fact, it is easy to show by analytic
875: computations using eigenvalue methods (see \S\ref{sec:Solve})
876: that for $N=3$ and $d=3$ no polynomial in Fermat form exhibits the
877: O'Raifeartaigh Mechanism.
878: 
879: We believe that this may be true much more generally. More precisely,
880: we conjecture that {\em all} cubic, R-symmetric superpotentials of the
881: Fermat form (\ref{eq:FermatCubic}) have at least a pair of
882: supersymmetric vacua. While a proof of this fact is far beyond the
883: scope of this work, we hope to return to this question in the future.
884: %\ref{app:Null}. 
885: 
886: \subsection{Symmetries of the Fermat Form}\label{subsec:symmetries}
887: 
888: Just as is often the case in gauge fixing, it is important to note
889: that our gauge condition (\ref{eq:GaugeCond}) does not completely fix
890: the $GL_N(\C)$ redundancy. For any generic $A_{\I_d}$, there is a finite
891: subgroup $H$ of $GL_N(\C)$ transformations which transforms the
892: coefficients in such a way as to keep $A_{\I_d}$ in the Fermat form
893: (\ref{eq:FermatForm}).  Namely, $h^m_i \in H$ are the solutions to
894: $N^2$ algebraic equations of degree $d$ in the $N^2$ complex
895: components $h_m^i$,
896: \begin{equation}
897:   \label{eq:Symmetry}
898: A_{m_1 \cdots m_d} h^{m_1}_{i} \cdots h^{m_d}_{i} = 1, ~~~~ A_{m_1
899:   \cdots m_d} h^{m_1}_{i} h^{m_2}_{j}  \cdots h^{m_d}_{j} = 0, ~~
900:   \mbox{for $i \neq j$}.
901: \end{equation}
902: If these were generic degree $d$ equations, Bezout's theorem would
903: tell us that they have $d^{N^2}$ solutions.  However, it turns out
904: that this is not the case for generic $A_{\I_d}$, and $d^{N^2}$ is
905: actually a strict upper bound on the order of $H$. We will show this
906: by exhibiting a large subgroup of $H$ which is independent of
907: $A_{\I_d}$ and whose order does not divide $d^{N^2}$. First, note that
908: permutations of the $\p^i$ certainly won't take us out of Fermat
909: form, so we expect that the symmetric group on $N$ variables $S_N
910: \subset H$. Next, note that multiplying any $\p^i$ by a $d^{th}$ root
911: of unity also won't ruin the special form, so we get a subgroup of $H$
912: isomorphic to $(\Z_d)^N$ given by,
913: \begin{equation}
914:   \label{eq:zdsymm}
915:   h_i^j = \delta_i^j \zeta^{n_i}, ~~~~ \zeta = e^{2\pi i/d}, ~~~~ 1 \le
916:   n_i \le d. 
917: \end{equation}
918: Thus, we see that $S_N \ltimes (\Z_d)^N$, which has order $N! \times
919: d^N$, must be a subgroup of H. Since $N!$ generally does not divide
920: $d^{N^2}$, we see that,
921: \begin{equation} 
922: N! \times d^N \le |H| < d^{N^2}
923: \end{equation}
924: In the simplest non-trivial case, $N=3$ and $d=3$, we numerically
925: solved (\ref{eq:Symmetry}) using Mathematica for various values of the
926: coefficients and found that $H$ is generically of order $648 = 2^3
927: \times 3^4 = 3! \times 3^3 \times 4$. In particular, there are indeed
928: elements of $H$ which depend on $A_{\I_d}$, which suggests that $|H|$
929: may generally be strictly larger than $S_N \ltimes (\Z_d)^N$. This is
930: not unexpected. Rather generally, the space of common solutions of
931: these polynomial equations is an algebraic variety. Different values
932: for the coefficients of the polynomials will nonetheless lead to
933: isomorphic algebraic varieties. $H$ is the modular group, and we will
934: not attempt to give a full characterization of $H$ beyond its
935: ``obvious''$S_N\ltimes (\Z_d)^N$ subgroup. We will content ourselves
936: with exploiting the constraints that stem from this latter subgroup.
937: 
938: To summarize, after fixing gauge and imposing R-symmetry, each model
939: is uniquely described up to a discrete symmetry $H$ containing $S_N
940: \ltimes (\Z_3)^N$ by a cubic superpotential in Fermat form
941: (\ref{eq:FermatCubic}) and a quadratic K\"{a}hler potential
942: (\ref{eq:KpotWZ}). Thus, a distribution of field theory landscape
943: sectors is a probability distribution on the space of couplings,
944: \begin{equation}
945:   \label{eq:InvtDist}
946: f(a_i, b_{ijk}, g_{i \jb}) d a_i d b_{ijk} d g_{i \jb}
947: \end{equation}
948: which is invariant under $H \supset S_N \ltimes (\Z_3)^N$, 
949: \begin{equation}
950:   \label{eq:Hact}
951:   a_i \rightarrow a_m h^m_i, ~~~~~ b_{ijk} \rightarrow b_{mnp} h^m_i
952:   h^n_j h^p_k, ~~~~~ g_{i \jb} \rightarrow g_{m \nb} h^m_i \bar{h}^\nb_\jb.
953: \end{equation} 
954: We assume that this distribution is fixed by high energy
955: physics. Lacking any clear physical input which distinguishes
956: among the different vacua of a {\em given} low-energy model, we
957: simply treat these vacua democratically. In the following sections, we
958: use algebraic methods to compute some statistical properties of
959: various physical quantities averaged over the $2^N$ vacua of each
960: model as a function of the couplings. Of course, these
961: model averages can then be further averaged over the ensemble using
962: the high energy distribution function. 
963: 
964: \section{Eigenvalue Methods for Solving Algebraic Equations}\label{sec:Solve}
965: 
966: The solution of $N$ simultaneous linear equations in $N$ variables is
967: something we all learn to do from an early age using several different
968: methods. For example, we may decide to do it via Gaussian elimination
969: or via matrix methods - using determinants and Cramer's rule to invert
970: the matrix of coefficients. A natural question to ask is if there
971: exist analogues of these methods which could be used to solve
972: simultaneous equations of higher degree. Indeed, this is the case. The
973: method of Gaussian elimination has a vast generalization in the study
974: of algebraic elimination theory and Gr\"{o}bner bases methods (see
975: \cite{CLO, CLO2}). These methods provide algorithms which allow one to
976: systematically eliminate variables one by one in any given system of
977: equations at the cost of increasing the degrees of the resulting
978: system of equations in fewer variables and, with some further work,
979: find (approximate) solutions. Unfortunately, these are not directly
980: useful for us as we will be interested in the statistical properties
981: (such as the average and variance of the superpotential and other
982: observables) of such solutions as a function of the coefficients -
983: i.e. in the properties of solutions of {\em families} of such
984: equations.  It turns out that the generalization of matrix methods to
985: the case of simultaneous equations of higher degree does end up being
986: quite useful for these purposes.
987: 
988: \subsection{Rings, Ideals and a Toy Model - Quadratic $W$}
989: 
990: We will introduce these methods by applying them to the simple
991: case of a {\em generic} quadratic superpotential,
992: \begin{equation}
993:   \label{eq:quad}
994:   W = W_0 - A_i z_i + \frac{1}{2} C_{ij} z_i z_j = W_0 - A \cdot z +
995:   \frac{1}{2} z \cdot C \cdot z.
996: \end{equation}
997: Of course, the ``statistics of vacua'' in this case may seem to be,
998: well, vacuous - the critical point of this polynomial is given by the
999: unique solution of $N$ linear equations,
1000: \begin{equation}
1001:   \label{eq:critquad}
1002: \del_i W = -A_i + C_{ij} z_j = 0 \Rightarrow C \cdot z = A,
1003: \end{equation}
1004: which is, of course,
1005: \begin{equation}
1006:   \label{eq:linsol}
1007:   z = C^{-1} \cdot A,
1008: \end{equation}
1009: where the genericity condition is that $\det C \neq 0$, or $C \in GL_N(\C)$.
1010: However, if we are given a {\em distribution} of coefficients for $W$,
1011: we might be interested in the corresponding distribution of the values
1012: of $W$ at the critical point. To attack such problems more generally,
1013: it is essential (though, in this case, akin to trying to kill an ant
1014: with a machine gun) to recast these questions in the algebraic
1015: language of polynomial rings and ideals.
1016: 
1017: To begin with, let us quickly review what we will need about ideals,
1018: rings, and varieties. Roughly, a ring is a set $R$ whose algebraic
1019: properties mimic many of those of the integers $\Z$. That is, we may
1020: add, subtract, and multiply elements of $R$ and obtain new elements in
1021: $R$, though the same may not be true for division. Recall that an
1022: ideal $I$ contained in a ring $R$ is defined by the property that if
1023: $f, g \in I$ and $r, s \in R$ then $rf + sg \in I$.  In particular,
1024: this implies that the {\em quotient ring} of $R$ by $I$, $Q=R/I$ where
1025: we identify all elements of $R$ differing by an element in $I$, is
1026: well-defined. Note that all the rings that we consider are {\em
1027: Noetherian} rings, which means that all their ideals are finitely
1028: generated, so
1029: \begin{equation}
1030:   \label{eq:ideal}
1031: I = \< f_1, \ldots, f_m \> = \left\{ \left. \sum r_i f_i \right| r_i
1032: \in R \right\} = \< f_i \>.
1033: \end{equation}
1034: The rings we will be most interested are rings of polynomial functions
1035: in $N$ variables, $R = \C[z_1, \ldots, z_N]$, and their quotients by
1036: ideals $I$ generated by some polynomials $f_i(z_1, \ldots, z_N) \in
1037: R$. Thus, $I$ is just the set of all polynomials which vanish on the
1038: zero locus in $\C^N$ of the simultaneous set of polynomial equations
1039: $f_i(z_j) = 0$, $i = 1, \ldots, m$. This zero-locus is called the {\em
1040: algebraic variety} in $\C^N$ corresponding to the ideal $I$. Thus,
1041: non-vanishing elements of the quotient ring $Q=R/I$ are precisely the
1042: non-trivial residues of polynomial functions on $\C^N$ restricted to
1043: the variety - they define a notion of ``polynomial functions'' on the
1044: variety.
1045: 
1046: As we have already mentioned above, the polynomials $\del_i W$ which
1047: determine the critical point generate an ideal in the ring of
1048: polynomials in $N$ variables,
1049: \begin{equation}
1050:   \label{eq:idealWquad}
1051:   I = \< \del_i W \> = \< -A_i + C_{ij} z_j \> = \< -A + C \cdot z\>.
1052: \end{equation}
1053: Clearly, any linear combination of the generators $\del_i W$ with
1054: complex coefficients is also in the ideal. In particular, if $\det C
1055: \neq 0$ then $C$ is invertible and we can consider the $N$ elements of
1056: $I$, $z - C^{-1} \cdot A$, obtained by multiplying the $N$ generators
1057: $\del_i W$ by the matrix $C^{-1}$. In fact, as $C$ is invertible,
1058: these are also generators of the ideal $I$. Thus, in the quotient
1059: ring $Q=\C[z_1, \ldots, z_N] / I$, we can use the $N$ relations $z_i -
1060: {C^{-1}}_{ij} A_j=0$ to systematically eliminate the $z_i$ and rewrite
1061: any element of $Q$ in terms of multiples of the identity -
1062: constants. So, we see that
1063: \begin{equation}
1064:   \label{eq:ringquad}
1065:   Q = \C[z_1, \ldots, z_N] / I \cong \C,
1066: \end{equation}
1067: a one dimensional complex vector space, generated by constants. In
1068: particular, the residue of any polynomial $f(z_1, \ldots, z_N) \in
1069: \C[z_1, \ldots, z_N]$ in $Q$ is obtained by setting to zero all
1070: elements of the ideal $I$, that is, by substituting $z = C^{-1} \cdot
1071: A \in \C^N$ into $f$.  Thus, we see that {\em the residue of $f$ in
1072: $Q$ is precisely its value at the critical point of $W$}, and that the
1073: variety corresponding to $I$ is the single critical point $z = C^{-1}
1074: \cdot A \in \C^N$. Indeed, as the only functions on a point are
1075: constant functions - the value at the point - this is consistent with
1076: our intuition that the quotient ring $Q=\C[z_1, \ldots, z_N] / I$
1077: should be interpreted as the ``polynomial functions'' on a
1078: point. Further, note that as $Q$ is isomorphic as a ring to $\C$, the
1079: product structure of the ring of polynomials is preserved in the
1080: quotient. That is, we can think of the residue of $f$ in $Q$ as an
1081: operator on other elements of $Q$ which is multiplication by its value
1082: at the critical point. In particular, using the fact that $\del W, z -
1083: C^{-1} \cdot A \in I$, we can compute the residue of $W$ in $Q$ by,
1084: \begin{equation}
1085: W = W_0 - \frac{1}{2} A \cdot z + \frac{1}{2} z \cdot \del W \sim W_0
1086: - \frac{1}{2} A \cdot z \sim W_0 - \frac{1}{2} A \cdot C^{-1} \cdot A
1087: \in Q.
1088: \end{equation}
1089: Of course, we could have just plugged in our solution, but it turns
1090: out that this kind of computation generalizes to the case of higher
1091: degree in a way that is useful for computing statistical properties.
1092: 
1093: \subsubsection{Non-Generic Quadratic Superpotentials}
1094: 
1095: It is also interesting to ask what happens if $C$ is {\em not}
1096: generic, if $\det C = 0$. Then, $C$ has a non-trivial null space which
1097: is the space solutions of the homogeneous linear equations we get in
1098: the limit that we take $A_i \rightarrow 0$. We can describe this limit
1099: more precisely by lifting the equations to projective space, to
1100: $\CP{N}$. That is, we add a new coordinate $z_0$ (which we may think
1101: of physically as a chiral superfield corresponding to a (complexified)
1102: scale factor) and then consider the homogeneous system of $N$ linear
1103: equations $C_{ij} z_j - A_i z_0 = 0$ up to (complex) scaling in
1104: $\CP{N}$.  In the open set $U_0 = \{z_i \sim \lambda z_i \in \C^{N+1}-
1105: \{0\} | z_0 \neq 0 \} \subset \CP{N}$, the $u_i = \frac{z_i}{z_0}$ are
1106: good coordinates and the above equation reduces to the inhomogeneous
1107: equation with $z_i \rightarrow u_i$. Thus, if $\det C \neq 0$, the
1108: unique solution of the inhomogeneous equation in $U_0$ gives us the
1109: unique solution in $\CP{N}$ as well. However, the advantage of
1110: projectivizing the problem is that we have added points corresponding
1111: to solutions even in the non-generic case. If $\det C=0$ then we have
1112: at least one solution in $\CP{N}$ which has $z_0 =0$ given by a
1113: one-dimensional subspace of the null space of $C$. As the $u_i
1114: \rightarrow \infty$ as we take $z_0 \rightarrow 0$, we can think of
1115: such a solution as ``a solution at infinity'' from the perspective of
1116: $U_0$. Thus, we see that non-generic points in the space of couplings
1117: of the quadratic superpotential correspond to situations in which the
1118: supersymmetric vacuum has ``run off to infinity'' in field space. Now,
1119: since all of our field theory computations are approximations only
1120: valid locally in field space, these non-generic points should more
1121: properly be understood physically as points in the space of couplings
1122: in which those approximations break down.
1123: 
1124: \subsection{The General Case}
1125: 
1126: Now, let's consider the generalization of the above to the case that
1127: $W$ is of degree $d>2$,
1128: \begin{equation}
1129:   \label{eq:supWz}
1130: W(z_1, \ldots, z_N) = \sum_{n=1, \I_n}^d \frac{A_{\I_n}}{n} z_{i_1}
1131: \cdots z_{i_n}.
1132: \end{equation}
1133: The critical points of $W$ are then given by the simultaneous
1134: solutions of $N$ degree $d-1$ polynomials, $\del_i W = 0$, and, as we
1135: have already mentioned above, the polynomials $\del_i W$ generate an
1136: ideal in the ring of polynomials in $N$ variables,
1137: \begin{equation}
1138:   \label{eq:idealWz}
1139:   I = \< \del_i W \> = \< \sum_{n=1, \I_{n-1}}^d A_{i\I_{n-1}} z_{i_1} \cdots
1140:   z_{i_{n-1}} \>.
1141: \end{equation}
1142: Just as before, we may now try to simplify the generators by taking
1143: $\C$-linear combinations of them. In particular, it would be nice to
1144: simplify the form of the terms of highest degree as we did in
1145: obtaining the Fermat form. So, define the matrix $C$ by,
1146: \begin{equation}
1147:   \label{eq:CMatrix}
1148: A_{i \cdots i} = C_{ii}, ~~~~ A_{ij \cdots j} = A_{jij \cdots j} = \cdots =
1149: A_{j \cdots j i} = C_{ij},~~ \mbox{for $i \neq j$}.
1150: \end{equation}
1151: Now, just as in the case of the quadratic superpotential, if $\det C
1152: \neq 0 \Rightarrow C \in GL_N(\C)$, we can multiply the $\del_i W$ by
1153: the $C^{-1}$ and obtain a simpler set of generators,
1154: \begin{equation}
1155:   \label{eq:idealWzs}
1156:   I = \< \sum_{n=1, \I_{n-1}}^d \sum_{j=1}^N
1157:   C^{-1}_{ij} A_{j\I_{n-1}} z_{i_1} \cdots z_{i_{n-1}} \> = \< z_i^{d-1} +
1158:    (\mbox{terms of degree $\le d-2$ in any $z_j$}) \>.
1159: \end{equation}
1160: Of course, if $W$ is in Fermat form, then $C_{ij} = \delta_{ij}$ and
1161: so the generators $\del_i W$ themselves take precisely this
1162: form.\footnote{Now, while it is obvious that a superpotential in
1163:   Fermat form has $\det C \neq 0$, one might wonder if this condition
1164:   is in general necessary and sufficient for the existence of a
1165:   $GL_N(\C)$ coordinate transformation taking any given superpotential
1166:   to Fermat form. However, this is not at all clear, as $C$ involves
1167:   only nearly diagonal components of $A_{\I_d}$ while the
1168:   transformation to Fermat form (\ref{eq:GLNac}) involves all the
1169:   components of $A_{\I_d}$. We will not need to delve into this issue
1170:   further for our purposes.} As we have argued earlier, the Fermat
1171: Form is convenient for other reasons as well, so we will assume
1172: henceforth that $W$ is given in Fermat form. Just as before, the above
1173: form suggests that in the quotient ring $Q = \C[z_i, \ldots, z_N] / I$
1174: we should be able to use the generators in (\ref{eq:idealWzs}) to
1175: systematically eliminate all powers of $z_i$ greater than $(d-2)$ and
1176: write any element of $Q$ only in terms of monomials with powers of
1177: $z_i$ less than or equal to $(d-2)$. This is in fact the case as long
1178: as $W$ is a {\em generic} polynomial of degree $d$ (for a rigorous
1179: proof of this fact, see Chapter 3, Theorem 6.2 of \cite{CLO}). It is
1180: not difficult to show this explicitly in the case of interest $d=3$,
1181: but the result is not particularly enlightening and will not be
1182: presented here. We only note here that the computation requires that
1183: several large matrices in the coefficients be invertible, which we
1184: believe is a computational manifestation of the assumption of
1185: genericity. For the following, we will proceed assuming that $W$ is
1186: generic, and discuss what we mean by this more precisely at the end of
1187: the discussion.
1188: 
1189: Thus, for $W$ generic, the monomials $z_{1}^{d_1} \cdots z_{N}^{d_N}$
1190: with $0 \le d_i \le d-2$ form a basis for $Q$ as a vector space, and
1191: the relations will allow us to express the residue of any polynomial
1192: $f(z_1, \ldots, z_N) \in \C[z_1, \ldots, z_N]$ in $Q$ in terms of this
1193: basis. This is particularly simple in the case of the superpotential
1194: itself, as we have
1195: \begin{equation}
1196:   \label{eq:superpotQ}
1197:   W= \frac{1}{d} \sum_i z_i \del_i W + \sum_{n=0,\I_{n}}^{d-1}
1198:   \frac{d-n}{d} A_{\I_n} z_{i_1} \cdots z_{i_n} \rightarrow
1199:   \sum_{n=0,\I_{n}}^{d-1} \frac{d-n}{d} A_{\I_n} z_{i_1} \cdots z_{i_n} \in Q.
1200: \end{equation}
1201: Thus, we see that the quotient ring,
1202: \begin{equation}
1203:   \label{eq:quotringz}
1204:   \C[z_1, \ldots, z_N] / \< \del_i W \> \cong \C^{(d-1)^N},
1205: \end{equation}
1206: is a $(d-1)^N$-dimensional vector space over $\C$ generated by the
1207: images of the monomials $z_{1}^{d_1} \cdots z_{N}^{d_N}$ with $0 \le
1208: d_i < d-1$. Further, as we mentioned earlier, this quotient ring
1209: should be understood as the ``polynomial functions'' on the variety
1210: cut out by the generators of the ideal, which in this case are the
1211: functions on a set of $(d-1)^N$ points. In particular, in analogy with
1212: with the quadratic case, we expect that the residue of any polynomial
1213: $f(z_1, \ldots, z_N) \in \C[z_1, \ldots, z_N]$ in $Q$ should be
1214: related to the values of $f$ at the critical points of $W$. However,
1215: the fundamental difference between the case $d>2$ and the simple case
1216: of a quadratic superpotential is that here,\ {\em the residue of $f$
1217:   in $Q$ is a vector}, and it is not immediately obvious how the
1218: components of that vector in any given basis are related to its values
1219: at the critical points. However, as $Q$ is a ring, $f$ also acts by
1220: multiplication on $Q$.  Since $Q$ is a vector space, and as
1221: multiplication by $f$ is $\C$-linear, $f$ is also a {\em linear
1222:   operator} on $Q$, which we can represent as a matrix $\mf{f}$ in our
1223: basis of monomials in an obvious way. In particular, we could consider
1224: $f=z_i$ as an operator or matrix $\mf{z_i}$ in this sense. Clearly,
1225: the action of $\mf{z_i}$ on most basis vectors (for $d_i < d-2$) is
1226: totally obvious,
1227: \begin{equation}
1228: \mf{z_i} : z_{1}^{d_1} \cdots z_i^{d_i} \cdots z_{N}^{d_N} \rightarrow
1229: z_{1}^{d_1} \cdots z_i^{d_i+1} \cdots z_{N}^{d_N}, ~~~ d_i < d-2.
1230: \end{equation}
1231: However, for $d_i = d-2$ we must use the relations in the ideal to
1232: re-express the resulting monomial as a linear combination of the basis
1233: elements. Assuming that this is done for each $z_i$, we obtain $N$
1234: matrices $\mf{z_i}$, each representing multiplication by one of the
1235: $z_i$. We can consider the characteristic equation for each of these
1236: matrices,
1237: \begin{equation}
1238:   \label{eq:chareq}
1239:   0 = \det{(\mf{z_i} - \lambda \Bid)} = P_i(\lambda),
1240: \end{equation}
1241: which is a polynomial equation of degree $(d-1)^N$ in $\lambda$.
1242: Generically, this equation will have $(d-1)^N$ distinct roots,
1243: corresponding to the $(d-1)^N$ eigenvalues of the matrix $\mf{z_i}$. In
1244: particular, the matrix $\mf{z_i}$ must be diagonal in the corresponding
1245: basis of eigenvectors. Since $Q$ is a commutative ring, all
1246: matrices corresponding to multiplication by functions must mutually
1247: commute, and this must in particularly be true for the $N$ matrices
1248: $\mf{z_i}$. Thus, all the matrices corresponding to multiplication by
1249: functions $f \in \C[z_1,\ldots,z_N]$ can be simultaneously
1250: diagonalized in the basis of eigenvectors of the $\mf{z_i}$. In particular,
1251: in this eigenbasis, $Q$ splits {\em as a ring} into the direct sum of
1252: $(d-1)^N$ rings isomorphic $\C$, 
1253: \begin{equation}
1254:   \label{eq:ringstruct}
1255:   Q = \C[z_1, \ldots, z_N] / \< \del_i W \> \cong
1256:   \underbrace{\C \oplus \cdots \oplus \C}_{(d-1)^N}.
1257: \end{equation}
1258: each of which one may naturally associate with the ring of functions
1259: on one of the critical points of $W$.  Thus, generalizing our result
1260: from the toy model, the eigenvalue of $\mf{f}$ associated with each
1261: eigenvector corresponds to the value of $f$ at the corresponding
1262: critical point. If we take $f=z_i$, then the eigenvalues of $\mf{z_i}$
1263: are precisely the $z_i$ coordinates of the critical points. Further,
1264: this means that the trace of $\mf{f}$ in {\em any basis} is the sum of
1265: the values of $f$ at the critical points of $W$. In particular, we can
1266: do this trace explicitly in the monomial basis. Now, as powers of $f$
1267: correspond to powers of the corresponding matrix $\mf{f}$, we can take
1268: their traces to compute the sums of powers of the values of $f$ at the
1269: critical points as well. Thus, we see that by computing traces in the
1270: monomial basis, we can compute all holomorphic moments of $f$ at the
1271: critical points of a generic superpotential $W$. In other words, we
1272: can effectively compute all (holomorphic) statistical properties of
1273: the values of any polynomial $f$ at the critical points of $W$.
1274: 
1275: \subsubsection{Resultants and Genericity} \label{subsubsec:generic}
1276: 
1277: There are two basic ways in which the generic situation of $(d-1)^{N}$
1278: vacua can break down. Roots of the system of polynomials can coincide,
1279: or roots can run off to infinity. The former is something that already
1280: was implicit in the setup of \cite{Arkani-Hamed:2005yv}, where the $N$
1281: chiral fields were assumed to be decoupled. When roots run together,
1282: the tunneling between the respective vacua is no longer suppressed,
1283: and one should not really count them as independent vacua. It is not
1284: difficult to modify the following considerations to deal with this
1285: situation, but we will not discuss this further here (see \cite{CLO}
1286: for details).  Having roots ``run off to infinity'' is a phenomenon
1287: that can occur only when interactions between the fields are turned
1288: on, and our analysis breaks down in this case. 
1289: 
1290: It turns out \cite{CLO} that both the collision and expulsion of roots
1291: can be captured by a single genericity condition.\footnote{It is a
1292:   certain integral polynomial in the coefficients $A_{\I_n}$ of $W$
1293:   known as the u-resultant.} As it is only the latter non-genericity
1294: which is fatal, it is useful to have a precise criterion when it
1295: obtains. In the case of a quadratic superpotential, we saw that the
1296: critical point runs off to infinity precisely when $\det C = 0$.  Just
1297: as before, we can describe this more precisely by considering the
1298: homogenization of the polynomial equations $\del_i W=0$ and lifting
1299: them to to equations in $\CP{N}$. It turns out (see \cite{CLO} Chapter
1300: 3) that there is an analogous integral polynomial in the coefficients
1301: of the monomials in $W$ of highest degree $A_{\I_d}$ known as a
1302: multi-polynomial resultant which plays the same role as $\det C$ in
1303: degree $>2$. The vanishing of this particular resultant indicates that
1304: some solutions have ``run off to infinity'', or equivalently, the
1305: presence of solutions in the compliment of the open set $U_0$. We will
1306: refer to the vanishing locus of the resultant as the {\it
1307:   discriminant locus}.  While we will not describe this resultant in
1308: general, we will be able to compute it exactly in an example $d=N=3$
1309: and see the advertised behavior explicitly in the next section.
1310: Finally, note that the cubic polynomial superpotentials we are
1311: interested in are approximations valid only within some small polydisk
1312: about the origin. Indeed, in the real situation, we don't literally
1313: have roots running off to infinity. Once they leave the polydisk we
1314: are considering, our approximation of truncating to the renormalizable
1315: terms in the Lagrangian breaks down.  We simply no longer {\it trust}
1316: those solutions which have wandered too far from the origin.
1317: 
1318: \section{Holomorphic Moments and the Statistics of SUSY
1319:   Vacua}\label{sec:Stats} 
1320: 
1321: We will now restrict our consideration to the physically relevant case of
1322: a renormalizable, R-symmetric, cubic superpotential for $N$ chiral
1323: fields in Fermat form,
1324: \begin{equation}
1325:   W(z_i) = \sum_{i=1}^N \left( \tfrac{1}{3} z_i^3 - a_i z_i \right) 
1326: - \sum_{i < j < k}^N b_{ijk} z_i z_j z_k,
1327: \end{equation}
1328: where we recall that $b_{ijk}$ is symmetric with vanishing diagonals,
1329: i.e. $b_{iii} = b_{iij} = b_{iji} = b_{jii} =0$. Dimensional analysis
1330: as well as the $S_N\ltimes (\Z_3)^N$ symmetry will be important tools
1331: for the analysis to follow. As the superpotential $W$ has mass dimension
1332: $3$, the $z_i$ must have mass dimension $1$ while the $a_i$ have mass
1333: dimension $2$ and the $b_{ijk}$ are dimensionless. Further,
1334: $\boldsymbol{\zeta} \in (\Z_3)^N$ acts as,
1335: \begin{equation}
1336:  z_i \rightarrow z_i \zeta_i, ~~~~  a_i \rightarrow a_i \zeta_i^{-1},
1337:  ~~~~ b_{ijk} \rightarrow b_{ijk} \zeta_i^{-1} \zeta_j^{-1}
1338:  \zeta_k^{-1}, ~~~~ \boldsymbol{\zeta} = \{ \zeta_i \},
1339: \end{equation}
1340: where $\zeta_i$ is in the $i^{th}$ $\Z_3$, under which $z_i$ has
1341: charge $+1$ while the $a_i$ and $b_{ijk}$ have charge $-1$.  Note that
1342: the equations for the critical points of the cubic in Fermat form are,
1343: \begin{equation}
1344:   \label{eq:critp}
1345:   z_i^2 - a_i - \sum_{j<k} b_{ijk} z_j z_k=0,
1346: \end{equation}
1347: and a basis for the quotient ring $Q$ is given by the $2^N$ monomials,
1348: \begin{equation}
1349:   \label{eq:basisN}
1350:   \{ 1,z_{i}, z_{i} z_{j}, \ldots, z_1 \cdots z_N \},
1351: \end{equation}
1352: where at most one power of each $z_i$ appears in each monomial. The
1353: residue of $W$ in $Q$ written in this basis is,
1354: \begin{equation}
1355:   \label{eq:Wres}
1356:   W = -\frac{2}{3} \sum_{i=1}^N a_i z_i.
1357: \end{equation}
1358: Note that $\tr{\mf{W}}$ and all other holomorphic moments of odd
1359: powers of $\mf{W}$ or $\mf{z_i}$ vanish due to the R-symmetry.
1360: Explicitly, since the relations (\ref{eq:critp}) are even, any odd
1361: power of the operator $\mf{z_i}$ maps monomials of even degree to
1362: those of odd degree and vice-versa and therefore never has any
1363: non-zero diagonal components.
1364: 
1365: We will be interested in determining the distribution of values of $W$
1366: among the $2^{N}$ vacua which appear for generic values of the
1367: couplings. These are the eigenvalues of $\mf{W}$.  By the R-symmetry,
1368: the values of $W$ at the critical points occur in pairs, $\pm
1369: \lambda$, and are the roots of the characteristic equation,\footnote{
1370:   Note that in general, the number of roots that run off to infinity
1371:   as one approaches the discriminant locus is controlled by the order
1372:   of the pole of $F_{2^{N-1}}(a,b)$. For a pole of order $m$, $2m$
1373:   roots run off to infinity. If the order of the pole is less than the
1374:   maximal ($2^{N-1}$), then some of the roots remain finite, even in
1375:   the limit.}
1376: \begin{equation}
1377:    0 = \det(\tfrac{3}{2}\mf{W} -\lambda \Bid) \equiv P(\lambda^{2}) =
1378:    (\lambda^{2})^{2^{N-1}} - 2^{N-1}\sum_{k=1}^{2^{N-1}}
1379:    F_{k}(a,b)(\lambda^{2})^{2^{N-1}-k} .
1380: \end{equation}
1381: %Note that the normalization factor of $\tfrac{3}{2}$ was thrown in for
1382: %convenience, so that the coefficient of the leading term is 1. 
1383: We can express the $F_k (a,b)$ in terms of the holomorphic moments of
1384: $\mf{W}$ by noting that,
1385: \begin{equation}
1386:    \det(\tfrac{3}{2}\mf{W} -\lambda \Bid) = \exp
1387:    \tr{\log(\tfrac{3}{2}\mf{W} -\lambda \Bid)} = \lambda^{2^{N}}
1388:    \exp{\tr{-\sum_{k=1}^{2^{N-1}} \frac{1}{2k} \left( \frac{3}{2}
1389:    \frac{\mf{W}}{\lambda} \right)^{2k}} } ,
1390: \end{equation}
1391: where the last equality is only meaningful for positive powers of
1392: $\lambda$ and encodes the form of the $F_k (a,b)$.  Dimensional
1393: analysis and invariance under the $S_N\ltimes (\Z_3)^N$ symmetry can
1394: be used to restrict the form of the $F_{k}(a,b)$. They must be
1395: homogeneous polynomials in the $a_{i}$, of degree $3k$, whose
1396: coefficients are rational functions of the $b_{ijk}$. Of particular
1397: importance to us is $F_{1}$, which is proportional to the
1398: ``holomorphic variance'' of $\mf{W}$. Using the symmetries, we may
1399: parameterize this variance as,
1400: \begin{equation}\label{eq:F1def}
1401: \vvev{\tfrac{9}{4}\mf{W}^2} \equiv 2^{-N} \tr{\tfrac{9}{4}\mf{W}^2} = F_{1}(a,b) = f(b)\sum_{i}a_{i}^{3} +\sum_{i<j<k}
1402:  g^{ijk}(b)a_{i}a_{j}a_{k}+\sum_{i\neq j} h^{ij}(b) a_{i}^{2}a_{j},
1403: \end{equation}
1404: where the normalization factor converts the trace into the average
1405: value of $\tfrac{9}{4} W^2$ among the $2^{N}$ critical points.  Now,
1406: as $F_{1}$ must be invariant under $S_{N}\ltimes (\BZ_{3})^N$, we see
1407: that $f(b)$ must also be invariant, while $g^{ijk}(b)$ is symmetric
1408: with vanishing diagonals and must have charge $+1$ under the $i^{th}$,
1409: $j^{th}$, and $k^{th}$ $\BZ_3$, and $h^{ij}(b)$ has vanishing diagonal
1410: and must have charge $-1$ under the $i^{th}$ and $+1$ under the
1411: $j^{th}$ $\BZ_3$. While we will not be able to compute the coefficient
1412: functions $f(b)$, $g^{ijk}(b)$, and $h^{ij}(b)$ in closed form for
1413: arbitrary large $N$, we can certainly do so for any small fixed $N$.
1414: We will now turn to the simple example of the case $N=3$.
1415: 
1416: \subsection{Example I - $W$ odd, $d=3$, $N=3$} \label{subsec:Neq3}
1417: 
1418: Consider the case of the generic, odd, cubic superpotential in three
1419: variables in Fermat form,
1420: \begin{equation}
1421:   \label{eq:wd}
1422:   W= \frac{1}{3} \left( z_1^3 + z_2^3 + z_3^3 \right) -\left( a_1 z_1 + a_2
1423:   z_2 + a_3 z_3 + b z_1 z_2 z_3 \right).
1424: \end{equation}
1425: The equations for its critical points are,
1426: \begin{equation}
1427:   \label{eq:wcrit}
1428:   z_1^2 - a_1 - b z_2 z_3 = 0, ~~~~
1429:   z_2^2 - a_2 - b z_1 z_3 = 0, ~~~~
1430:   z_3^2 - a_3 - b z_1 z_2 = 0.
1431: \end{equation}
1432: A basis for the quotient ring $Q$ is given by the $2^3$ monomials,
1433: \begin{equation}
1434:   \label{eq:basismon}
1435:   \{ 1, z_1, z_2, z_3, z_1 z_2, z_1 z_3, z_2 z_3, z_1 z_2 z_3 \}.
1436: \end{equation}
1437: In this basis, the residue of $W$ in $Q$ is,
1438: \begin{equation}
1439:   \label{eq:wdres}
1440:   W= -\frac{2}{3} \left( a_1 z_1 + a_2
1441:   z_2 + a_3 z_3 \right).
1442: \end{equation}
1443: It is an easy exercise to determine the form of the matrix $\mf{W}$,
1444: \begin{equation}
1445: \mf{W} = -\frac{2}{3} \left(
1446: \begin{array}{cccccccc}
1447: 0 & {a_1}^2 & {a_2}^2 & {a_3}^2 & 0 & 0 & 0 &
1448:   \frac{3 a_1 a_2 a_3 b}{1 - b^3} \\
1449: a_1 & 0 & 0 & 0 & \frac{{a_2}^2 + a_1 a_3 b^2}{1 - b^3} & 
1450:    \frac{{a_3}^2 + a_1 a_2 b^2}{1 - b^3} & \frac{2 a_2 a_3 b}{1 - b^3} & 0\\
1451: a_2 & 0 & 0 & 0 & \frac{{a_1}^2 + a_2 a_3 b^2}{1 - b^3} & \frac{2 a_1 a_3
1452:   b}{1 - b^3} &  \frac{{a_3}^2 + a_1 a_2 b^2}{1 - b^3} & 0\\
1453: a_3 & 0 & 0 & 0 & \frac{2 a_1 a_2 b}{1 - b^3} & \frac{{a_1}^2 + a_2 a_3 b^2}{1 - b^3} & 
1454:    \frac{{a_2}^2 + a_1 a_3 b^2}{1 - b^3} & 0\\
1455: 0 & a_2 & a_1 & a_3 b  & 0 & 0 & 0 & \frac{{a_3}^2 + 2 a_1 a_2
1456:   b^2}{1 - b^3}\\ 
1457: 0 & a_3 & a_2 b  & a_1 & 0 & 0 & 0 & \frac{{a_2}^2 + 2 a_1 a_3 b^2}{1 - b^3}\\
1458: 0 & a_1 b  & a_3 & a_2 & 0 & 0 & 0 & \frac{{a_1}^2 + 2 a_2 a_3 b^2}{1 - b^3}\\
1459: 0 & 0 & 0 & 0 & a_3 & a_2 & a_1 & 0 
1460: \end{array}
1461: \right).
1462: \end{equation}
1463: Now, it is clear that $\tr{\mf{W}} = 0$ as expected from the
1464: R-symmetry. More interesting is the fact that we can just as easily compute,
1465: \begin{equation}
1466:   \label{eq:trwsquared}
1467: \vvev{ \tfrac{9}{4} \mf{W}^2} = 2^{-3}\tr{\tfrac{9}{4}\mf{W}^2} = 
1468:       \left( \frac{(a_1^3 + a_2^3 + a_3^3) \left( 1 -  \frac{1}{4} b^3
1469:       \right) + \frac{9}{2} a_1 a_2 a_3 b^2}{1 - b^3} \right),
1470: \end{equation}
1471: as well as all higher moments of $W$. In particular, we can explicitly
1472: read off the coefficient functions of \eqref{eq:F1def} for the $N=3$ case,
1473: \begin{equation}
1474: \begin{split}
1475: \label{eq:fghN3}
1476:    f(b)&= \frac{1}{4}\frac{4-b^{3}}{1-b^{3}},\\
1477:    g(b)&= \frac{9 b^{2}}{2(1-b^{3})},\\
1478:    h(b)&= 0.
1479: \end{split}
1480: \end{equation}
1481: Further, if we wish to explicitly compute the values of $W$ at the
1482: critical points, we can use these moments to compute the
1483: characteristic polynomial for $W$ and numerically solve it.
1484: Indeed, this is easy to do using Mathematica, and one can verify
1485: that numerical solution of the above equations through other means
1486: agree with the eigenvalue methods. Further, using similar techniques,
1487: we can also compute the moments of any polynomial over these vacua. 
1488: 
1489: Finally, we note that \eqref{eq:fghN3} are singular in the limit that
1490: $b$ approaches a third root of unity. It is easy to check numerically
1491: that six of the eight critical points run off to infinity in this
1492: limit, and that these three points indeed comprise the discriminant
1493: locus of the $N=3$ case. Thus, we see explicitly in this case that the
1494: multi-polynomial resultant we discussed earlier must be proportional to
1495: $(1-b^3)$ for $N=3$.
1496: 
1497: \subsection{Example II - The Fermat Cubic for Large $N$ and small $b_{ijk}$}
1498: \label{subsec:LargeN}
1499: 
1500: Since the dimension of the vector space $Q$ grows exponentially with
1501: the number of variables, the matrix methods introduced above become
1502: quickly intractable, even numerically, for $N \gtrsim 10 $. In
1503: particular, explicit computations require that we can compute
1504: holomorphic moments like
1505: \begin{equation}
1506:   \label{eq:trWsq}
1507:   \vvev{\tfrac{9}{4} \mf{W}^2} = 2^{-N} \tr{\tfrac{9}{4} \mf{W}^2} =
1508:   \sum_{i,j} a_i a_j (2^{-N} \tr{\mf{z_i} \mf{z_j}}) = \sum_{i,j} a_i
1509:   a_j \vvev{\mf{z_i} \mf{z_j}} ,
1510: \end{equation}
1511: which in turn require that we can compute the moments of products of
1512: the $\mf{z_i}$'s. This could be done if we could easily compute the
1513: matrix elements of $\mf{z_i}$ in this basis. However, as we mentioned
1514: earlier, this is very computationally intensive, and involves the
1515: inversion of exponentially large matrices.
1516: 
1517: Thus, it is useful to understand if there exists a limit in which the
1518: above methods become more tractable at large $N$, at least
1519: perturbatively.  Now, note that the decoupled limit with $b_{ijk}=0$
1520: considered in \cite{Arkani-Hamed:2005yv} is certainly such a simple
1521: case, so the $b_{ijk}$ seem to be natural candidates for small
1522: parameters. Indeed, radiative stability of the K\"{a}hler potential
1523: requires that generically $|A_{ijk}| \sim \mcO(N^{-1})$ while at most
1524: $\mcO(N)$ of them may be $\mcO(1)$. That is, the $b_{ijk}$'s are {\it
1525:   generically small}. There are, of course, a large number of them, so
1526: this large-$N$ suppression of the magnitudes of the $b_{ijk}$ is
1527: compensated by the sum. But say that we make the further assumption
1528: that perturbation theory is good --- that is, that the magnitude of
1529: the $b_{ijk}$ is further suppressed by some small parameter $\delta$.
1530: Then it makes sense to expand our expressions for $\vvev{f(\mf{z})}$
1531: as a (formal) power series in the $b_{ijk}$.
1532: 
1533: \subsubsection{Computations of Holomorphic Moments for Small $b_{ijk}$}
1534: 
1535: To begin with, we can use dimensional analysis as well as
1536: the $S_N\ltimes (\Z_3)^N$ symmetry to constrain the form of successive
1537: terms in the power series expansions for $\vvev{f(\mf{z})}$. Let us
1538: consider $\vvev{\mf{z_i} \mf{z_j}}$ as an example. First note that if we
1539: have $i=j$,
1540: \begin{equation}
1541:   \label{eq:trzsq}
1542:   \vvev{\mf{z_i}^2} = \vvev{a_i \Bid + \sum_{j<k} b_{ijk} \mf{z_j}
1543:   \mf{z_k}} = a_i + \sum_{j<k} b_{ijk} \vvev{\mf{z_j} \mf{z_k}},
1544: \end{equation}
1545: so we only need focus on the case $i \neq j$. Using the symmetries and
1546: dimensional analysis, it is easy to check that the first few terms in
1547: the (formal) power series expansion of $\vvev{\mf{z_i}\mf{z_j}}$ for
1548: $i \neq j$ must take the form, 
1549: \begin{equation}
1550:   \label{eq:trztwo}
1551:   \vvev{\mf{z_i} \mf{z_j}} = \a^{(2)}_1 \sum_k b_{ijk}^2 a_k + \sum_{k<l}
1552:         b_{ikl}b_{jkl} \left[ \a^{(3)}_1 ( b_{jkl}a_i + b_{ikl}a_j ) +
1553:         \a^{(3)}_2 ( b_{ijk} a_l + b_{ijl} a_k ) \right] +\CO(b^{4}),
1554: \end{equation}
1555: %The $f_{ijk}$ must obey,
1556: %\begin{equation}
1557: %  f_{ijk}(b)=f_{jik}(b), ~~~ \boldsymbol{\zeta}: f_{ijk}(b) \rightarrow
1558: %  f_{ijk}(b) \zeta_i \zeta_j \zeta_k.
1559: %\end{equation}
1560: %These symmetries constrain the form of the (formal) power series
1561: %expansion of $f_{ijk}(b)$ in the $b$'s,
1562: %\begin{equation}
1563: %  \label{eq:termsinsq}
1564: %  f_{ijk}(b) = \a_1 (b_{ijk})^2 + \a_2 \sum_{m,n}
1565: %  (\delta_{ik}b_{imn}b_{jmn}b_{jmn} + (i \leftrightarrow j)) + \a_3
1566: %  \sum_m b_{ijm} b_{ikm} b_{jkm} + \cdots,
1567: %\end{equation}
1568: where the $\a^{(n)}_i \in \Q$ are rational numbers which we will see
1569: are {\em independent} of $N$ for any fixed $\a^{(n)}_i$ and $N$
1570: sufficiently large. The $\a^{(n)}_i$ can be found by computing the
1571: diagonal matrix elements of the operator $\mf{z_i} \mf{z_j}$ in the
1572: monomial basis, summing them, and dividing by $2^N$. Of course, the
1573: only way one obtains a diagonal element is if the relations
1574: \eqref{eq:critp} are used, so such diagonal elements only arise from
1575: the action of $\mf{z_i} \mf{z_j}$ on basis elements containing $z_i$
1576: or $z_j$. Each use of the relation results either in the introduction
1577: of one factor of $a_i$ or $b_{ijk}$, so the $\a^{(n)}_i$ can be
1578: computed using the relations precisely $(n+1)$ times. For example, the
1579: contribution of the monomial $z_i$ to $\a^{(n)}_1$ is found by,
1580: \begin{equation}
1581: \begin{split}
1582:   \label{eq:matelmts}
1583:   \mf{z_i} \mf{z_j} \cdot z_i & =  a_i z_j + \sum_{k < l \neq j}
1584:   b_{ikl} z_k z_l z_j +  \sum_{l \neq (i,j)} b_{ijl} \mf{z_j} \mf{z_l}
1585:   \cdot z_j  \\ 
1586: & = \cdots + \sum_{l \neq
1587:   (i,j)} b_{ijl} \Bigl( a_j z_l + \sum_{m < n \neq (l,j)}
1588:   b_{jmn} z_m z_n z_l +  \sum_{m \neq (j,l)} b_{jlm} \mf{z_l} \mf{z_m}
1589:   \cdot z_l \Bigr) \\
1590: & = \cdots + \sum_{l \neq (i,j)} \sum_{m \neq (l,j)} b_{ijl} b_{jlm} a_l
1591:   z_m + \mcO(b^3) \\
1592: & = \cdots + \sum_k b_{ijk}^2 a_k z_i + \cdots.
1593: \end{split}
1594: \end{equation}
1595: In fact, we can greatly simplify this computation by noting that the
1596: $S_N$ symmetry implies that we only need to compute how many times any
1597: \emph{single} term of a sum over dummy indices appears in the diagonal
1598: matrix elements. For example, to find $\a^{(2)}_1$, we only need to
1599: ask how many times the single term $b_{ijk}^2 a_k$ with $i$,$j$, and
1600: $k$ all fixed appears on the diagonal of $\mf{z_i} \mf{z_j}$. This is
1601: very easy to do. First of all, this term only involves $a_k$ and one
1602: particular coefficient $b_{ijk}$. As these two coefficients arise in
1603: the relations \eqref{eq:critp} only in terms involving $z_i$, $z_j$,
1604: and $z_k$, we can completely ignore all the other variables, which
1605: just go along for the ride.\footnote{ Formally, this is equivalent to
1606:   working in $Q$ modulo the ideal $\langle z_l \rangle$, $l \neq
1607:   i,j,k$.} Thus, to calculate $\a^{(2)}_1$, we only need to compute
1608: the coefficient of the term $b_{ijk}^2 a_k$ in the trace of $\mf{z_i}
1609: \mf{z_j}$ restricted to the $2^3$ monomials of the form $z_i^{a_1} z_j^{a_2}
1610: z_k^{a_3}$, $a_i = 0,1$ in our basis \eqref{eq:basisN}, ignoring all
1611: terms involving any other variables (and divide the result by $2^{3}$).
1612: That is, we have shown that this coefficient can be computed by
1613: considering just the case $N=3$. For instance, we can expand
1614: \eqref{eq:trwsquared} in a power series in $b$ and find $\a^{(2)}_1 =
1615: \tfrac{3}{4}$.
1616: 
1617: We can more easily compute this using a diagrammatic technique.  To
1618: compute the contribution of $\mf{z_i} \mf{z_j} \cdot z_i^{a_1}
1619: z_j^{a_2} z_k^{a_3}$ to $\a^{(2)}_1$, first draw a labeled external
1620: line for $\mf{z_i}$ and $\mf{z_j}$ as well as all the $z_n$'s present
1621: in the basis element.  Each $b_{ijk}$ corresponds to a vertex which
1622: takes in two identically labeled lines corresponding to one of its
1623: indices and outputs two lines corresponding to the other two indices,
1624: while $a_k$ is an external sink for a pair of $z_k$ lines. Construct
1625: all possible graphs with two $b_{ijk}$ vertices, and one $a_k$
1626: external sink from the given lines, noting that graphs which differ by
1627: a choice of which lines of the same index are contracted at a vertex
1628: or sink are equivalent. Count the number of distinct graphs you drew
1629: and divide by $2^3$ - this is the contribution. It turns out that
1630: there is precisely one possible graph for each basis element with
1631: either $z_i$ or $z_j$ in it, so we again find $\a^{(2)}_1 =
1632: \tfrac{3}{4}$.
1633: 
1634: The above method, of course, generalizes in the obvious way to the
1635: computation of any $\a^{(n)}_i$ by restricting ourselves to just the
1636: relevant $N=n+1$ variables. For example, to compute the coefficient
1637: $\a^{(3)}_1$ of $b_{ikl} b_{jkl}^2 a_i$, we consider $N=4$ and graphs
1638: with one $b_{ikl}$ vertex and two $b_{jkl}$ vertices as well as a
1639: single sink $a_i$. Note that it is useful to start with monomials of
1640: lowest total degree and proceed to those of higher degree as any graph
1641: associated with a given basis monomial also contributes to all
1642: monomials it divides. Thus, without much trouble, one can compute the
1643: terms cubic in $b_{ijk}$,
1644: \begin{equation}
1645: \vvev{\mf{z_{i}}\mf{z_{j}}}= \tfrac{3}{4} \Bigl[  \sum_k b_{ijk}^2 a_k
1646:      + \sum_{k<l} b_{ikl}b_{jkl} \left[
1647:           (b_{jkl}a_i + b_{ikl}a_j) + 2( b_{ijk} a_l + b_{ijl} a_k )
1648:         \right] +\CO(b^{4})
1649:   \Bigr],
1650: \end{equation}
1651: so $\a^{(3)}_1 = \tfrac{3}{4}$, $\a^{(3)}_2 = \tfrac{3}{2}$. Using
1652: this result as well as \eqref{eq:trzsq} we can compute,
1653: \begin{equation}
1654: \begin{split}
1655:   \label{eq:Wsqhol}
1656:   \vvev{\tfrac{9}{4}\mf{W}^2} & = \left( \sum_i a_i^2
1657:   \vvev{\mf{z_i}^2} + 2 \sum_{i<j} a_i a_j \vvev{\mf{z_i} \mf{z_j}}
1658:   \right)
1659:    = \sum_i a_i^3  + \sum_{i<j} \left( 2 a_i a_j + \sum_{l}
1660:   a_l^2 b_{ijl} \right) \vvev{\mf{z_i} \mf{z_j}} \\ 
1661: & = \left(1+\tfrac{3}{4}\sum_{j<k<l}b_{jkl}^{3}\right) \sum_i a_i^3 
1662: + \sum_{i<j<k} \tfrac{9}{2} \left(
1663:   b_{ijk}^2+2\sum_{l}b_{ijl}b_{jkl}b_{ikl} \right) a_i a_j a_k \\
1664:   &\qquad + \sum_{i\neq j} \left( \tfrac{3}{4}
1665:   \sum_{k<l}b_{ikl}b_{jkl}^{2} \right)  a_{i}^{2}a_{j} + \mcO(b^4).
1666: \end{split}
1667: \end{equation}
1668: Thus, we can read off the coefficients functions of
1669: \eqref{eq:F1def} up to terms of order $b^4$, 
1670: \begin{equation}
1671: \begin{split}
1672:   \label{eq:fgh}
1673: f(b) &=1+\tfrac{3}{4}\sum_{j<k<l}b_{jkl}^{3}+\dots, \\ 
1674: g^{ijk}(b) &=  \tfrac{9}{2} \left(
1675:   b_{ijk}^2+2\sum_{l}b_{ijl}b_{jkl}b_{ikl} \right) +\dots, \\
1676: h^{ij}(b) &= \tfrac{3}{4} \sum_{k<l}b_{ikl}b_{jkl}^{2}+\dots.
1677: \end{split}
1678: \end{equation}
1679: As noted, each additional power of $b$ brings an additional sum over a
1680: dummy index, which has $\CO(N)$ terms. If $|b_{ijk}| \sim N^{-1}\d$,
1681: then the term of order $b^{k}$ goes like $N \d^{k}$, so this is a
1682: systematic expansion in powers of $\d$. 
1683: 
1684: We can continue and compute higher holomorphic moments in an expansion
1685: in powers of $b_{ijk}$ - for instance, the following holomorphic
1686: moments will be useful (with $i \neq j \neq k \neq l$),
1687: \begin{equation}\label{eq:quarticmoments}
1688: \begin{split}
1689:    \vvev{\mf{z_{i}}^{4}}-\vvev{\mf{z_{i}}^{2}}^{2} &=
1690:     \sum_{j<k}b_{ijk}^{2}a_{j}a_{k}+\CO(b^{3})\\ 
1691:     \vvev{\mf{z_{i}}^{3}\mf{z_{j}}}
1692:         -\vvev{\mf{z_{i}}^{2}}\vvev{\mf{z_{i}}\mf{z_{j}}} &=
1693:     \sum_{k}b_{ijk}^{2}a_{i}a_{k}+\CO(b^{3})\\ 
1694:     \vvev{\mf{z_{i}}^{2}\mf{z_{j}}^{2}}
1695:         -\vvev{\mf{z_{i}}^{2}}\vvev{\mf{z_{j}}^{2}} &=
1696:     \sum_{k<l}b_{ikl}b_{jkl}a_{k}a_{l} +\CO(b^{3})\\ 
1697:     \vvev{\mf{z_{i}}^{2}\mf{z_{j}}^{2}}
1698:         -\vvev{\mf{z_{i}}\mf{z_{j}}}^{2} &= a_{i} a_{j} +
1699:     \sum_{k<l}b_{ikl}b_{jkl}a_{k}a_{l} +\CO(b^{3}) \\ 
1700:     \vvev{\mf{z_{i}}^{2}\mf{z_{j}}\mf{z_{k}}}
1701:         -\vvev{\mf{z_{i}}^{2}}\vvev{\mf{z_{j}}\mf{z_{k}}} &=
1702:     b_{ijk}a_{j}a_{k} +\CO(b^{3})\\ 
1703:     \vvev{\mf{z_{i}}^{2}\mf{z_{j}}\mf{z_{k}}}
1704:         -\vvev{\mf{z_{i}}\mf{z_{j}}}\vvev{\mf{z_{i}}\mf{z_{k}}} &=
1705:     b_{ijk}a_{j}a_{k} +
1706:     \tfrac{3}{4}\sum_{l}b_{jkl}^{2}a_{i}a_{l}+\CO(b^{3})\\ 
1707:     \vvev{\mf{z_{i}}\mf{z_{j}}\mf{z_{k}}\mf{z_{l}}}
1708:         -\vvev{\mf{z_{i}}\mf{z_{j}}}\vvev{\mf{z_{k}}\mf{z_{l}}} &= \CO(b^{3}).
1709: \end{split}
1710: \end{equation}
1711: Therefore, we see that it is possible to effectively compute
1712: holomorphic moments and any desired holomorphic statistical property
1713: in the limit of large $N$ as a systematic expansion in small
1714: $b_{ijk}$. 
1715: 
1716: For any \emph{fixed} $N$, we can, with more effort, crank out some
1717: explicit formul\ae, capturing the dependence on the discriminant
1718: locus. For $N=3$, we see explicitly, that each of the moments in
1719: \eqref{eq:quarticmoments} has a double pole at the discriminant locus.
1720: \begin{equation}\label{eq:QuarticNeq3Moments}
1721: \begin{split}
1722:    \vvev{\mf{z_{1}}^{4}}-\vvev{\mf{z_{1}}^{2}}^{2} &=
1723:    \frac{ b^{2}(8(2+b^{3})a_{2}a_{3}+3 b^{4}a_{1}^{2})}{16(1-b^{3})^{2}}\\ 
1724:     \vvev{\mf{z_{1}}^{3}\mf{z_{2}}}
1725:         -\vvev{\mf{z_{1}}^{2}}\vvev{\mf{z_{1}}\mf{z_{2}}} &=
1726:     \frac{b^{2}((16-b^{3})a_{1}a_{3}+12b a_{2}^{2})}{16(1-b^{3})^{2}}\\ 
1727:     \vvev{\mf{z_{1}}^{2}\mf{z_{2}}^{2}}
1728:         -\vvev{\mf{z_{1}}^{2}}\vvev{\mf{z_{2}}^{2}} &=
1729:     \frac{b^{3}((16-b^{3})a_{1}a_{2}+12b a_{3}^{2})}{16(1-b^{3})^{2}}\\ 
1730:    \vvev{\mf{z_{1}}^{2}\mf{z_{2}}^{2}}-\vvev{\mf{z_{1}}\mf{z_{2}}}^{2}&=
1731:    \frac{8(2+b^{3})a_{1}a_{2}+3 b^{4}a_{3}^{2}}{16(1-b^{3})^{2}}\\ 
1732:     \vvev{\mf{z_{1}}^{2}\mf{z_{2}}\mf{z_{3}}}
1733:         -\vvev{\mf{z_{1}}^{2}}\vvev{\mf{z_{2}}\mf{z_{3}}} &=
1734:     \frac{b(8(2+b^{3})a_{2}a_{3}+3b^{4} a_{1}^{2})}{16(1-b^{3})^{2}}\\ 
1735:     \vvev{\mf{z_{1}}^{2}\mf{z_{2}}\mf{z_{3}}}
1736:         -\vvev{\mf{z_{1}}\mf{z_{2}}}\vvev{\mf{z_{1}}\mf{z_{3}}} &=
1737:     \frac{b((16-b^{3})a_{2}a_{3}+12b a_{1}^{2})}{16(1-b^{3})^{2}}\\ 
1738: \end{split}
1739: \end{equation}
1740: 
1741: \section{Holomorphic Moments and Scanning}
1742: \label{sec:Scanning}
1743: 
1744: Now, let us consider the relationship between the holomorphic moments
1745: we have computed and the {\em real-valued} statistical properties of
1746: the vacua.  $W=W_{1}+i W_{2}$ is complex. Averaged over the $2^{N}$
1747: vacua, the mean value is, of course, zero.
1748: \begin{equation*}
1749: \vvev{W_{1}}=\vvev{W_{2}}=0
1750: \end{equation*}
1751: The variances involve the sums of squares of the the eigenvalues, and
1752: so are encoded in $\tfrac{4}{9} 2 F_1 = \vvev{\mf{W}^2}$. The variances
1753: we are interested in are $\delta W_{1}^{2}$, $\delta W_{2}^{2}$ and
1754: the covariance, $\vvev{W_{1}W_{2}}$. Two linear combinations of these
1755: are ``holomorphic'' in the eigenvalues,
1756: \begin{subequations}\label{eq:Wvariances}
1757: \begin{align}
1758:    \delta W_{1}^{2}-\delta W_{2}^{2}&= \Real{\vvev{\mf{W}^2}}\\
1759:    \vvev{W_{1}W_{2}}&= \Imag{\vvev{\mf{W}^2}}
1760: \end{align}
1761: The remaining linear combination requires more detailed knowledge,
1762: though it is easy to find a lower bound from the Cauchy-Schwarz
1763: inequality,
1764: \begin{equation}
1765:   \delta W_{1}^{2}+\delta W_{2}^{2} \geq |\vvev{\mf{W}^2}|
1766: \end{equation}
1767: \end{subequations}
1768: In the following, we will turn to the implications of these
1769: formul\ae\ when one averages over some ensemble of couplings. A lower
1770: bound, like (\ref{eq:Wvariances}c) will be quite sufficient to prove
1771: that a given coupling ``scans''. But one requires something of an {\it
1772:   upper} bound in order to prove that a coupling {\it doesn't} scan.
1773: These methods generalize straightforwardly to other coupling which
1774: depend holomorphically on the $\phi_{i}$ (superpotential couplings for
1775: the Standard model fields, holomorphic gauge couplings, {\it etc.}).
1776: 
1777: \subsection{Ensembles of Theories and the Scanning of $\Lambda$}
1778: 
1779: Let us return to \eqref{eq:F1def},  
1780: \begin{equation}
1781:   \label{eq:VarW} 
1782: \vvev{\mf{W}^2} = \tfrac{4}{9} \Bigl[ f(b)\sum_{i}a_{i}^{3} +\sum_{i<j<k}
1783:  g^{ijk}(b)a_{i}a_{j}a_{k}+\sum_{i\neq j} h^{ij}(b) a_{i}^{2}a_{j} \Bigr],
1784: \end{equation}
1785: which encodes the variance of $W$ among the $2^{N}$ vacua for fixed
1786: values of the couplings. What we would like to do now is discuss what
1787: happens if we have some ensemble of theories. For example, rather than
1788: studying F-theory vacua for {\it fixed} flux, we might wish to study
1789: the ensemble of all possible fluxes.
1790: 
1791: No one, currently, has a compelling proposal for what probability
1792: measure to choose for that ensemble. Neither do we. What we hope to do
1793: in this section is to explore the implications one can extract from
1794: such a choice, given the results of the previous sections.
1795: 
1796: The authors of \cite{Arkani-Hamed:2005yv} worked with $b_{ijk}\equiv
1797: 0$, and assumed that the $a_{i}$ were $N$ independent random
1798: variables, chosen from some common distribution. In their limit, the
1799: second two terms of \eqref{eq:VarW} vanish, and we have
1800: \begin{equation*}
1801: \vvev{\mf{W}^2} =\tfrac{4}{9} \sum_{i=1}^{N}a_{i}^{3}
1802: \end{equation*}
1803: Since there are $N$ terms, and each is an independent random variable,
1804: the sum grows like $\CO(N)$. At first glance, for nonzero $b_{ijk}$,
1805: the situation appears to change dramatically. The second term in
1806: \eqref{eq:VarW} contains $\CO(N^{3})$, and the third term contains
1807: $\CO(N^{2})$ independent random variables. At least, that would be the
1808: case if the $b_{ijk}$ coefficients were all generically of $\CO(1)$.
1809: However, as we have argued, radiative stability requires that the
1810: $b_{ijk}$ be generically small.  So, despite appearances, each of the
1811: terms in \eqref{eq:F1def} is $\CO(N)$. The cosmological constant scans
1812: (and, in fact, the variance grows like $\CO(N)$) in this more general
1813: context, just as it did in the model of \cite{Arkani-Hamed:2005yv}
1814: where $b_{ijk}\equiv 0$.
1815: 
1816: \subsection{Other Couplings}
1817: 
1818: Other holomorphic couplings which depend on the $\phi_{i}$ can be
1819: treated similarly. Consider some such coupling, $c(\phi)$, which might
1820: be a holomorphic gauge coupling, or a coupling in the superpotential
1821: for the Standard Model fields. It is reasonable to assume that
1822: $c(\phi)$ has definite parity under the $\BZ_{4}$ R-symmetry that
1823: constrained the form of $W(\phi)$.  The crucial distinction in
1824: \cite{Arkani-Hamed:2005yv} is between those cases where $c(\phi)$ is
1825: odd and hence $\vvev{c(\phi)}=0$ (like the superpotential, $W$) versus
1826: those for which $c(\phi)$ is even. The former couplings ``scan,''
1827: whereas the latter do not: the standard deviation of $c(\phi)$ among
1828: the $2^{N}$ vacua is much smaller that its mean value.
1829: 
1830: As with the superpotential, we assume that $c(\phi)$ has a Taylor
1831: expansion, convergent in a polydisk of radius $M_{c}$. For simplicity,
1832: we will take it to be dimensionless; the case of the $\mu$ parameter
1833: in the Standard Model is an easy generalization. So we have,
1834: \begin{equation}\label{e:cexp}
1835: \begin{split}
1836:   c_{\mathrm{ev}}(\phi) &= c_{0} + \sum_{i\leq j}
1837:   c_{ij}\phi_{i}\phi_{j}/M_{c}^{2}+\dots\\
1838:   c_{\mathrm{odd}}(\phi) &= \sum_{i} c_{i}\phi_{i}/M_{c} +\dots
1839: \end{split}
1840: \end{equation}
1841: depending on the parity of $c(\phi)$.  As in our previous discussion,
1842: radiative stability constrains the form of the coefficients in this
1843: expansion for large-$N$. The $c_{ijk\dots}$ must be {\it generically
1844:   small}. Specifically, we have constraints of the form
1845: \begin{equation}
1846: \begin{split}
1847:    g^{i\overline{\imath}}g^{j \overline{\jmath}}c_{ij}
1848:    \overline{c}_{\overline{\imath} \overline{\jmath}}&\sim \CO(N)\\ 
1849:    \det{}_{i \overline {\imath} }( g^{j \overline{\jmath} } c_{ij}
1850:    \overline{c}_{\overline{\imath}\overline{\jmath}}) & \sim \CO(1)
1851: \end{split}
1852: \end{equation}
1853: and so forth for the higher coefficients. Thus, the mean value of the
1854: $(2k)^{th}$ term in the series \eqref{e:cexp} (the mean value of the
1855: odd terms vanish), rather than going like $N^{2k} (M_{r}/M_{c})^{2k}$
1856: actually goes like $N (M_{r}/M_{c})^{2k}$. So, we have a systematic
1857: expansion in powers of $\epsilon^{2}=(M_{r}/M_{c})^2$. For
1858: sufficiently small $\epsilon$, the leading, $\epsilon^{2}$ term can
1859: compensate for the overall factor of $N$, while the subleading terms
1860: are negligible. More explicitly, if we let $a_i = M_r^2 \tilde{a}_i$
1861: we have,
1862: \begin{subequations}\label{eq:cvev}
1863: \begin{gather}
1864:     \vvev{c_{\mathrm{odd}}}=0\\
1865:     \vvev{c_{\mathrm{ev}}}= c_{0}+ \epsilon^{2}
1866:     \left[f(b)\sum_{i}c_{ii}\tilde{a}_{i}+ \sum_{i\neq j}h^{ij}(b)c_{ii}\tilde{a}_{j}+
1867:     \frac{1}{2}\sum_{i\neq j\neq k} g^{ijk}(b) c_{ij}\tilde{a}_{k}\right]
1868:     +\CO(\epsilon^{4})
1869: \end{gather}
1870: \end{subequations}
1871: where the $\CO(\epsilon^{4})$ terms represent the contributions of
1872: quartic and higher terms in $c(\phi)$. Note that the rational
1873: functions of $b$: $f(b)$, $g^{ijk}(b)$ and $h^{ij}(b)$ are the
1874: \emph{same} ones from \eqref{eq:F1def} that appeared in the
1875: computation of the variances of the superpotential and were computed
1876: up to terms of $\mcO(b^4)$ in \eqref{eq:fgh}. It is easy to check
1877: that, as announced, the leading term behaves as $N\epsilon^{2}$, with
1878: the subleading terms suppressed by higher powers of $\epsilon^{2}$.
1879: 
1880: The variance is calculated similarly. In the odd case,
1881: \begin{equation}
1882:    \vvev{c_{\mathrm{odd}}^{2}} = \epsilon^{2}
1883:    \left[f(b)\sum_{i}c_{i}^{2}\tilde{a}_{i}+ \sum_{i\neq
1884:    j}h^{ij}(b)c_{i}^{2}\tilde{a}_{j}+ \frac{1}{2}\sum_{i\neq j\neq k}
1885:    g^{ijk}(b) c_{i}c_{j}\tilde{a}_{k}\right] +\CO(\epsilon^{4}).
1886: \end{equation}
1887: Thus, given a distribution for the couplings we can definitively show
1888: that this coupling scans if this variance is non-vanishing.
1889: 
1890: In the even case, let $\hat{c} = c - \vvev{c}$. The variance
1891: is then given by,
1892: \begin{equation}\label{eq:cvar}
1893:   \vvev{\hat{c}^{2}} = \sum_{i,j,k,l}
1894:   \frac{c_{ij}c_{kl}}{M_c^4}\left[ \vvev{\mf{z_{i}} \mf{z_{j}}
1895:   \mf{z_{k}} \mf{z_{l}}}
1896:   - \vvev{\mf{z_{i}}\mf{z_{j}}}\vvev{\mf{z_{k}}\mf{z_{l}}} \right]
1897:   +\CO(\epsilon^{6}).
1898: \end{equation}
1899: As in \S\ref{subsec:LargeN}, we can easily compute \eqref{eq:cvar} in
1900: an expansion in powers of $b_{ijk}$ using \eqref{eq:quarticmoments}.
1901: For finite $N$, we can do better -- for $N=3$, we have
1902: \eqref{eq:QuarticNeq3Moments}. As with the superpotential, we can
1903: write $\hat{c} = c_{1}+ic_{2}$, and we have
1904: \begin{subequations}\label{eq:cvars}
1905: \begin{align}
1906:  \delta c_{1}^{2} - \delta c_{2}^{2} &= \Real \vvev{\hat{c}^{2}},\\
1907:  \vvev{c_{1}c_{2}}&= \Imag \vvev{\hat{c}^{2}},\\
1908:  \delta c_{1}^{2} +\delta c_{2}^{2} &\geq |\vvev{\hat{c}^{2}}|.
1909: \end{align}
1910: \end{subequations}
1911: The right-hand-side of (\ref{eq:cvev}b) goes like $N\epsilon^{2}$,
1912: whereas the right-hand-side of \eqref{eq:cvar} goes like
1913: $N\epsilon^{4}$.  We would like to conclude that the standard
1914: deviations, $\delta c_{1}/c_{1}$ and $\delta c_{2}/c_{2}$, behave as
1915: $\tfrac{\sqrt{N}\epsilon^{2}}{N\epsilon^{2}}=\tfrac{1}{\sqrt{N}}$.
1916: 
1917: Unfortunately, (\ref{eq:cvars}c) is only a \emph{lower bound}, so more
1918: detailed information is necessary to \emph{really show} that this
1919: coupling does \emph{not} scan. Indeed, this is a potentially serious drawback to the whole notion of a ``friendly landscape.'' If the lower bound in (\ref{eq:cvars}c) drastically underestimates the true variance, then these couplings \emph{will} vary appreciably over this ensemble of vacua. And anthropic arguments, based on holding them fixed while varying other coupling, like the cosmological constant, are incorrect. 
1920: 
1921: \section{Generalizations}\label{sec:Generalizations}
1922: 
1923: The ``space'' of vacua discussed here is the \emph{complex} affine
1924: algebraic variety, $\C[z_1, \ldots, z_N] / \< \del_i W \>$. As such,
1925: it was amenable to the techniques of complex algebraic geometry. If we
1926: were studying $N$ real scalar fields, $\phi_{i}$, with potential,
1927: $V(\phi)$, we would be faced with a problem in \emph{real} algebraic
1928: geometry. This problem is harder, both because real algebraic geometry
1929: is harder, and less well-developed than the complex case and because
1930: the characterization of the desired space of vacua is more subtle. We
1931: are not interested in $\BR[x_1, \ldots, x_N] / \< \del_i V \>$. We are
1932: only interested in \emph{minima} of $V$, as opposed to all critical
1933: points. At large $N$, ``most'' critical points of $V$ are actually
1934: saddle points, and it's algebraically a little awkward to pick out
1935: just the minima.
1936: 
1937: A more interesting generalization is the case in which some of our
1938: complex chiral multiplets are charged under a $U(1)^{k}$ gauge
1939: symmetry. Physically, this leaves open the possibility of
1940: supersymmetry-breaking, in this ``moduli'' sector, via the inclusion
1941: of Fayet-Iliopoulos terms. Mathematically, this gauging moves us from
1942: the realm of complex affine algebraic algebraic geometry to that of
1943: \emph{toric} geometry. The toric version of our problem is nearly as
1944: well-developed mathematically as the affine case we have discussed. It
1945: would be very interesting to generalize our considerations to that
1946: case.
1947: 
1948: Finally, we have deliberately eschewed discussion of the microphysics
1949: that determines the (ensemble of) couplings in our low-energy
1950: effective Lagrangian. Douglas and collaborators \cite{ Douglas:2003um,
1951:   Ashok:2003gk, Douglas:2004kp, Douglas:2004zu, Denef:2004ze,
1952:   Douglas:2004qg, Douglas:2004kc, Douglas:2004zg, Denef:2004cf,
1953:   Douglas:2005df}, for instance, have pursued the idea of treating all
1954: possible choices of fluxes in an F-theory compactification
1955: ``democratically,'' assigning equal weight to the low-energy theory
1956: that arises from each choice of flux. Further developments along this
1957: vein appear in \cite{Giryavets:2004zr, DeWolfe:2004ns,
1958:   Blumenhagen:2004xx, Misra:2004ky, Conlon:2004ds, Kumar:2004pv,
1959:   Acharya:2005ez}.  It has been argued \cite{DeWolfe:2005uu} that in
1960: the context of type IIA flux compactifications this may not be a
1961: reasonable choice, as there are instances where the number of possible
1962: choices of flux is infinite. Independent of this more subtle question,
1963: we believe that our methods will prove useful in analyzing the vacuum
1964: structure of any give theory \emph{whenever} one has a large number,
1965: $N$, of light chiral multiplets.
1966: 
1967: \section{Acknowledgments}
1968: 
1969: We would like to thank N. Arkani-Hamed, A. Bergman, G. Farkas, W.
1970: Fischler, S. Keel, S. Paban, B. Sturmfels, J. Vaaler, and C. Vafa for
1971: discussions. JD would like to thank the Harvard Theory Group and the
1972: Perimeter Institute for hospitality while this work was in progress.
1973: 
1974: This material is based upon work supported by the National Science
1975: Foundation under Grant Nos. PHY-0071512 and PHY-0455649, and by grant
1976: support from the US Navy, Office of Naval Research, Grant Nos.
1977: N00014-03-1-0639 and N00014-04-1-0336, Quantum Optics Initiative.
1978: 
1979: 
1980: %\appendix
1981: %\section{The Nullstellensatz and the O'Raifeartaigh Mechanism}
1982: %\label{app:Null}
1983: 
1984: \bibliographystyle{utphys}
1985: \bibliography{poly}
1986: \end{document}
1987: