math0307321/CU.tex
1: \documentclass[times,10pt,twocolumn]{article}
2: \usepackage{latex8}
3: \usepackage{times}
4: \usepackage{amsmath,amsthm,amsfonts,amssymb,amscd}
5: \usepackage{fancyhdr}
6: \input epsf
7: 
8: \newtheorem{lemma}{Lemma}
9: \newtheorem{theorem}[lemma]{Theorem}
10: \newtheorem{corollary}[lemma]{Corollary}
11: \newtheorem{definition}{Definition}
12: \newtheorem{proposition}[lemma]{Proposition}
13: \newtheorem{question}{Question}
14: \numberwithin{lemma}{section}
15: \numberwithin{definition}{section}
16: \numberwithin{question}{section}
17: \newcommand\C{{\mathbb{C}}}
18: \newcommand\R{{\mathbb{R}}}
19: \newcommand\Quat{{\mathbb{H}}}
20: \newcommand\M{{\mathbb{M}}}
21: \newcommand\F{{\mathbb{F}}}
22: \newcommand\Z{{\mathbb{Z}}}
23: \newcommand\Tr{{\mathop\textup{Tr }}}
24: 
25: \setcounter{page}{438}
26: \pagestyle{fancy}
27: \fancyhead{}
28: \renewcommand{\headrulewidth}{0pt}
29: \fancyfoot[CE,CO]{\thepage}
30: 
31: \begin{document}
32: 
33: \title{A Group-theoretic Approach to Fast Matrix Multiplication}
34: 
35: \author{Henry Cohn\\
36: Microsoft Research\\
37: One Microsoft Way\\
38: Redmond, WA 98052-6399\\
39: cohn@microsoft.com\\
40: \and
41: Christopher Umans\\
42: Department of Computer Science\\
43: California Institute of Technology\\
44: Pasadena, CA 91125\\
45: umans@cs.caltech.edu\\
46: }
47: 
48: \maketitle
49: 
50: \begin{abstract}
51: We develop a new, group-theoretic approach to bounding the
52: exponent of matrix multiplication. There are two components to
53: this approach: (1) identifying groups $G$ that admit a certain
54: type of embedding of matrix multiplication into the group algebra
55: $\C[G]$, and (2) controlling the dimensions of the irreducible
56: representations of such groups. We present machinery and examples
57: to support (1), including a proof that certain families of groups
58: of order $n^{2 + o(1)}$ support $n \times n$ matrix
59: multiplication, a necessary condition for the approach to yield
60: exponent $2$.  Although we cannot yet completely achieve both (1)
61: and (2), we hope that it may be possible, and we suggest potential
62: routes to that result using the constructions in this paper.
63: \end{abstract}
64: 
65: \Section{Introduction}
66: 
67: \thispagestyle{fancy}
68: 
69: \fancyfoot[LE,LO]{\ \\ \parbox{6.875in}{\scriptsize{\ \\ Copyright
70: \copyright\ 2003 IEEE. Reprinted from Proceedings of the 44th
71: Annual Symposium on Foundations of Computer Science. This
72: material is posted here with permission of the IEEE.  Such
73: permission of the IEEE does not in any way imply IEEE endorsement
74: of any of Cornell University's products or services.  Internal or
75: personal use of this material is permitted.  However, permission
76: to reprint/republish this material for advertising or promotional
77: purposes or for creating new collective works for resale or
78: redistribution must be obtained from the IEEE by writing to
79: pubs-permissions@ieee.org.  By choosing to view this document,
80: you agree to all provisions of the copyright laws protecting
81: it.}}}
82: 
83: Strassen \cite{S} made the startling discovery that one can
84: multiply two $n \times n$ matrices in only $O(n^{2.81})$ field
85: operations, compared with $2n^3$ for the standard algorithm.  This
86: immediately raises the question of the exponent of matrix
87: multiplication: what is the smallest number $\omega$ such that for
88: each $\varepsilon>0$, matrix multiplication can be carried out in
89: at most $O(n^{\omega+\varepsilon})$ operations? Clearly $\omega
90: \ge 2$.  It is widely believed that $\omega=2$, but the best bound
91: known is $\omega < 2.38$, due to Coppersmith and Winograd
92: \cite{CW}, following a sequence of improvements to Strassen's
93: original algorithm (see \cite[p.~420]{BCS} for the history). It
94: is known that all the standard linear algebra problems (for
95: example, computing determinants, solving systems of equations,
96: inverting matrices, computing LUP decompositions---see Chapter~16
97: of \cite{BCS}) have the same exponent as matrix multiplication,
98: which makes $\omega$ a fundamental number for understanding
99: algorithmic linear algebra. In addition, there are non-algebraic
100: algorithms whose complexity is expressed in terms of $\omega$
101: (see, e.g., Section~16.9 in \cite{BCS}).
102: 
103: Several fairly elaborate techniques for bounding $\omega$ are
104: known, but since 1990 nobody has been able to improve on them. In
105: this paper:
106: \begin{itemize}
107: \item We develop a new approach to bounding $\omega$ that imports the
108: problem into the domain of group theory and representation theory.
109: The approach is relatively simple and almost entirely separate
110: {}from the existing machinery built up since Strassen's original
111: algorithm.
112: 
113: \item We demonstrate the feasibility of the group theory aspect of the
114: approach by identifying a family of groups for which a parameter
115: that mirrors $\omega$ approaches 2. We also exhibit techniques
116: for bounding this critical parameter and prove non-trivial bounds
117: for a number of diverse groups and group families.
118: 
119: \item We pose a question in representation theory (Question~\ref{fundamentalq}
120: below) that represents a potential barrier to directly obtaining
121: non-trivial bounds on $\omega$ using this approach. We do not
122: know the answer to this question. A positive answer would
123: illuminate a path that might lead to $\omega = 2$ using the
124: techniques that we present in this paper.
125: \end{itemize}
126: 
127: Our approach is reminiscent of a question asked by Coppersmith
128: and Winograd (in Section~11 of \cite{CW}) about avoiding ``three
129: disjoint equivoluminous subsets'' in abelian groups, which would
130: lead to $\omega=2$ if it has a positive answer. However, our
131: technique is completely different, and our framework seems to
132: have more algebraic structure to make use of (whereas theirs is
133: more combinatorial).
134: 
135: \SubSection{Analogy with fast polynomial multiplication}
136: 
137: There is a close analogy between the framework we propose in this
138: paper and the well-known algorithm for multiplying two degree $n$
139: polynomials in $O(n\log n)$ operations using the Fast Fourier
140: Transform (FFT). In this section we elucidate this analogy to give
141: a high-level description of our technique.
142: 
143: Suppose we wish to multiply the polynomials $A(x) = \sum_{i =
144: 0}^{n-1} a_i x^i$ and $B(x) = \sum_{i = 0}^{n-1} b_i x^i$. The
145: naive way to do this is to compute $n^2$ products of the form
146: $a_ib_j$, and from these the $2n-1$ coefficients of the product
147: polynomial $A(x)\cdot B(x)$. Of course a far better algorithm is
148: possible; we describe it below in language that easily translates
149: into our framework for matrix multiplication.
150: 
151: \fancyfoot[CE,CO]{\thepage}
152: \fancyfoot[LE,LO]{}
153: 
154: Let $G$ be a group and let $\C[G]$ be the group algebra---that is,
155: every element of $\C[G]$ is a formal sum $\sum_{g \in G} a_g g$
156: with $a_g \in \C$, and the product of two such elements is
157: $$
158: \left(\sum_{g \in G}a_g g \right) \cdot \left ( \vphantom{\sum_{g
159: \in G}a_g g} \sum_{h \in G}b_h h \right ) = \sum_{f \in G}\left
160: (\sum_{gh = f} a_gb_{h} \right ) f.
161: $$
162: 
163: We often identify the element $\sum_{g \in G} a_g g$ with the
164: vector of its coefficients. If $G$ is the cyclic group of order
165: $m$, then the product of two elements $a = (a_g)_{g \in G}$ and
166: $b = (b_g)_{g \in G}$ is a {\em cyclic convolution} of the
167: vectors $a$ and $b$. The important observation is that a cyclic
168: convolution is almost what is needed to compute the coefficients
169: of the product polynomial $A(x)\cdot B(x)$---the only problem is
170: that it wraps around.  To avoid this problem, we embed $A(x)$ and
171: $B(x)$ as elements $\bar{A}, \bar{B} \in \C[G]$ as follows: Let
172: $z$ be a generator of $G$, which we assume to be a cyclic group
173: of order $m > 2n-1$, and define
174: \[
175: \bar{A} = \sum_{i = 0}^{n-1}a_iz^i \qquad \textup{and}\qquad
176: \bar{B} = \sum_{i = 0}^{n-1}b_iz^i.
177: \]
178: Since the group size $m$ is large enough to avoid wrapping
179: around, we can read off the coefficients of the product
180: polynomial from the element $\bar{A}\bar{B} \in \C[G]$: the
181: coefficient of $x^i$ in $A(x)B(x)$ is the coefficient of the
182: group element $z^i$ in $\bar{A}\bar{B}$. This is a wordy account
183: of a so-far simple correspondence, but the payoff is near. The
184: {\em Discrete Fourier Transform} (DFT) for $\C[G]$ is an
185: invertible linear transformation $D:\C[G] \rightarrow \C^{|G|}$,
186: which turns multiplication in $\C[G]$ into pointwise
187: multiplication of vectors in $\C^{|G|}$. We can therefore compute
188: the product $\bar{A}\bar{B}$ by first computing $D(\bar{A})$ and
189: $D(\bar{B})$ and then computing the inverse DFT of their
190: pointwise product. Thus, using the $O(m \log m)$ Fast Fourier
191: Transform algorithm, we can perform multiplication in $\C[G]$
192: (and therefore polynomial multiplication, via the embedding
193: above) in $O(m \log m)$ operations.
194: 
195: One of the main results of the present paper is that {\em matrix
196: multiplication can be embedded into group algebra multiplication
197: in an analogous way}. The embedding is not as simple as the
198: embedding of polynomial multiplication, but it has a natural and
199: clean description in terms of a property of subsets of $G$ (which
200: we often take to be subgroups). In particular, if $S, T$, and $U$
201: are subsets of $G$ and $A = (a_{s, t})_{s \in S, t \in T}$ and $B
202: = (b_{t, u})_{t \in T, u \in U}$ are $|S| \times |T|$ and $|T|
203: \times |U|$ matrices, respectively, then we define
204: \[\bar{A} = \sum a_{s, t}s^{-1}t \qquad\textup{and}\qquad \bar{B}
205: = \sum b_{t, u} t^{-1}u.\] If $S, T, U$ satisfy the {\em triple
206: product property} (see Definition~\ref{definition:realize}), then
207: we can read off the entries of the product matrix $AB$ from
208: $\bar{A}\bar{B} \in \C[G]$: entry $(AB)_{s, u}$ is simply the
209: coefficient of the group element $s^{-1}u$.
210: 
211: In the case of polynomial multiplication, the simplicity of the
212: embedding obscures the fact that if $G$ is too large (e.g., if
213: $|G| = n^2$ rather than $O(n)$), then the benefit of the entire
214: scheme is destroyed. Avoiding this pitfall turns out to be the
215: main challenge in the new setting. We wish to embed matrix
216: multiplication into a group algebra over a {\em small} group $G$,
217: as the size of $G$ is a lower bound on the complexity of
218: multiplication in $\C[G]$. It is not surprising, for example,
219: that $n \times n$ matrix multiplication can be embedded into the
220: group algebra of a group of order $n^3$. We show that abelian
221: groups cannot beat $n^3$ and {\em we identify families of
222: non-abelian groups of size $n^{2 + o(1)}$ that admit such an
223: embedding.}
224: 
225: It might seem that this result together with the above trick for
226: performing group algebra multiplication (i.e., taking the DFT,
227: multiplying in the Fourier domain, and transforming back) would
228: imply that $\omega = 2$. There are, however, two complications
229: introduced by the fact that we are forced to work with non-abelian
230: groups. The first is that we know of fast algorithms to compute
231: the DFT only for limited classes of non-abelian groups (see
232: Section~13.5 in \cite{BCS}). However, the DFT is linear, and
233: because of the recursive structure of divide and conquer matrix
234: multiplication algorithms, linear transformations applied before
235: and after the recursive step are ``free.'' For example, in
236: Strassen's original matrix multiplication algorithm, the number
237: of matrix additions and scalar multiplications in the recursive
238: step does not affect the bound on $\omega$. So this potential
239: complication is in fact no problem at all.
240: 
241: The second complication is that for $\C[G]$ when $G$ is
242: non-abelian, multiplication in the Fourier domain is {\em not}
243: simply pointwise multiplication of vectors in $\C^{|G|}$. Instead
244: it is {\em block-diagonal matrix multiplication}, where the
245: dimensions of the blocks are the dimensions of the irreducible
246: representations of $G$. We thus obtain a reduction of $n \times n$
247: matrix multiplication to a number of smaller matrix
248: multiplications of varying sizes, which gives rise to an
249: inequality involving the exponent $\omega$ of matrix
250: multiplication. If the size of $G$ were exactly $n^2$, then this
251: inequality would imply that $\omega = 2$. However, the smallest
252: one can make $|G|$ is $n^{2 + o(1)}$, and then the question of
253: whether the inequality implies $\omega = 2$ turns on the
254: representation theory of $G$. We show that when $|G| = n^{2 +
255: o(1)}$, even slight control over the dimension of the largest
256: irreducible representation is sufficient to achieve $\omega = 2$.
257: Some control is necessary to avoid trivialities such as reducing
258: to an even larger matrix multiplication problem. We can achieve
259: that much control; the issue of whether it is possible to achieve
260: more control is the subject of Question~\ref{fundamentalq}.
261: 
262: \SubSection{Outline}
263: 
264: Following some preliminaries below,
265: Sections~\ref{section:realizing} through~\ref{section:bounds} are
266: devoted to outlining our approach. In
267: Sections~\ref{section:linear} and~\ref{variety}, we show that a
268: variety of different types of groups support matrix multiplication
269: within our framework, and in the process demonstrate a number of
270: useful proof techniques. Section~\ref{section:linear} highlights
271: linear groups, whose representation theory makes them especially
272: attractive for our purposes. Section~\ref{section:Lie} describes
273: a parallel with Lie groups and gives a construction that suggests
274: that finite linear groups may indeed be a fruitful line of
275: inquiry. In Section~\ref{section:wreath} we consider wreath
276: product constructions, and in Section~\ref{section:direct} we use
277: the combinatorial notion of Sperner capacity to demonstrate the
278: surprising fact that the $k$-fold direct product of a group may
279: support $n^k \times n^k$ matrix multiplication even when the group
280: itself fails to support $n \times n$ matrix multiplication. This
281: suggests a potential route to answering
282: Question~\ref{fundamentalq} in the affirmative. We end by
283: mentioning some open problems and variants of our overall
284: approach in Section~\ref{section:conclusions}.
285: 
286: \SubSection{Preliminaries}
287: 
288: Let $\langle n, m, p \rangle$ denote the structural tensor for
289: rectangular matrix multiplication of $n \times m$ by $m \times p$
290: matrices, and let $R$ denote the tensor rank function. (See
291: \cite{BCS} for background on matrix multiplication and tensors.
292: We will use this material only in the proof of
293: Theorem~\ref{theorem:bound}.)  We will typically work over the
294: field of complex numbers; if we use another field $F$, we will
295: write $\langle n, m, p \rangle_F$. As usual $\omega$ will denote
296: the exponent of matrix multiplication over $\C$.
297: 
298: We will use the following basic fact from representation theory:
299: the group algebra $\C[G]$ of a finite group $G$ decomposes as the
300: direct product
301: $$
302: \C[G] \cong  \C^{d_1 \times d_1} \times \dots \times \C^{d_k
303: \times d_k}
304: $$
305: of matrix algebras of orders $d_1,\dots,d_k$.  These numbers are
306: called the character degrees of $G$, or the dimensions of the
307: irreducible representations.  It follows from computing the
308: dimensions of both sides that $|G| = \sum_i d_i^2$.  See
309: \cite{JL} and \cite{H} for background on representation theory.
310: 
311: \Section{Realizing matrix multiplication via groups}
312: \label{section:realizing}
313: 
314: In this section we describe the embedding of matrix multiplication
315: into group algebra multiplication, and we identify a property of
316: groups $G$ that implies that the group algebra of $G$ admits such
317: an embedding. If $S$ is a subset of a group, let $Q(S)$ denote
318: the right quotient set of $S$, i.e.,
319: $$
320: Q(S) = \{s_1 s_2^{-1} : s_1,s_2 \in S\}.
321: $$
322: 
323: \begin{definition}
324: \label{definition:realize} A group $G$ {\em realizes} $\langle
325: n_1,n_2,n_3 \rangle$ if there are subsets $S_1,S_2,S_3 \subseteq
326: G$ such that $|S_i| = n_i$, and for $q_i \in Q(S_i)$, if
327: $$
328: q_1q_2q_3 = 1
329: $$
330: then $q_1=q_2=q_3=1$. We call this condition on $S_1,S_2,S_3$ the
331: {\em triple product property}. If we wish to emphasize the
332: specific subsets, we say that $G$ {\em realizes $\langle
333: n_1,n_2,n_3 \rangle$ through} $S_1,S_2,S_3$.
334: \end{definition}
335: 
336: In most of our examples, matrix multiplication will be realized
337: through subgroups $H_1$, $H_2$, $H_3$ of $G$, rather than
338: arbitrary subsets.  In that case, the triple product property is
339: especially simple, because $Q(H_i) = H_i$: it states that if
340: $h_1h_2h_3=1$ with $h_i \in H_i$, then $h_1=h_2=h_3=1$.  An
341: equivalent formulation replaces $h_1h_2h_3=1$ with $h_1h_2=h_3$.
342: 
343: Perhaps the simplest example comes from the product $C_{n} \times
344: C_m \times C_p$ of cyclic groups, which clearly realizes $\langle
345: n,m,p \rangle$ through $C_n \times \{1\} \times \{1\}$, $\{1\}
346: \times C_m \times \{1\}$, and $\{1\} \times \{1\} \times C_p$.  We
347: will see a number of less trivial examples shortly.
348: 
349: \begin{lemma}
350: \label{lemma:permute} If $G$ realizes $\langle
351: n_1,n_2,n_3\rangle$, then it does so for every permutation of
352: $n_1,n_2,n_3$.
353: \end{lemma}
354: 
355: \begin{proof}
356: Suppose $G$ realizes $\langle n_1,n_2,n_3\rangle$ through
357: $S_1,S_2,S_3$, and suppose $s_i,s_i' \in S_i$. We need to show
358: that the order in which $1$, $2$, and $3$ appear in the equation
359: $$
360: s_1's_1^{-1} s_2's_2^{-1} s_3's_3^{-1} = 1
361: $$
362: is irrelevant.  Conjugating by $s_1' s_1^{-1}$ shows that it is
363: equivalent to
364: $$
365: s_2's_2^{-1} s_3's_3^{-1} s_1's_1^{-1} = 1,
366: $$
367: so we can perform a cyclic shift.  To get a transposition, we
368: take the inverse of the initial equation, which yields
369: $$
370: s_3s_3'^{-1} s_2s_2'^{-1} s_1s_1'^{-1} = 1,
371: $$
372: i.e., a transposition of $1$ with $3$ (the roles of $s$ and $s'$
373: have been reversed, but that is irrelevant).  These two
374: permutations generate all permutations of $\{1,2,3\}$.
375: \end{proof}
376: 
377: \begin{lemma}
378: \label{lemma:shortexact} If $N$ is a normal subgroup of $G$ that
379: realizes $\langle n_1, n_2, n_3\rangle$ and $G/N$ realizes
380: $\langle m_1, m_2, m_3\rangle$, then $G$ realizes $\langle
381: n_1m_1, n_2m_2, n_3m_3\rangle$.
382: \end{lemma}
383: 
384: \begin{proof}
385: Suppose $N$ realizes $\langle n_1, n_2, n_3\rangle$ through
386: $S_1,S_2,S_3$, and suppose $T_1,T_2,T_3$ are lifts to $G$ of the
387: three subsets of $G/N$ that realize $\langle m_1, m_2,
388: m_3\rangle$. Then we claim that $G$ realizes $\langle n_1m_1,
389: n_2m_2, n_3m_3\rangle$ through the pointwise products
390: $S_1T_1,S_2T_2,S_3T_3$.  We need to check that for $s_i,s_i' \in
391: S_i$ and $t_i,t_i' \in T_i$,
392: $$
393: (s_1't_1')(s_1t_1)^{-1} (s_2't_2')(s_2t_2)^{-1}
394: (s_3't_3')(s_3t_3)^{-1} = 1
395: $$
396: iff $s_i=s_i'$ and $t_i=t_i'$ for all $i$.  If we reduce this
397: equation modulo $N$, we find that $t_i=t_i'$ modulo $N$, and
398: hence also in $G$.  The equation in $G$ then becomes
399: $$
400: s_1's_1^{-1} s_2's_2^{-1} s_3's_3^{-1} = 1,
401: $$
402: {}from which we deduce $s_i=s_i'$, as desired.
403: \end{proof}
404: 
405: One useful special case of Lemma~\ref{lemma:shortexact} is that if
406: $G_1$ realizes $\langle n_1, m_1, p_1\rangle$ and $G_2$ realizes
407: $\langle n_2, m_2, p_2\rangle$, then $G_1 \times G_2$ realizes
408: $\langle n_1n_2, m_1m_2, p_1p_2\rangle$.
409: 
410: Our first theorem describes the embedding of matrix multiplication
411: into group algebra multiplication:
412: 
413: \begin{theorem}
414: \label{theorem:reduction} Let $F$ be any field.  If $G$ realizes
415: $\langle n,m,p \rangle$, then the number of field operations
416: required to multiply $n \times m$ with $m \times p$ matrices over
417: $F$ is at most the number of operations required to multiply two
418: elements of $F[G]$.  Furthermore, $\langle n,m,p\rangle_F \le
419: F[G]$.
420: \end{theorem}
421: 
422: For the definition of the restriction relation $\le$ in the last
423: sentence, see Section~14.3 of \cite{BCS}.
424: 
425: \begin{proof}
426: Let $G$ realize $\langle n,m,p \rangle$ through subsets $S,T,U$.
427: Suppose $A$ is an $n \times m$ matrix, and $B$ is an $m \times p$
428: matrix.  We will index the rows and columns of $A$ with the sets
429: $S$ and $T$, respectively, those of $B$ with $T$ and $U$, and
430: those of $AB$ with $S$ and $U$.
431: 
432: Consider the product
433: $$
434: \left(\sum_{s\in S, t \in T} A_{st} s^{-1} t\right)
435: \left(\sum_{t' \in T, u \in U} B_{t'u} t'^{-1} u\right)
436: $$
437: in the group algebra.  We have
438: $$
439: (s^{-1} t) (t'^{-1} u) = s'^{-1} u'
440: $$
441: iff $s=s'$, $t=t'$, and $u=u'$, so the coefficient of $s^{-1} u$
442: in the product is
443: $$
444: \sum_{t \in T} A_{st} B_{tu} = (AB)_{su}.
445: $$
446: Thus, one can simply read off the matrix product from the group
447: algebra product by looking at the coefficients of $s^{-1} u$ with
448: $s \in S, u \in U$, and the assertions in the theorem statement
449: follow.
450: \end{proof}
451: 
452: \Section{The pseudo-exponent}
453: 
454: The {\em pseudo-exponent} of a group measures the quality of the
455: embedding afforded by Theorem \ref{theorem:reduction} in a single,
456: well-behaved parameter, which in some ways mirrors the exponent
457: $\omega$ of matrix multiplication.
458: 
459: \begin{definition}
460: The pseudo-exponent $\alpha(G)$ of a non-trivial finite group $G$
461: is the minimum of
462: $$
463: \frac{3 \log |G|}{\log nmp}
464: $$
465: over all $n,m,p$ (not all $1$) such that $G$ realizes $\langle
466: n,m,p \rangle$.  The pseudo-exponent of the trivial group is $3$.
467: \end{definition}
468: 
469: When it is clear from the context which group is intended, we
470: often write $\alpha$ instead of $\alpha(G)$.  Note that in the
471: special case that $G$ realizes $\langle n, n, n \rangle$, its
472: pseudo-exponent satisfies $\alpha \le \log_n |G|$.  In general,
473: if $G$ realizes $\langle n,m,p \rangle$, then
474: $$
475: \alpha \le \log_{\sqrt[3]{nmp}} |G|.
476: $$
477: 
478: \begin{lemma}
479: The pseudo-exponent of a finite group $G$ is always greater than
480: $2$ and at most $3$. If $G$ is abelian, then it is exactly $3$.
481: \end{lemma}
482: 
483: \begin{proof}
484: The upper bound of $3$ is trivial: use the subgroups $H_1=H_2 =
485: \{1\}$ and $H_3=G$.
486: 
487: For the lower bounds, suppose $G$ realizes $\langle n_1,n_2,n_3
488: \rangle$ (with $n_1n_2n_3>1$) through subsets $S_1,S_2,S_3$.  It
489: follows from the definition of realization that the map $(x,y)
490: \mapsto x^{-1} y$ is injective on $S_1 \times S_2$ and its image
491: intersects the quotient set $Q(S_3)$ only in the identity.  Thus,
492: $|G| \ge n_1n_2$, and $|G| > n_1n_2$ unless $n_3=1$.  Similarly,
493: $|G| \ge n_2n_3$ with equality only if $n_1=1$, and $|G| \ge
494: n_1n_3$ with equality only if $n_2=1$. Thus, $|G|^3 >
495: (n_1n_2n_3)^2$, so $\alpha(G) > 2$.
496: 
497: If $G$ is abelian, then the product map $S_1 \times S_2 \times S_3
498: \to G$ must be injective, so $|G| \ge n_1n_2n_3$ and $\alpha(G)
499: \ge 3$.
500: \end{proof}
501: 
502: The pseudo-exponent is well-behaved with respect to group
503: extensions:
504: 
505: \begin{lemma}
506: \label{lemma:upperbound} If $N$ is a normal subgroup of $G$, then
507: $\alpha(G) \le \max(\alpha(N),\alpha(G/N))$.
508: \end{lemma}
509: 
510: \begin{proof}
511: Suppose $N$ realizes $\langle n_1, n_2, n_3\rangle$ and $G/N$
512: realizes $\langle m_1, m_2, m_3\rangle$.  Then
513: Lemma~\ref{lemma:shortexact} implies that the pseudo-exponent of
514: $G$ is at most
515: $$
516: \frac{3\log |G|}{\log n_1m_1n_2m_2n_3m_3} = \frac{3\log |N| +
517: 3\log |G/N| }{\log n_1n_2n_3 + \log m_1m_2m_3},
518: $$
519: which is bounded above by the larger of
520: $$
521: \frac{3\log |N|}{\log n_1n_2n_3} \qquad \textup{and}\qquad
522: \frac{3\log |G/N|}{\log m_1m_2m_3},
523: $$
524: as desired.
525: \end{proof}
526: 
527: Non-abelian groups can have pseudo-exponent less than $3$.  The
528: smallest example is the symmetric group $S_3$ on $3$ elements. It
529: realizes $\langle 2, 2, 2 \rangle$ through its three subgroups of
530: order $2$, so it has pseudo-exponent at most $\log_2 6$ (and one
531: can check that it is exactly $\log_2 6$). Next, we generalize
532: this construction to show that it is possible to come arbitrarily
533: close to pseudo-exponent $2$, as follows.
534: 
535: Given a triangular array of points in the plane, as in
536: Figure~\ref{fig-triangle}, we consider the group of permutations
537: of the points, together with three subgroups, one for each side
538: of the triangle.  Each subgroup permutes the set of points on
539: each line parallel to its side of the triangle.  The proof of
540: Theorem~\ref{theorem:pseudoexponent2}, while not phrased in
541: geometric terms, shows that these subgroups satisfy the triple
542: product property.
543: 
544: \begin{figure}
545: \begin{center}
546: \leavevmode \epsfbox[-2 -2 82 92]{triangle.ps}
547: \end{center}
548: \caption{A triangular array of points.} \label{fig-triangle}
549: \end{figure}
550: 
551: \begin{theorem}
552: \label{theorem:pseudoexponent2} The pseudo-exponent of
553: $S_{n(n+1)/2}$ is at most
554: $$
555: 2 +\frac{2-\log 2}{\log n} + O\left(\frac{1}{(\log n)^2}\right).
556: $$
557: \end{theorem}
558: 
559: \begin{proof}
560: There are $n(n+1)/2$ triples $(a,b,c)$ with $a,b,c \ge 0$ and
561: $a+b+c=n-1$.  We view $S_{n(n+1)/2}$ as the group of permutations
562: of these triples.  Let $H_i$ be the subgroup that fixes the $i$-th
563: coordinate.  The size of this subgroup is $1!2!\dots n!$, so the
564: pseudo-exponent bound is
565: $$
566: \frac{\log (n(n+1)/2)!}{\log 1!2!\dots n!} = 2 + \frac{2-\log
567: 2}{\log n} + O\left(\frac{1}{(\log n)^2}\right),
568: $$
569: assuming these subgroups satisfy the triple product property.  For
570: that, we need to prove that if $h_1h_2h_3=1$ with $h_i \in H_i$,
571: then $h_1=h_2=h_3=1$.
572: 
573: Suppose $h_1h_2h_3=1$ with $h_i \in H_i$.   We will order the
574: triples lexicographically, so that $(0,0,n-1)$ is the smallest
575: triple and $(n-1,0,0)$ is the largest, and prove by induction
576: using this ordering that $h_1$, $h_2$, and $h_3$ fix every triple.
577: 
578: Suppose all triples smaller than $(a,b,c)$ are fixed by each of
579: $h_1, h_2, h_3$ (in the base case, the set of such triples is
580: empty). The permutation $h_3$ cannot send $(a,b,c)$ to a smaller
581: triple, since all smaller triples are fixed points, so $h_3$ must
582: send it to $(a+i,b-i,c)$ with $i \ge 0$.  Then $h_2$ sends that
583: to $(a+i+j,b-i,c-j)$ for some $j$. The only way $h_1$ can return
584: to $(a,b,c)$ is if $i+j=0$, so that must be the case. However,
585: $h_1$ fixes $(a,b-i,c+i)$ for $i>0$ (since such a triple is
586: smaller than $(a,b,c)$), so we must have $i=0$.  It follows that
587: $(a,b,c)$ is fixed by each of $h_1,h_2,h_3$, so by induction all
588: triples are fixed and hence $h_1=h_2=h_3=1$.
589: \end{proof}
590: 
591: The same holds for all symmetric groups, since one can look at the
592: largest subgroup of the form $S_{n(n+1)/2}$.
593: 
594: \Section{Relating the pseudo-exponent to $\omega$}
595: \label{section:bounds}
596: 
597: In this section we relate the pseudo-exponent $\alpha$ to the
598: exponent of matrix multiplication $\omega$.  As with many of the
599: results since Strassen's algorithm, our main theorems are stated
600: as bounds on $\omega$, rather than explicit algorithms, but of
601: course algorithms are implicit in the proofs.
602: 
603: \begin{theorem}
604: \label{theorem:bound} Suppose $G$ has pseudo-exponent $\alpha$,
605: and the character degrees of $G$ are $\{d_i\}$. Then
606: $$
607: |G|^{\omega/\alpha} \le \sum_{i} d_i^{\omega}.
608: $$
609: \end{theorem}
610: 
611: The intuition is simple: the problem of multiplying matrices of
612: size $|G|^{1/\alpha}$ reduces to multiplication in $\C[G]$, which
613: is equivalent to multiplying a collection of matrices of sizes
614: $d_i$.  These multiplications should take about $d_i^\omega$
615: operations, so $\sum_i d_i^\omega$ should be an approximate upper
616: bound for the number of operations required to multiply matrices
617: of size $|G|^{1/\alpha}$, i.e., roughly $|G|^{\omega/\alpha}$. It
618: is convenient that when one makes this idea precise, these crude
619: approximations become exact bounds.
620: 
621: \begin{proof}
622: Suppose $G$ realizes $\langle n, m, p \rangle$ with $nmp =
623: |G|^{3/\alpha}$ (it follows from the definition of the
624: pseudo-exponent that $G$ realizes such a tensor). By
625: Theorem~\ref{theorem:reduction},
626: \begin{equation}
627: \label{equation:reduction} \langle n, m, p \rangle \le \C[G]
628: \simeq \bigoplus_i \langle d_i, d_i, d_i \rangle.
629: \end{equation}
630: We will need two facts about the rank of matrix multiplication:
631: for all $n',m',p'$,
632: $$
633: (n'm'p')^{\omega/3} \le R(\langle n', m', p' \rangle)
634: $$
635: (Proposition~15.5 in \cite{BCS}), and for each $\varepsilon>0$
636: there exists $C>0$ such that for all $k$,
637: $$
638: R(\langle k, k, k \rangle) \le C k^{\omega+\varepsilon}
639: $$
640: (Proposition~15.1 in \cite{BCS}).
641: 
642: The $\ell$-th tensor power of \eqref{equation:reduction} is
643: $$
644: \langle n^\ell, m^\ell, p^\ell \rangle \le
645: \bigoplus_{i_1,\dots,i_\ell} \langle d_{i_1}\dots
646: d_{i_\ell},d_{i_1}\dots d_{i_\ell}, d_{i_1}\dots d_{i_\ell}
647: \rangle,
648: $$
649: if we use
650: $$
651: \langle n_1,m_1,p_1\rangle \otimes \langle n_2,m_2,p_2\rangle
652: \simeq \langle n_1n_2,m_1m_2,p_1p_2 \rangle.
653: $$
654: It follows from taking the rank of both sides that
655: $$
656: |G|^{\ell\omega/\alpha} \le C \left(\sum_i
657: d_i^{\omega+\varepsilon}\right)^\ell,
658: $$
659: and if we take the $\ell$-th root and let $\ell$ go to infinity,
660: then we deduce that
661: $$
662: |G|^{\omega/\alpha} \le \sum_{i} d_i^{\omega+\varepsilon}.
663: $$
664: Finally, because this inequality holds for all $\varepsilon>0$, it
665: must hold for $\varepsilon=0$ as well, by continuity.
666: \end{proof}
667: 
668: Notice that if $\alpha(G)$ were $2$, then this theorem would imply
669: that $\omega=2$ (using $\sum_i d_i^2 = |G|$, the Cauchy-Schwarz
670: inequality, and the fact that every non-trivial group has at least
671: two irreducible representations). In general, though, we need to
672: control the character degrees of $G$. The maximum possible
673: character degree for any non-trivial group is $(|G|-1)^{1/2}$; we
674: show below that an upper bound of $|G|^{1/2 - \varepsilon}$ for
675: fixed $\varepsilon > 0$ would be sufficient to obtain $\omega =
676: 2$ from a family of groups with pseudo-exponent approaching $2$
677: (and that even a much weaker bound suffices).
678: 
679: We define $\gamma(G)$, or simply $\gamma$ when $G$ is clear from
680: the context, so that $|G|^{1/\gamma}$ is the maximum character
681: degree of $G$ ($\gamma(G)=\infty$ if $G$ is abelian). Ideally,
682: we'd like the exponent of matrix multiplication $\omega$ to be
683: bounded above by the pseudo-exponent $\alpha$. The following
684: corollary shows that in the region near 2, this actually happens,
685: with a correction factor that depends on $\gamma$.
686: 
687: \begin{corollary}
688: \label{corollary:upperbound} Let $G$ be a finite group. If
689: $\alpha(G) < \gamma(G)$, then
690: $$
691: \omega \le \alpha \left (\frac{\gamma - 2}{\gamma - \alpha} \right
692: ).
693: $$
694: \end{corollary}
695: 
696: \begin{proof}
697: Let $\{d_i\}$ denote the character degrees.  Then by
698: Theorem~\ref{theorem:bound},
699: \begin{eqnarray*}
700: |G|^{\omega/\alpha} &\le& \sum_i d_i^{\omega -2}d_i^2\\
701: &\le& |G|^{(\omega - 2)/\gamma}\sum_i d_i^2\\
702: & = & |G|^{1+(\omega - 2)/\gamma},
703: \end{eqnarray*}
704: which implies $\omega(1/\alpha-1/\gamma) \le 1-2/\gamma.$
705: Dividing by $1/\alpha - 1/\gamma$ (which is positive by
706: assumption) yields the stated result.
707: \end{proof}
708: 
709: Like $\alpha(G)$, we have $\gamma(G)>2$ for all $G$, and
710: Corollary~\ref{corollary:upperbound} shows that our approach
711: amounts to a race between $\alpha(G)$ and $\gamma(G)$ to see
712: which approaches $2$ faster. The most attractive form of this
713: corollary is the following special case:
714: 
715: \begin{corollary}
716: \label{cor:race} Suppose there exists a family $G_1,G_2,\dots$ of
717: finite groups such that $\alpha(G_i) = 2+o(1)$ as $i \to \infty$,
718: and furthermore $\alpha(G_i)-2 = o(\gamma(G_i)-2)$.  Then the
719: exponent of matrix multiplication is $2$.
720: \end{corollary}
721: 
722: These corollaries are weakenings of Theorem~\ref{theorem:bound},
723: the advantage being that they only require knowledge of
724: $\gamma(G)$, which is typically easier to work with than the
725: complete set of character degrees that is required for
726: Theorem~\ref{theorem:bound}.
727: 
728: It is reasonable to ask whether the requirement $\alpha < \gamma$
729: which occurs in Corollary~\ref{corollary:upperbound} is
730: necessary. It turns out that it is, because if $\alpha \ge
731: \gamma$, then for all $\omega>0,$
732: $$
733: |G|^{\omega/\alpha} \le |G|^{\omega/\gamma} \le \sum_i d_i^\omega,
734: $$
735: where the second inequality holds because $|G|^{1/\gamma} = d_i$
736: for some $i$. Then the inequality in Theorem~\ref{theorem:bound}
737: holds even for $\omega = 3$. The necessity of $\alpha < \gamma$
738: makes perfect sense, because when it fails to hold, the approach
739: amounts to a reduction of matrix multiplication to several
740: instances, one of which is as large as the original instance. In
741: fact, the construction in the proof of
742: Theorem~\ref{theorem:pseudoexponent2} succumbs to this problem:
743: there we proved that $\alpha(S_{n(n+1)/2}) \le 2+O(1/\log n)$,
744: but it turns out that $\gamma(S_{n(n+1)/2}) = 2 + \Theta(1/(n\log
745: n))$ (see \cite{VK}).  However, there exist non-abelian groups
746: for which $\alpha < \gamma$ and $\alpha < 3$; one example is the
747: group in Proposition~\ref{proposition:order80} below.
748: 
749: If we {\em do} have access to the complete set of character
750: degrees then there is a relatively simple condition to check to
751: determine whether the inequality in Theorem~\ref{theorem:bound}
752: yields a non-trivial bound on $\omega$. The condition is that
753: $|G|^{3/\alpha} > \sum_i{d_i^3}$. To see this observe that the
754: inequality in Theorem~\ref{theorem:bound} is equivalent to
755: \begin{equation} \label{eqn:log}
756: \frac{\omega}{\alpha} \log |G| \le \log \sum_i d_i^\omega.
757: \end{equation}
758: The right-hand side is convex as a function of $\omega$, and the
759: left-hand side is linear. Furthermore, as $\omega \to \infty$,
760: the right-hand side is asymptotic to
761: $$
762: \frac{\omega}{\gamma} \log |G|,
763: $$
764: which is smaller than the left-hand side when $\alpha < \gamma$
765: (which is the non-trivial case).  Therefore \eqref{eqn:log} gives
766: no information about $\omega$ in the interval $[2,3]$ unless it
767: rules out $\omega=3$, which is equivalent to the above stated
768: condition. We do not have examples of groups meeting this
769: condition.
770: 
771: We are thus led to pose the following question in representation
772: theory:
773: 
774: \begin{question}
775: \label{fundamentalq} Does there exist a finite group that realizes
776: $\langle n,m,p \rangle$ and has character degrees $\{d_i\}$ such
777: that
778: $$
779: nmp > \sum_i d_i^3?
780: $$
781: \end{question}
782: 
783: It is possible that there is a theorem in representation theory
784: that implies that the answer to this question is ``no.'' In that
785: case the approach we have outlined cannot be used directly to
786: obtain bounds on $\omega$; however even in this case there are
787: variants of our approach that would not be ruled out (see, e.g.,
788: Subsection~\ref{section:extensions}). On the other hand, a
789: positive answer might point the direction to a proof that $\omega
790: = 2$ using our approach: it would seem strange if the best bound
791: groups could prove were some constant strictly between~$2$
792: and~$3$, and the condition in Corollary~\ref{cor:race} for
793: $\omega=2$ feels very natural.
794: 
795: \Section{Linear groups} \label{section:linear}
796: 
797: Matrix groups over finite fields are an important class of finite
798: groups. They are especially attractive for our purposes because
799: their representation sizes, as measured by $\gamma$, are well
800: behaved. We will focus on the case of $SL_n(\F_q)$ for simplicity,
801: although we see no reason why it should perform better than other
802: linear groups.  If $n>1$ is held fixed, $\gamma(SL_n(\F_q))$
803: approaches $2+2/n$ as $q$ tends to infinity (which can be deduced
804: from \cite{Green}, according to a private
805: communication from G.\ Lusztig). Thus, if one could prove that
806: $\alpha(SL_n(\F_q)) = 2+o(1)$ for some fixed $n$, then
807: Corollary~\ref{corollary:upperbound} would imply $\omega=2$. Even
808: if one lets $n$ grow, one might still hope that $\alpha$ would
809: tend to $2$ faster than $\gamma$.  We cannot prove that $\alpha$
810: even approaches $2$ at all as $n,q \to \infty$, but comparison
811: with Theorem~\ref{theorem:slnLie} below suggests that it does. In
812: this section we concentrate on the case of $SL_2(\F_q)$.
813: 
814: For later reference, we collect here the character degrees of
815: $SL_2(\F_q)$:
816: \begin{table}[h]
817: \begin{tabular}{c|c|c}
818: Degree & Multiplicity ($q$ odd) & Multiplicity ($q$ even)\\ \hline
819: $q+1$ & $(q-3)/2$ & $(q-2)/2$\\
820: $q$ & $1$ & $1$\\
821: $q-1$ & $(q-1)/2$ & $q/2$\\
822: $(q+1)/2$ & $2$ & $0$\\
823: $(q-1)/2$ & $2$ & $0$\\
824: $1$ & $1$ & $1$
825: \end{tabular}
826: \end{table}
827: 
828: (See Exercise~28.2 and its solution in \cite{JL} for $q$ even, and
829: \cite{LR} for $q$ odd, but note that \cite{LR} has a typo in the
830: multiplicity for degree $q+1$ at the bottom of the first column
831: on page~122.)
832: 
833: \begin{proposition}
834: \label{proposition:sl2} The group $SL_2(\F_q)$ of order $q^3-q$
835: realizes $\langle q, q, q \rangle$.
836: \end{proposition}
837: 
838: Unfortunately, this pseudo-exponent bound tends to $3$ as $q \to
839: \infty$, but at least it is always strictly better than $3$.  (We
840: can also prove similarly that $\alpha(SL_n(\F_q)) < 3$.)
841: 
842: \begin{proof}
843: Consider the three parabolic subgroups
844: $$
845: H_1 = \left\{ \left(
846: \begin{array}{cc}
847: 1 & x\\
848: 0 & 1 \end{array} \right) : x \in \F_q\right\},
849: $$
850: $$
851: H_2 = \left\{ \left(
852: \begin{array}{cc}
853: 1 & 0\\
854: y & 1 \end{array} \right) : y \in \F_q\right\},
855: $$
856: and
857: $$
858: H_3 = \left\{ \left(
859: \begin{array}{cc}
860: 1+z & z\\
861: -z & 1-z \end{array} \right) : z \in \F_q\right\}.
862: $$
863: We need to check that for $h_i \in H_i$, if $h_1h_2=h_3$, then
864: $h_1=h_2=h_3=1$. To check that, we multiply to get
865: $$
866: \left(
867: \begin{array}{cc} 1 & x\\ 0 & 1
868: \end{array} \right)\left( \begin{array}{cc} 1 & 0\\ y & 1
869: \end{array} \right) = \left( \begin{array}{cc}
870: 1+xy & x\\
871: y & 1 \end{array} \right).
872: $$
873: That can be of the form
874: $$
875: \left(
876: \begin{array}{cc}
877: 1+z & z\\
878: -z & 1-z \end{array} \right)
879: $$
880: only if $x=y=z=0$, as desired.
881: \end{proof}
882: 
883: One might hope that $SL_n(\F_q)$ realizes
884: $$
885: \langle q^{n(n-1)/2}, q^{n(n-1)/2}, q^{n(n-1)/2}\rangle
886: $$
887: through three conjugates of the group of upper-triangular matrices
888: with $1$'s on the diagonal. However, that fails for $q=2$ and
889: $n=3$, according to calculations using the computer program GAP
890: (see~\cite{GAP}); furthermore, no subgroups of these orders work
891: for $q=2$ and $n=3$.
892: 
893: \begin{proposition}
894: \label{proposition:sl2fq2} The group $SL_2(\F_{q^2})$ of order
895: $q^6-q^2$ realizes $\langle q^2, q^2, q^3-q \rangle$.
896: \end{proposition}
897: 
898: \begin{proof}
899: Let $x \mapsto \bar x$ denote the Frobenius automorphism of
900: $\F_{q^2}$ over $\F_q$.  The three subgroups we will use are
901: $$
902: H_1 = \left\{ \left(
903: \begin{array}{cc}
904: 1 & x\\
905: 0 & 1 \end{array} \right) : x \in \F_{q^2}\right\},
906: $$
907: $$
908: H_2 = \left\{ \left(
909: \begin{array}{cc}
910: 1 & 0\\
911: y & 1 \end{array} \right) : y \in \F_{q^2}\right\},
912: $$
913: and
914: \begin{eqnarray*}
915: H_3 &=& SU_2(\F_q)\\
916: &=& \left\{ \left(
917: \begin{array}{cc}
918: a & b\\
919: -\bar b & \bar a \end{array} \right) : a,b \in \F_{q^2}, a\bar a
920: + b\bar b = 1\right\}.
921: \end{eqnarray*}
922: Note that to check that $|H_3| = q^3-q$, one just needs to count
923: solutions to $a\bar a + b\bar b = 1$. For a fixed $b$ with $b\bar
924: b \ne 1$, there are $q+1$ corresponding choices of $a$ that work;
925: if $b\bar b =1$, then $a=0$.  There are $(q^2-1)-(q+1)$ non-zero
926: choices of $b$ with $b \bar b \ne 1$ (to which we must add
927: $b=0$), and $q+1$ with $b\bar b = 1$.  Thus, there are
928: $(q^2-q-1)(q+1)+(q+1) = q^3-q$ elements of $H_3$.
929: 
930: As in the previous proof, checking the triple product property
931: amounts to checking that
932: $$
933: \left( \begin{array}{cc}
934: 1+xy & x\\
935: y & 1 \end{array} \right) = \left(
936: \begin{array}{cc}
937: a & b\\
938: -\bar b & \bar a \end{array} \right)
939: $$
940: implies $x=y=b=0$ and $a=1$, which is a trivial calculation.
941: \end{proof}
942: 
943: Proposition~\ref{proposition:sl2fq2} proves that
944: $$
945: \liminf_{q \to \infty} \alpha(SL_2(\F_q)) \le 18/7,
946: $$
947: which is substantially better than~$3$ but still not near $2$.
948: Using Theorem~\ref{theorem:bound} and the character degrees of
949: $SL_2(\F_q)$, one can show that if
950: $$
951: \liminf_{q \to \infty} \alpha(SL_2(\F_q)) < 9/4,
952: $$
953: then Question~\ref{fundamentalq} has a positive answer.
954: 
955: \Section{Lie groups} \label{section:Lie}
956: 
957: In the category of Lie groups, one can set up a theory parallel to
958: that of the previous sections. We do not know how to use it to
959: bound the exponent of matrix multiplication (because of course
960: Lie groups of positive dimension are infinite). However, we have
961: had more luck constructing examples using Lie groups than with
962: finite linear groups, and this success seems a good reason to be
963: optimistic about matrix groups over finite fields. All examples
964: involving Lie groups can be skipped by a reader who cares only
965: about finite groups and matrix multiplication.
966: 
967: Recall that $Q(S)$ denotes the right quotient set of $S$.
968: 
969: \begin{definition}
970: \label{definition:Liepseudoexponent}  Let $G$ be a Lie group,
971: with submanifolds $M_1, M_2, M_3$ such that for $q_i \in Q(M_i)$,
972: if $q_1q_2q_3=1$ then $q_1=q_2=q_3=1$. We say that $G$ has {\em
973: Lie pseudo-exponent} at most
974: $$
975: \frac{\dim(G)}{(\dim(M_1)+\dim(M_2)+\dim(M_3))/3}.
976: $$
977: \end{definition}
978: 
979: We usually take the submanifolds to be Lie subgroups.  If $G$ and
980: the three subgroups are algebraic groups defined over a number
981: field, then it is natural to ask what pseudo-exponent may be
982: achieved when one reduces modulo a prime ideal, to get a finite
983: quotient group. If the triple product property still holds, then
984: as the finite field size tends to infinity, the pseudo-exponent
985: bound of this finite group approaches the Lie pseudo-exponent.
986: However, the triple product property may not be preserved, as we
987: will show after the following theorem.
988: 
989: \begin{theorem}
990: \label{theorem:slnLie} The group $SL_n(\R)$ has Lie
991: pseudo-exponent at most $2+2/n$.
992: \end{theorem}
993: 
994: \begin{proof}
995: The three subgroups are the group $U$ of upper-triangular
996: matrices with $1$'s on the diagonal, the group $L$ of
997: lower-triangular matrices with $1$'s on the diagonal, and
998: $SO_n(\R)$.  Each subgroup has dimension $n(n-1)/2$, and
999: $SL_n(\R)$ has dimension $n^2-1$, so assuming the triple product
1000: property holds, the Lie pseudo-exponent is at most
1001: $$
1002: \frac{n^2-1}{n(n-1)/2} = 2+\frac{2}{n}.
1003: $$
1004: 
1005: Let $M \in SO_n(\R)$, $A \in U$, and $B \in L$.  We wish to prove
1006: that if $MA=B$, then $M=A=B=I$.  Let $e_1,\dots,e_n$ be the
1007: standard basis of $\R^n$.  We will prove by induction on $i$ that
1008: $Me_i=e_i$.  Once we know that $M=I$, it follows that $A=B$, and
1009: thus $A=B=I$ because $U$ and $L$ are disjoint except for the
1010: identity.  ($A=B=I$ will also follow directly from the proof that
1011: $M=I$.)
1012: 
1013: Let $A_i$ and $B_i$ denote the $i$-th columns of $A$ and $B$, and
1014: denote their $j$-th entries by $A_{ij}$ and $B_{ij}$.  Note that
1015: this indexing of rows and columns is opposite to the standard
1016: convention, but it will be more convenient in this proof.
1017: Because $MA=B$, we have
1018: $$
1019: M A_i = B_i.
1020: $$
1021: 
1022: We start with the base case $i=1$.  Since $A$ is in $U$, we have
1023: $A_1 = e_1$.  Thus, $|B_1| = |MA_1| = |Me_1| = |e_1| = 1$, since
1024: $M$ is an orthogonal matrix.  Because $B_{11} = 1$, the only way
1025: $|B_1|$ can be $1$ is if $B_1=e_1$.  Thus, $Me_1 = e_1$.
1026: 
1027: Now suppose that $Me_j = e_j$ for all $j<i$.  Because $A$ is in
1028: $U$,
1029: $$
1030: A_i = e_i + \sum_{j < i} A_{ij} e_j,
1031: $$
1032: and because $B$ is in $L$,
1033: $$
1034: B_i = e_i + \sum_{j > i} B_{ij} e_j.
1035: $$
1036: Now the induction hypothesis implies that
1037: $$
1038: B_i = MA_i = Me_i + \sum_{j < i} A_{ij} e_j,
1039: $$
1040: so
1041: $$
1042: Me_i = e_i +\sum_{j > i} B_{ij} e_j - \sum_{j < i} A_{ij}e_j.
1043: $$
1044: Since $M$ is orthogonal, $|Me_i|=|e_i|=1$.  The coefficient of
1045: $e_i$ in $Me_i$ is already $1$, so the other coefficients must be
1046: zero and thus $Me_i=e_i$, as desired.
1047: \end{proof}
1048: 
1049: The same holds for $SL_n(\C)$ with $SO_n(\R)$ replaced by $SU_n$,
1050: but not by $SO_n(\C)$: the orthogonal matrix
1051: $$
1052: \left(\begin{array}{ccc} 1 & \frac{-1+i}{2} & \frac{1+i}{2}\\
1053: 1 & \frac{1+i}{2} & \frac{-1+i}{2}\\
1054: -i & 1 & 1
1055: \end{array}\right)
1056: $$
1057: equals
1058: $$
1059: \left(\begin{array}{ccc} 1 & 0 & 0\\
1060: 1 & 1 & 0\\
1061: -i & \frac{1-i}{2} & 1
1062: \end{array}\right)
1063: \left(\begin{array}{ccc} 1 & \frac{-1+i}{2} & \frac{1+i}{2}\\
1064: 0 & 1 & -1\\
1065: 0 & 0 & 1
1066: \end{array}\right).
1067: $$
1068: Of course the same obstacle arises over finite fields (a sum of
1069: non-zero squares may vanish).
1070: 
1071: \Section{Additional examples} \label{variety}
1072: 
1073: In this section we explore a variety of different types of groups,
1074: and prove non-trivial pseudo-exponent bounds for them. We hope
1075: that these examples (together with the ones we have already seen)
1076: will serve as something of a tool kit for constructing a group
1077: that might answer Question~\ref{fundamentalq}, and possibly even a
1078: family of groups that prove $\omega = 2$.
1079: 
1080: \SubSection{Solvable groups} \label{section:solvable}
1081: 
1082: Non-abelian simple (or almost simple) groups appear to be a
1083: fruitful source of groups with small pseudo-exponents.  However,
1084: solvable groups also do quite well. In this section, we will
1085: construct solvable groups that have Lie pseudo-exponent tending
1086: to $2$, and finite solvable groups with pseudo-exponent bounds of
1087: $2.5$ and $2.4811\dots$ (which, GAP tells us, is the best
1088: pseudo-exponent attained using three subgroups in any group of
1089: order up to~$100$).
1090: 
1091: Let $F$ be a field, and $\langle,\rangle$ a symmetric bilinear
1092: form on $F^n$.  Define multiplication in
1093: $$
1094: G = \{(x,y,\alpha): x,y \in F^n, \alpha \in F \}
1095: $$
1096: via
1097: $$
1098: (x,y,\alpha) (u,v,\beta) = (x+u,y+v,\alpha+\beta+2\langle u, y
1099: \rangle),
1100: $$
1101: and define the three subgroups
1102: $$
1103: H_1 = \{(x,0,0): x \in F^n\},
1104: $$
1105: $$
1106: H_2 = \{(0,y,0): y \in F^n\},
1107: $$
1108: and
1109: $$
1110: H_3 = \{(z,z,\langle z,z \rangle): z \in F^n\}.
1111: $$
1112: 
1113: \begin{proposition}
1114: If the only element $z \in F^n$ satisfying $\langle z,z \rangle =
1115: 0$ is $z=0$, then $H_1$, $H_2$, and $H_3$ satisfy the triple
1116: product property. \label{prop:inner}
1117: \end{proposition}
1118: 
1119: \begin{proof}
1120: We simply need to check that $H_3$ avoids all elements of the form
1121: $(x,0,0)(0,y,0) = (x,y,0)$, except when $x=y=0$.  The only way
1122: such an element can be in $H_3$, i.e., of the form $(z,z,\langle
1123: z,z \rangle)$, is if $x=y=z$ and $\langle z,z \rangle = 0$.  That
1124: means $z=0$ and thus $x=y=0$, as desired.
1125: \end{proof}
1126: 
1127: When $F = \R$, the group described above is a Heisenberg group,
1128: and we obtain the following bound:
1129: 
1130: \begin{corollary}
1131: In the above framework, with $F = \R$, and $\langle,\rangle$ the
1132: standard inner product, the Lie group $G$ has Lie pseudo-exponent
1133: at most $2 + 1/n$.
1134: \end{corollary}
1135: 
1136: \begin{proof}
1137: It is clear that Proposition \ref{prop:inner} is satisfied; the
1138: group dimension is $2n + 1$, and the three subgroups each have
1139: dimension $n$.
1140: \end{proof}
1141: 
1142: When $F$ is a finite field, the group described above is an
1143: extraspecial group, and we obtain the following bound:
1144: 
1145: \begin{corollary}
1146: In the above framework, with $F = \F_q$ of odd characteristic, $n
1147: = 2$, and $\langle x, y \rangle = x_1y_1 - wx_2y_2$ for some $w
1148: \in F$ that is not a square, the finite group $G$ has
1149: pseudo-exponent at most $2.5$.
1150: \end{corollary}
1151: 
1152: Here, $x_i$ denotes the $i$-th coordinate of the vector $x$.
1153: 
1154: \begin{proof}
1155: Note that $\langle z, z \rangle = 0$ implies $z_1^2 = w z_2^2$,
1156: which by our choice of $w$ can only happen when $z = 0$. Thus
1157: Proposition \ref{prop:inner} is satisfied. The group has order
1158: $q^5$, and the three subgroups have size $q^2$, leading to a
1159: pseudo-exponent bound of $2.5$ as claimed.
1160: \end{proof}
1161: 
1162: A slight variant of this construction works for even $q$ as well,
1163: but the pseudo-exponent bound is identical so we omit the details.
1164: 
1165: One quite different example is the following Frobenius group of
1166: order $80$.  We found the group by a brute force search using GAP,
1167: and Michael Aschbacher supplied the following humanly
1168: understandable proof that it works.
1169: 
1170: Let $C_5 \subset \F_{16}^\times$ be the unique subgroup of order
1171: $5$. Consider its semidirect product $G = C_5 \ltimes \F_{16}$
1172: with the additive group of $\F_{16}$, where multiplication is
1173: defined by
1174: $$
1175: (\alpha,x)(\beta,y) = (\alpha\beta, \beta x + y).
1176: $$
1177: 
1178: \begin{proposition}
1179: \label{proposition:order80} The group $G = C_5 \ltimes \F_{16}$
1180: realizes $\langle 5, 5, 8\rangle$, and thus $\alpha(G) \le
1181: 3\log_{200}80 = 2.4811\dots$.
1182: \end{proposition}
1183: 
1184: \begin{proof}
1185: Let
1186: $$
1187: H_1 = \{(\alpha,0) : \alpha \in C_5\}
1188: $$
1189: and
1190: $$
1191: H_2 = \{(\alpha, \alpha-1) : \alpha \in C_5\}
1192: $$
1193: (i.e., $H_2$ is $H_1$ conjugated by $(1,1)$). Let
1194: $$
1195: H_3 = \{(1,x) : x \in \F_{16}, \Tr x = 0 \},
1196: $$
1197: where $\textup{Tr}$ denotes the trace from $\F_{16}$ to $\F_2$.
1198: These groups satisfy $|H_1| = |H_2| = 5$ and $|H_3| = 8$.  All we
1199: need to check is the triple product property.
1200: 
1201: We must verify that unless $\alpha$ and $\beta$ are both $1$, the
1202: product
1203: $$
1204: (\alpha,0)(\beta, \beta-1) = (\alpha\beta, \beta-1)
1205: $$
1206: is not in $H_3$.  For it to be in $H_3$, we must have $\alpha =
1207: \beta^{-1}$ and $\Tr (\beta-1) = 0$.  However,
1208: $$
1209: \Tr (\beta - 1) = \Tr \beta - \Tr 1 = \Tr \beta,
1210: $$
1211: and $\Tr \beta = 1$ for $\beta \in C\setminus\{1\}$ because the
1212: minimal polynomial over $\F_2$ of such a $\beta$ is
1213: $1+\beta+\beta^2+\beta^3+\beta^4$.
1214: \end{proof}
1215: 
1216: This proposition generalizes as follows (see \cite{Brown} for
1217: background on cohomology): Let $G$ be a group that acts on an
1218: abelian group $A$, $\theta : G \to A$ a $1$-cocycle, and $B
1219: \subseteq A$ a subgroup. If $\theta(g) \in B$ implies $g=1$ for
1220: all $g \in G$, then the semidirect product $G \ltimes A$ realizes
1221: $\langle |G|,|G|,|B|\rangle$ via the subgroups $G \times \{0\}$,
1222: $\{(g,\theta(g)) : g \in G\}$, and $\{1\} \times B$.  (In
1223: Proposition~\ref{proposition:order80}, the $1$-cocycle is a
1224: coboundary.) Unfortunately, we do not know any other good
1225: examples.
1226: 
1227: Unlike the cases of extraspecial groups and matrix groups, we do
1228: not know how to generalize Proposition~\ref{proposition:order80}
1229: to achieve Lie pseudo-exponent arbitrarily near $2$. The best we
1230: know how to do is the following. Let $\Quat$ be the quaternions,
1231: and $U \subset \Quat^\times$ be the group of unit quaternions
1232: (which is isomorphic to $SU(2)$). Then within the semidirect
1233: product $U \ltimes \Quat$, the three subgroups $U \times \{0\}$,
1234: $\{(u,u-1) : u \in U\}$, and $\{(0,x): \Tr x = 0 \}$ satisfy the
1235: triple product property and prove that the Lie pseudo-exponent of
1236: $U \ltimes \Quat$ is at most $7/3$.
1237: 
1238: \SubSection{Wreath products} \label{section:wreath}
1239: 
1240: In this section we present another family of groups that achieves
1241: pseudo-exponent $2 + o(1)$. This family is described in terms of
1242: the wreath product: if $A$ is a group, then the wreath product $A
1243: \wr S_n$ is the semidirect product $S_n \ltimes A^n$, where $S_n$
1244: acts on $A^n$ by permuting the coordinates (and the
1245: multiplication is of course via $(\pi, u)(\pi',v) = (\pi\pi',
1246: \pi'u + v)$).
1247: 
1248: \begin{theorem}
1249: Let $A$ be the cyclic group of order $2n$, and let $G_n = A\wr
1250: S_n$. Then
1251: $$
1252: \alpha(G_n) \le \gamma(G_n) = 2 + \frac{1+\log 2}{\log n} +
1253: O\left(\frac{1}{(\log n)^2}\right).
1254: $$
1255: \end{theorem}
1256: 
1257: \begin{proof}
1258: We view $G_n$ as the semidirect product $S_n \ltimes A^n$, and
1259: will use the three subgroups
1260: \begin{eqnarray*}
1261: H_1 & = & \{(\pi, 0) : \pi \in S_n\}, \\
1262: H_2 & = & \{(\pi, \pi u - u) : \pi \in S_n\}, \quad\textup{and} \\
1263: H_3 & = & \{(\pi, \pi v - v) : \pi \in S_n\},
1264: \end{eqnarray*}
1265: where $u = (1, 2, \dots, n)$, and $v = (n, n-1, \dots, 1)$.
1266: 
1267: As each subgroup has size $n!$ in a group of size $n!(2n)^n$,
1268: $$
1269: \alpha \le \frac{\log (n!(2n)^n)}{\log n!},
1270: $$
1271: assuming the triple product property holds. The largest character
1272: degree of $G_n$ is $|S_n| = n!$ (see Theorem~25.6 in \cite{H})
1273: and so $|G|^{1/\gamma} = n!$, which implies
1274: $$
1275: \gamma = \frac{\log (n!(2n)^n)}{\log n!}.
1276: $$
1277: By Stirling's formula,
1278: $$
1279: \frac{\log (n!(2n)^n)}{\log n!} = 2 + \frac{1+\log 2}{\log n} +
1280: O\left(\frac{1}{(\log n)^2}\right),
1281: $$
1282: so all that remains is to verify the triple product property.
1283: 
1284: Suppose $h_1 = (\pi',0) \in H_1$ and $h_2 = (\pi, \pi u -u ) \in
1285: H_2$.  Their product is $(\pi'\pi,\pi u - u)$, and if it equals
1286: $h_3 = (\sigma, \sigma v - v) \in H_3$, then $\pi u - u = \sigma
1287: v - v$. The $i$-th coordinate of $\pi u - u$ is $\pi(i) - i$, and
1288: that of $\sigma v - v$ is $(n+1 - \sigma(i)) - (n+1-i) =
1289: i-\sigma(i)$. Thus, $h_1h_2=h_3$ implies $\pi(i)+\sigma(i) = 2i$
1290: for all $i$. This is an equation in $A$, and hence holds only
1291: modulo $2n$. However, $\pi(i)$, $\sigma(i)$, and $i$ are all in
1292: $\{1,\dots,n\}$, so the equation holds in the integers as well.
1293: Because $\pi(1)$ and $\sigma(1)$ are both at least $1$, we
1294: conclude from $\pi(1)+\sigma(1)=2$ that $\pi(1)=\sigma(1)=1$.
1295: Then $\pi(2)$ and $\sigma(2)$ must be at least $2$, and
1296: $\pi(2)+\sigma(2)=4$, so $\pi(2)=\sigma(2)=2$, etc.  We conclude
1297: that $\pi$ and $\sigma$ are both trivial, as is $\pi'$ because
1298: $\pi'\pi=\sigma$.  Thus, $h_1=h_2=h_3=1$, as desired.
1299: \end{proof}
1300: 
1301: This construction is an improvement over
1302: Theorem~\ref{theorem:pseudoexponent2}, because it achieves
1303: essentially the same pseudo-exponent bound, while at the same time
1304: $\alpha \le \gamma$. A more complicated variant of this
1305: construction achieves a comparable pseudo-exponent and has
1306: $\alpha < \gamma$.
1307: 
1308: \SubSection{Direct products and the Sperner capacity}
1309: \label{section:direct}
1310: 
1311: It is natural to attempt to improve the pseudo-exponent of a
1312: finite group $G$ by forming some group derived from it, such as a
1313: power $G^k$.  We know that $\gamma(G^k) = \gamma(G)$, so that
1314: parameter becomes no smaller.  Lemma~\ref{lemma:upperbound}
1315: implies that $\alpha(G^k) \le \alpha(G)$, and in this section we
1316: show that it is possible to achieve $\alpha(G^k) < \alpha(G)$.
1317: 
1318: We will be led for the first time since
1319: Lemma~\ref{lemma:shortexact} to realize matrix multiplication
1320: through quotient sets that are {\em not} subgroups.
1321: Proposition~\ref{prop:dihedral} below proves that this
1322: complication is necessary to determine the pseudo-exponents of
1323: certain groups.
1324: 
1325: Let $D_m$ be the dihedral group generated by $x$ and $y$, with the
1326: relations $y^2 = x^m = 1$ and $yxy = x^{-1}$.
1327: 
1328: \begin{proposition}
1329: \label{prop:dihedral} For every $m$, $D_m$ realizes $\langle
1330: 2,2,2\lfloor m/3 \rfloor\rangle$, and hence $\alpha(D_m)<3$ for
1331: $m\ge 9$.  If $m$ is a prime greater than $3$, then no three
1332: subgroups prove $\alpha(D_m)<3$.
1333: \end{proposition}
1334: 
1335: \begin{proof}
1336: Let $S_1 = \langle y \rangle$ be the subgroup generated by $y$,
1337: $S_2 = \langle yx^2 \rangle$, and $S_3 = \{x^{3k}, yx^{3k+1} : 0
1338: \le k < (m-2)/3\}$.  Then one can check by simple case analysis
1339: that $D_m$ realizes $\langle 2,2,2\lfloor m/3 \rfloor\rangle$
1340: through $S_1,S_2,S_3$.  Note that $S_3$ is a subgroup iff $m$ is a
1341: multiple of $3$.
1342: 
1343: When $m$ is prime, all subgroups of $D_m$ have order $1$, $2$,
1344: $m$, or $2m$, and it is easy to rule out each case (except when
1345: $m=3$, in which case three subgroups of order~$2$ prove
1346: $\alpha(D_3)<3$).
1347: \end{proof}
1348: 
1349: Proposition~\ref{prop:dihedral} is not optimal: $D_5$ realizes
1350: $\langle 2, 2, 3 \rangle$ through $\{1, y\}, \{1, yx\}, \{1, x^2,
1351: yx^4\}$.  However, we have checked using GAP that it is optimal
1352: for $m=4$, and thus $\alpha(D_4)=3$.
1353: 
1354: We now use the combinatorial notion of {\em Sperner capacity} to
1355: show that $\alpha(D_4^k)<3$ for large $k$, despite the fact that
1356: $\alpha(D_4)=3$.
1357: 
1358: \begin{proposition}
1359: If $S \subseteq (\Z/m\Z)^k$ is a subset in which no two distinct
1360: vectors differ by an element of $\{0,1\}^k$, then $D_m^k$ realizes
1361: $\langle 2^k, 2^k, |S|\rangle$.
1362: \end{proposition}
1363: 
1364: \begin{proof}
1365: We identify $\Z/m\Z$ with the subgroup $\langle x \rangle
1366: \subseteq D_m$ (via $i \leftrightarrow x^i$), so that $S
1367: \subseteq \langle x \rangle^k \subseteq D_m^k$. The subgroups
1368: $\langle y \rangle$ and $\langle yx \rangle$ of $D_m$ have
1369: pointwise product $\langle y \rangle\langle yx \rangle =
1370: \{1,y,yx,x\}$.  Therefore the condition on differences of elements
1371: in $S$ implies that $\langle y \rangle^k$, $\langle yx
1372: \rangle^k$, and $S$ satisfy the triple product property, since
1373: $(\langle y \rangle^k \langle yx \rangle^k) \cap \langle x
1374: \rangle^k = \{1,x\}^k$, and $Q(S) \subseteq \langle x \rangle^k$
1375: avoids $\{1,x\}^k$.
1376: \end{proof}
1377: 
1378: The problem of making $S$ as large as possible has been studied
1379: before;  a generalization of this problem is known as the Sperner
1380: capacity of a directed graph \cite{Garg,Korn}.  It is known that
1381: $|S| \le (m-1)^k$ (see Theorem~1.2 in \cite{Alon}, which extends
1382: several earlier papers \cite{Blok, Cald}), and that
1383: $$
1384: |S| = (m-1)^{(1-o(1))k}
1385: $$
1386: can be achieved by the following construction:
1387: 
1388: Assume that $(m-1)$ divides $k$, and take $S$ to be the set of all
1389: vectors in $(\Z/m\Z)^k$ with exactly $k/(m-1)$ occurrences of each
1390: element of $\{0,1,\dots,m-2\}$. Now suppose we have $u, v \in S$
1391: with $u - v \in \{0,1\}^k$.  For each coordinate $i$ such that
1392: $u_i=0$, we have $v_i \in \{0,m-1\}$ because $u_i-v_i \in
1393: \{0,1\}$, and thus $v_i=0$.  Then whenever $u_i=1$, it follows
1394: that $v_i=1$, because all $k/(m-1)$ cases in which $v_i=0$ have
1395: $u_i=0$ as well. Repeating this argument yields $u=v$, as desired.
1396: 
1397: We conclude that direct products {\em can} help:
1398: 
1399: \begin{corollary}
1400: We have $\alpha(D_4^k) \le (3 + o(1)) \log_{12}{8}$, which
1401: approaches $3\log_{12}{8} = 2.51\dots$ as $k \rightarrow \infty$.
1402: \end{corollary}
1403: 
1404: This pseudo-exponent bound comes tantalizingly close to settling
1405: Question~\ref{fundamentalq}: if
1406: $$
1407: \liminf_{k \to \infty} \alpha(D_4^k) < 3 \log_{12} 8,
1408: $$
1409: then the answer to the question is ``yes,'' and our methods do in
1410: fact prove $\omega<3$ (at least).  The same holds in general for
1411: $D_{2n}$ (which has $n-1$ characters of degree $2$ and $4$ of
1412: degree $1$) ; the Sperner capacity construction proves that
1413: $\alpha(D_{2n}^k) \le (3+o(1)) \log_{8n-4} 4n$, and if
1414: $$
1415: \liminf_{k \to \infty} \alpha(D_{2n}^k) < 3\log_{8n-4} 4n,
1416: $$
1417: then the answer to Question~\ref{fundamentalq} is ``yes.''
1418: 
1419: Also, note that Lemma~\ref{lemma:upperbound} implies that for all
1420: $G$,
1421: $$
1422: \liminf_{k \to \infty} \alpha(G^k) = \inf_{k \ge 1} \alpha(G^k).
1423: $$
1424: Thus, even if the answer to Question~\ref{fundamentalq} is ``no,''
1425: there are combinatorial consequences.  For example, knowing that
1426: $\alpha(D_{2n}^k) \ge 3\log_{8n-4} 4n$ for all $n$ and $k$ would
1427: give a new proof of the Sperner capacity bound $|S| \le (m-1)^k$
1428: above, in the case of even $m$.
1429: 
1430: \Section{Concluding comments} \label{section:conclusions}
1431: 
1432: \SubSection{Open questions}
1433: 
1434: The most pressing question arising in this paper is
1435: Question~\ref{fundamentalq}, which represents a potential barrier
1436: to obtaining non-trivial bounds on $\omega$ using our
1437: techniques.  However, there are numerous other open questions that
1438: are relevant to Question~\ref{fundamentalq} and the ultimate goal
1439: of proving $\omega=2$.
1440: 
1441: \paragraph{Matrix groups.}
1442: As pointed out in Section~\ref{section:linear}, matrix groups
1443: seem to be one of the most promising families of examples, but we
1444: still know very little about them.  Can our bounds for
1445: $\alpha(SL_2(\F_q))$ be improved?  We see no reason why they
1446: should be optimal.  Recall that beating $9/4$ asymptotically would
1447: settle Question~\ref{fundamentalq}.  We know even less about
1448: $SL_n(\F_q)$ (only that $\alpha(SL_n(\F_q))<3$), so any
1449: non-trivial construction would be of interest.  The only other
1450: finite matrix groups that we have studied are those closely
1451: connected to $SL_n$ (such as $PSL_n$ or $GL_n$), but there are a
1452: number of other families.  What can one say about the
1453: pseudo-exponents of the groups in these families?
1454: 
1455: \paragraph{Quotient sets.}
1456: The examples in Subsection~\ref{section:direct} show that
1457: quotient sets sometimes outperform subgroups.  For which groups
1458: does this occur?  Are there general constructions of useful
1459: quotient sets other than via Sperner capacity?  Can they be used
1460: to improve our constructions for $S_n$ or the wreath product?
1461: What about matrix groups?
1462: 
1463: \paragraph{Lie groups.}
1464: Can one use Lie groups to prove anything about $\omega$
1465: directly?  Do results on the Lie pseudo-exponent imply anything
1466: about the pseudo-exponents of related finite groups?  Compact Lie
1467: groups seem more closely analogous to finite groups than
1468: non-compact Lie groups are, so studying them might be
1469: illuminating.  (All of the Lie groups in this paper are
1470: non-compact.)
1471: 
1472: \paragraph{Group extensions.}
1473: Extensions of groups with pseudo-exponent $3$ can have
1474: substantially smaller pseudo-exponents, as demonstrated by the
1475: solvable groups in Subsection~\ref{section:solvable}.  (Recall
1476: that solvable groups are formed from abelian groups by taking
1477: repeated extensions.)  Is there a general way to lower $\alpha$ or
1478: raise $\gamma$ by taking extensions?  As a first step, can one
1479: find a family of solvable groups with pseudo-exponents tending to
1480: $2$?
1481: 
1482: \paragraph{Powers of groups.}
1483: The simplest case of group extensions is taking powers of a
1484: group.  Given $G$, what can one say about the \textit{asymptotic
1485: pseudo-exponent\/} $\inf_{k \ge 1} \alpha(G^k)$ of $G$?  As noted
1486: in Subsection~\ref{section:direct}, $\gamma(G^k)=\gamma(G)$, so
1487: if there exists a group such that $\inf_{k \ge 1} \alpha(G^k) =
1488: 2$, then $\omega=2$ by Corollary~\ref{cor:race}.
1489: 
1490: \SubSection{Extensions} \label{section:extensions}
1491: 
1492: It is natural to attempt to extend our methods in various ways.
1493: For example, one might try to obtain bounds on border ranks of
1494: tensors, perhaps by using deformations of group algebras. It is
1495: also reasonable to ask whether our approach (given its reliance
1496: on representation theory) works in finite characteristic, as well
1497: as over $\C$. As Theorem~\ref{theorem:reduction} indicates, one
1498: can just as easily embed matrix multiplication into $F[G]$ rather
1499: than $\C[G]$, where $F$ has characteristic $p$. As long as $p$
1500: does not divide $|G|$, the representation theory of $G$, and all
1501: other aspects of our approach, work out identically, assuming $F$
1502: is algebraically closed. Sch\"onhage has shown that the exponent
1503: of matrix multiplication over arbitrary fields depends only on the
1504: characteristic (see Corollary~15.18 in \cite{BCS}), so we lose
1505: nothing by requiring that $F$ be algebraically closed.
1506: 
1507: We conclude by mentioning a particular variant of our approach
1508: that does not require any control of the character degrees, and
1509: thus may still be viable even if there is a negative answer to
1510: Question~\ref{fundamentalq}. We have found less structure to make
1511: use of, and it seems less attractive, but it uses similar ideas.
1512: Suppose we have distinct elements $x_{i, j}, y_{k, \ell} \in G$,
1513: for $1\le i \le n$, $1\le j,k \le m$, and $1\le \ell \le p$, such
1514: that
1515: \begin{equation}
1516: x_{i,j}y_{j, \ell} \sim x_{i', k}y_{k', \ell'} \; \Leftrightarrow
1517: \; i=i', k=k', \ell = \ell', \label{eq:conjugacy}
1518: \end{equation}
1519: where $\sim$ denotes conjugacy of elements. Then we embed matrix
1520: $A = (a_{i, j})$ as $\bar{A} = \sum_{i, j} a_{i, j}x_{i, j} \in
1521: \C[G]$, and matrix $B = (b_{k, \ell})$ as $\bar{B} = \sum_{k,
1522: \ell} b_{k, \ell}y_{k, \ell} \in \C[G]$.  We can pursue a similar
1523: strategy to compute $AB$. In this case, however, in the Fourier
1524: domain, we need only to compute the {\em trace} of each of the
1525: matrix products in the block-diagonal matrix multiplication. That
1526: requires only $\sum_i d_i^2 = |G|$ multiplications, and so we can
1527: conclude that the rank of $\langle n,m,p \rangle$ is at most
1528: $|G|$.
1529: 
1530: Let $G$ be a group with subsets $S_1, S_2$ and $S_3$ satisfying
1531: the triple product property. If we replace {\em conjugacy} with
1532: {\em equality} in \eqref{eq:conjugacy}, then it can be satisfied
1533: by taking $\{x_{i, j}\} = S_1 S_2^{-1}$ (where $i$ indexes $S_1$
1534: and $j$ indexes $S_2$) and $\{y_{k, \ell}\} = S_2 S_3^{-1}$ ($k$
1535: indexes $S_2$ and $\ell$ indexes $S_3$), so it is possible that
1536: the techniques we have developed in this paper could help with
1537: this variant as well, although in general we find it difficult to
1538: work with conjugacy constraints.
1539: 
1540: \section*{Acknowledgements}
1541: 
1542: We are grateful to Michael Aschbacher, Noam Elkies, Bobby
1543: Kleinberg, L\'aszl\'o Lov\'asz, Amin Shokrollahi, David Vogan, and
1544: Avi Wigderson for helpful discussions.
1545: 
1546: \begin{thebibliography}{10}\setlength{\itemsep}{-1ex}\small
1547: 
1548: \bibitem{Alon}
1549: N.~Alon.
1550: \newblock On the capacity of digraphs.
1551: \newblock {\em European J.\ Combinatorics}, 19:1--5, 1998.
1552: 
1553: \bibitem{Blok}
1554: A.~Blokhuis.
1555: \newblock On the {Sperner} capacity of the cyclic triangle.
1556: \newblock {\em J.\ Algebraic Combinatorics}, 2:123--124, 1993.
1557: 
1558: \bibitem{Brown}
1559: K.~S. Brown.
1560: \newblock {\em Cohomology of Groups}.
1561: \newblock Number~87 in Graduate Texts in Mathematics. Springer-Verlag, 1982.
1562: 
1563: \bibitem{BCS}
1564: P.~{B\"urgisser}, M.~Clausen, and M.~A. Shokrollahi.
1565: \newblock {\em Algebraic Complexity Theory}, volume 315 of {\em Grundlehren der
1566:   mathematischen Wissenschaften}.
1567: \newblock Springer-Verlag, 1997.
1568: 
1569: \bibitem{Cald}
1570: R.~Calderbank, P.~Frankl, R.~L. Graham, W.~Li, and L.~Shepp.
1571: \newblock The {Sperner} capacity of the cyclic triangle for linear and
1572:   nonlinear codes.
1573: \newblock {\em J.\ Algebraic Combinatorics}, 2:31--48, 1993.
1574: 
1575: \bibitem{CW}
1576: D.~Coppersmith and S.~Winograd.
1577: \newblock Matrix multiplication via arithmetic progressions.
1578: \newblock {\em J. Symbolic Computation}, 9:251--280, 1990.
1579: 
1580: \bibitem{GAP}
1581: The GAP~Group.
1582: \newblock {\em {GAP -- Groups, Algorithms, and Programming, Version 4.3}},
1583:   2002.
1584: \newblock \texttt{(http://www.gap-\break system.org)}.
1585: 
1586: \bibitem{Garg}
1587: L.~Gargano, J.~{K\"orner}, and U.~Vaccaro.
1588: \newblock {Sperner} theorems on directed graphs and qualitative independence.
1589: \newblock {\em J.\ Combinatorial Theory Series A}, 61:173--192, 1992.
1590: 
1591: \bibitem{Green}
1592: J.~A. Green.
1593: \newblock The characters of the finite general linear groups.
1594: \newblock {\em Transactions of the American Mathematical Society}, 80:402--447,
1595:   1955.
1596: 
1597: \bibitem{H}
1598: B.~Huppert.
1599: \newblock {\em Character Theory of Finite Groups}.
1600: \newblock Number~25 in de Gruyter Expositions in Mathematics. Walter de
1601:   Gruyter, Berlin, 1998.
1602: 
1603: \bibitem{JL}
1604: G.~James and M.~Liebeck.
1605: \newblock {\em Representations and Characters of Groups}.
1606: \newblock Cambridge University Press, Cambridge, second edition, 2001.
1607: 
1608: \bibitem{Korn}
1609: J.~{K\"orner} and G.~Simonyi.
1610: \newblock A {Sperner}-type theorem and qualitative independence.
1611: \newblock {\em J.\ Combinatorial Theory}, 59:90--103, 1992.
1612: 
1613: \bibitem{LR}
1614: J.~Lafferty and D.~Rockmore.
1615: \newblock Fast fourier analysis for {$SL_2$} over a finite field and related
1616:   numerical experiments.
1617: \newblock {\em Experimental Mathematics}, 1:115--139, 1992.
1618: 
1619: \bibitem{S}
1620: V.~Strassen.
1621: \newblock Gaussian elimination is not optimal.
1622: \newblock {\em Numerical Mathematics}, 13:354--356, 1969.
1623: 
1624: \bibitem{VK}
1625: A.~M. Vershik and S.~V. Kerov.
1626: \newblock Asymptotics of the largest and the typical dimensions of irreducible
1627:   representations of a symmetric group.
1628: \newblock {\em Functional Analysis and its Applications}, 19:21--31, 1985.
1629: 
1630: \end{thebibliography}
1631: 
1632: \end{document}
1633: