cs0302028/body.tex
1: \section{Introduction}
2: 
3: The notion of a random Boolean function occurs many times, both
4: implicitly and explicitly, in the literature of theoretical computer
5: science.  Not long after Shannon~\cite{Sh38} pointed out the relevance
6: of Boolean algebra to the design of switching circuits, Riordan and
7: Shannon~\cite{RiSh42} obtained a lower bound to the complexity (the
8: size of series-parallel relay circuits, or of formulas with the
9: connectives ``and'', ``or'' and ``not'') of ``almost all'' Boolean
10: functions, and this bound can naturally be applied to a ``random''
11: Boolean function when all $2^{2^n}$ Boolean functions of $n$ arguments
12: are assumed to occur with equal probability.  Lupanov~\cite{Lu61b}
13: later showed that Riordan and Shannon's lower bound is matched
14: asymptotically by an upper bound that applies to all Boolean functions,
15: so in this situation the average case is asymptotically equivalent to
16: the worst case.  This asymptotic equivalence of average and worst cases
17: also holds in many other situations involving circuits or formulas.
18: There are some complexity measures, however, such as the length of the
19: shortest disjunctive-normal-form formula, for which the average case
20: behaves quite differently from the worst case (see Glagolev~\cite{Gl67},
21: for example), and the complexity of a random Boolean function remains a
22: challenging open problem.  In these cases, probability distributions
23: other than the uniform distribution have also been considered; for
24: example, one may assume that each entry in the truth-table is
25: independently $1$ with probability $p$ and $0$ with probability $1-p$,
26: so that the uniform distribution is the special case $p=1/2$ (see
27: Andreev~\cite{An84}).
28: 
29: Another approach to the study of random Boolean functions is to put a
30: probability distribution on formulas, and let that induce a probability
31: distribution on the functions that they compute.  This may be done by
32: using a ``growth process'' (defined below) to grow random formulas.
33: 
34: Valiant~\cite{Va84} considered such a growth process, and showed that
35: the resulting probability distribution tends to the distribution
36: concentrated on a single function:  the threshold function that assumes
37: the value $1$ if and only at least $n/2$ of its $n$ arguments assume
38: the value $1$.  This result was used to obtain a non-constructive upper
39: bound on the minimum possible size of a  formula for computing this
40: threshold function.  This argument in fact gives the best upper bound
41: currently known for this and similar threshold functions.
42: 
43: The choice of the initial probability distribution on formulas
44: dictates the probability distribution on functions.   To facilitate the
45: design and use of growth process, as in the case above, deriving a
46: characterization based on the initial conditions is an important problem.
47: 
48: One such result in this framework is due to Savick\'y~\cite{Sa90}.
49: Savick\'y formulated broad conditions under which the distribution of
50: the random function computed by a formula of depth $i$ tends to the
51: uniform distribution on all Boolean functions of $n$ variables as
52: $i\to\infty$.
53: 
54: Savick\'y~\cite{Sa95} has also shown that in some cases the rate of
55: approach of the probability of computing a particular function $f$ to
56: the uniform probability $2^{-2^n}$ gives information about $f$:  it is
57: fastest for the linear functions, and slowest for the ``bent''
58: functions (which are furthest, in Hamming distance, from the linear
59: functions).  For some other models of random formulas, Lefmann and
60: Savick\'y~\cite{LeSa97} and Savick\'y~\cite{Sa98} have shown that the
61: logarithm of the probability of computing a particular function is
62: related to the complexity of that function (as measured by the size of
63: the smallest formula computing that function).  Finally, we should
64: mention that Razborov~\cite{Ra88} has used random formulas in yet
65: another model to show that some large graphs with Ramsey properties
66: have representations by formulas of exponentially smaller size.  This
67: result, which has been improved quantitatively by
68: Savick\'y~\cite{Sa95b}, shows that Ramsey properties are possessed by
69: graphs that are far from random.
70: 
71: Our goal in this paper is to determine under what circumstances results
72: like Valiant's and Savick\'y's hold.  We show that for many growth
73: processes, the probability distribution on the computed function tends
74: to the uniform distribution on some set of functions (which may range
75: in size from a single function, as in Valiant's result, to all
76: functions, as in Savick\'y's).
77: 
78: 
79: 
80: \section{Definitions}\label{sec:defs}
81: Let $\cF_n$ denote the family of $n$-adic Boolean functions, let
82: $\cM_n$ denote the family of $n$-adic monotone Boolean functions, and
83: let $\cL_n$ denote the family of $n$-adic linear functions.  The set $B_n$
84: denotes Boolean cube of size $n$.
85: 
86: Let $k$ be a positive integer and $\alpha$ be a $k$-adic Boolean
87: function, which we call the \myem{connective}.  Let $A_0 =
88: \{x_1,x_2,...,x_n, \bar{x}_1,...,\bar{x}_n,0,1\}$ be the set comprising
89: the projection functions, their negations, and the constant functions,
90: and let $A_i = \{ \alpha(v_1,v_2,...,v_k)\ |\ v_i \in A_{i-1}\}$ be the
91: set comprising the formulas composed from $A_{i-1}$. A \myem{growth
92: process} is denoted by a pair $(\mu,\alpha)$, where $\mu$ is a
93: distribution on $A_0$ and $\alpha$ is a connective; $\mu$ is called the
94: \myem{initial distribution}.  A growth process gives rise to a
95: probability distribution $\pi_i$ on $A_i$ for each $i\ge 0$ in the
96: following way.  We take $\pi_0 = \mu$.  For $i\ge 1$, we take
97: $\pi_i(f)$ to be the probability that $\alpha(g_1, \ldots, g_k) = f$,
98: where $g_1, \ldots, g_k$ are independent random functions distributed
99: according to $\pi_{i-1}$ on $A_{i-1}$.
100: 
101: We shall assume that $\mu$ is a uniform distribution on a subset of
102: $A_0$.  This subset will always contain the $n$ projections; it may or
103: may not contain their $n$ negations; and it may contain neither, one,
104: or both of the two constants.  All of our results could be extended to
105: more general distributions $\mu$, but these assumptions allow us to
106: present the most interesting results with a minimum of notation.  They
107: also cover the results of Valiant~\cite{Va84} and
108: Savick\'y~\cite{Sa90}.  (Valiant's proof actually uses a non-uniform
109: distribution, but the same bound can be obtained by a simple
110: modification using a uniform distribution on the projections.)
111: 
112: The \myem{support} of a probability distribution $\pi$, denoted
113: $\Supp(\pi)$, is the set $\{f\ |\ \pi(f) > 0\}$.  The \myem{support}
114: of a growth process is the set of all functions $f\in \cF_n$ for
115: which $\pi_i(f) > 0$ for some $i > 0$: $\cup_i \Supp(\pi_i)$.
116: 
117: We are particularly interested in cases in which $\pi_i$ tends to
118: a \myem{limiting distribution} $\pi$ as $i\to\infty$.  (There are
119: also cases in which $\pi_{2i}$ and $\pi_{2i+1}$ tend to distinct
120: \myem{alternating limiting distributions}.) When a limiting
121: distribution exists, we can have $\pi(f) > 0$ only for $f$ in the
122: support of the growth process.  As Valiant's result indicates,
123: however, there may be functions in the support for which $\pi_i(f)\to
124: 0$, so that $\pi(f) = 0$.  The \myem{asymptotic support} of a growth
125: process with a limiting distribution $\pi$ is the set of functions
126: $f\in \cF_n$ for which $\pi(f) > 0$.
127: 
128: Additionally, we investigate how quickly the distribution $\pi_i$
129: approaches the limiting distribution as $i$ approaches infinity.
130: Namely, for some $\epsilon > 0$, the size of $i$ such that
131: $\max_f|\pi(f) - \pi_i(f)| < \epsilon$.  Almost all growth processes
132: that we study share the important characteristic:  for any $\epsilon >
133: 0$,\ $\max_f|\pi(f) - \pi_{O(\log(n))}(f)| < \epsilon$.  Note, unless
134: otherwise stated, the base of the logarithm is assumed to be $2$.
135: 
136: Growth processes in which the limiting distribution is concentrated on
137: one function are used extensively in probabilistic amplification
138: methods and can be analyzed by studying the properties of the
139: corresponding ``characteristic polynomial''.  Let $\{X_1,X_2,...,X_n\}$
140: be a set of random independent binary variables that are $1$ with
141: probability $p$ and let each $X_i$ represent the input $x_i$.  The
142: \myem{characteristic polynomial} of $f$ is defined by $A_f(p) =
143: \Pr[f(X_1,X_2,...X_n) =1]$ and is given by
144:  \[A_f(p) = \sum_{i=0}^n \beta_i\Ch{n}{i}p^i(1-p)^{n-i}\]
145: where $\beta_i$ is the fraction of assignments of weight $i$ for which
146: $f$ is true.  The characteristic polynomial was used by von
147: Neumann~\cite{vN56} and by Moore and Shannon~\cite{MoSh56} to study
148: reliable computation with unreliable components, as well as by
149: Valiant~\cite{Va84} (see also Boppana~\cite{Bo85,Bo89,DuZw97}).
150: 
151: To analyze growth processes whose limiting distribution is uniform over
152: a set of functions, we use a Fourier transform technique.  The Fourier
153: transform $\Delta_i$ of a probability distribution $\pi_i$ is defined by
154: \begin{equation}\label{eqn:fft}
155: \Delta_i(f) = \sum_{g \in \cF_n} (-1)^\iprod{f}{g} \pi_i(g)
156: \end{equation}
157: where $\pi_i(g)$ is the probability of selecting $g$ from $A_i$.  For
158: convenience, the inner product $\iprod{f}{g} = \sum_i f_ig_i$ is
159: defined to be over the integers, rather than over $\Zf_2$.  Unless
160: otherwise noted, Boolean $n$-adic functions are represented as Boolean
161: vectors from $B_{2^n}$.  The inverse Fourier transform is defined by
162: \begin{equation}\label{eqn:inv}
163: \pi_i(g) = \frac{1}{2^{2^n}}\sum_{f \in \cF_n} (-1)^\iprod{f}{g} \Delta_i(f).
164: \end{equation}
165: The Fourier transform was used by Razborov~\cite{Ra88} to derive his
166: results on Ramsey graphs, as well as by Savick\'y~\cite{Sa90}.
167: 
168: The Fourier transform plays a role in many of our results, but it needs
169: to be adapted in various ways to suit different cases.  When dealing
170: with linear functions, for example, we will have to represent the
171: functions $f$ and $g$ in definition \ref{eqn:fft} not as Savick\'y
172: does, by their truth-tables, but rather by their coefficients as
173: multivariate polynomials over $GF(2)$.  In other cases, when
174: establishing a limiting distribution that is uniform over a proper
175: subset of $\cF_n$, we shall need to use what we call ``restriction
176: lemmas'', which assert relationships that hold among the values of the
177: Fourier transform.
178: 
179: 
180: \section{Growing Linear Functions}\label{sec:lin}
181: A function $f$ is linear if it is of the form $f(x_1, \ldots, x_n) =
182: c_0 \oplus c_1 x_1 \oplus \cdots \oplus c_n x_n$ for some constants
183: $c_0, c_1, \ldots, c_n \in GF(2)$.  We may assume without loss of
184: generality that $\alpha$ depends on all its arguments, so that
185: $\alpha(y_1, \ldots, y_k) = c \oplus y_1 \oplus\cdots\oplus y_k$, where
186: $k\ge 2$.  The result of the growth process depends on the support of
187: of the initial distribution $\mu$, the parity of $k$, and the constant
188: term $c$.
189: 
190: To prove this we derive a recurrence for the Fourier coefficients of
191: the respective probability distribution $\pi_i$, from which we derive
192: the limiting distribution.  Since compositions of linear functions
193: are themselves linear, we represent the linear functions by their
194: vector $(c_0, c_1, \ldots, c_n)$ of coefficients, and the following
195: summations range over $\cL_n$.  Finally, let $w_1$ denote the constant
196: function $1$ ($w_1 = 100\ldots0$), whereas $\onebf = 11\ldots1$.
197: 
198: \begin{proposition}\label{prop:lin_rec}
199: Let $\alpha$ be a linear connective as described above and let $w \in \cL_n$.
200: The Fourier coefficients of the probability distribution $\pi_i$ of the
201: corresponding growth process are described by the recurrence relation
202: \[\Delta_{i+1}(w) = (-1)^{c\iprod{w_1}{w}}\Delta_i(w)^k.\]
203: \end{proposition}
204: \begin{proof}
205: \begin{eqnarray*}
206: \Delta_{i+1}(w) & = & \sum_{f \in \cL_n} \pi_{i+1}(f) (-1)^\iprod{f}{w}
207:                   =   \sum_{f \in \cL_n}
208:                       \sum_\Sarg{\gbf \in \cL_n^k}{\alpha(\gbf) = f}
209:                       \prod_{j=1}^k \pi_i(\gbf_j)(-1)^\iprod{f}{w}\\
210:                 & = & \sum_{\gbf \in \cL_n^k} \prod_{j=1}^k \pi_i(\gbf_j)
211:                                                (-1)^\iprod{\alpha(\gbf)}{w}
212:                   =   \sum_{\gbf \in \cL_n^k} \prod_{j=1}^k \pi_i(\gbf_j)
213:                        (-1)^\iprod{cw_1\oplus\bigoplus_{j=1}^k \gbf_j}{w} \\
214:                 & = & \sum_{\gbf \in \cL_n^k} \prod_{j=1}^k \pi_i(\gbf_j)
215:             (-1)^{\iprod{cw_1}{w}\oplus\bigoplus_{j=1}^k\iprod{\gbf_j}{w}}
216:                   =   (-1)^\iprod{cw_1}{w} \sum_{\gbf \in \cL_n^k}
217:                            \prod_{j=1}^k \pi_i(\gbf_j) (-1)^\iprod{\gbf_j}{w} \\
218:                 & = & (-1)^\iprod{cw_1}{w} \Delta_i(w)^k
219: \end{eqnarray*}
220: \end{proof}
221: 
222: Using proposition~\ref{prop:lin_rec}, the following theorems classify the
223: growth processes on linear connectives.
224: 
225: \begin{theorem}\label{thm:lin}
226: Let $\alpha(y) = c \oplus y_1 \oplus\cdots\oplus y_k$,\ $k > 1$, be a
227: linear $k$-adic connective, as defined above, and assume that the support
228: of $\mu$ does not contain negations of the projections.
229: \begin{enumerate}
230: \item If $\{0,1\} \cap \Supp(\mu) \not= \{0,1\}$,\ $k$ is odd and $c = 1$,
231:       then the growth process has alternating limiting distributions,
232:       each of which is uniform over one half of the support of the
233:       growth process (which consists of all linear functions for which
234:       $\bigoplus_{j=1}^n c_j = 1$).
235: \item In all other cases, the limiting distribution is uniform over the
236:       support of the growth process (which depends on $k$, $c$, and the
237:       presence of constants in the support).
238: \end{enumerate}
239: \end{theorem}
240: \begin{proof}
241: Two facts are key to this theorem: first, that $|\Delta_i(w)| \leq 1$,
242: and second, that if $|\Delta_i(w)| < 1$, then $\lim_{i \rightarrow \infty}
243: \Delta_i(w) = 0$.  Only the nonzero (magnitude 1) coefficients
244: contribute to limiting distribution (equation~\ref{eqn:inv});
245: fortunately, these are determined solely by the support of the initial
246: distribution.  Depending on which constants are part of the support,
247: there are either one, two, or four magnitude 1 coefficients:
248: \begin{eqnarray*}
249: \{0,1\}\cap \Supp(\mu) = \{0,1\} & \Rightarrow & \Delta_0(0) = 1, \\
250: \{0,1\}\cap \Supp(\mu) = \{0\}   & \Rightarrow & \Delta_0(0)=\Delta_0(w_1)= 1,\\
251: \{0,1\}\cap \Supp(\mu) = \{1\}   & \Rightarrow & \Delta_0(0) = 1, \
252:                                                  \Delta_0(\onebf) = -1,\\
253: \{0,1\}\cap \Supp(\mu) =\emptyset& \Rightarrow & \Delta_0(0) = \Delta_0(w_1)= 1,
254:                         \ \Delta_0(\onebf)=\Delta_0(w_1\oplus\onebf) = -1.
255: \end{eqnarray*}
256: 
257: If $k$ is odd and $c = 1$, the recurrence from
258: Proposition~\ref{prop:lin_rec} implies that $\Delta_{i+1}(w_1) =
259: -\Delta_i(w_1)$ and $\Delta_{i+1}(\onebf) = -\Delta_i(\onebf)$.  Hence,
260: if $\Supp(\mu) \cap \{0,1\} \not= \{0,1\}$, the resulting distribution
261: is alternating.  In the case where one of the constants is missing from
262: the support, only two coefficients have magnitude 1, and thus, the
263: alternating distributions are each uniform over half of $\cL_n$.  In
264: the case where both constants are missing, the alternating
265: distributions are each uniform over one quarter of $\cL_n$.
266: 
267: If $c = 0$, $k$ is even, or $\{0,1\}\cap \Supp(\mu) = \{0,1\}$, the
268: limiting distribution exists because the sign of the magnitude 1
269: coefficients does not alternate.  We can read off the limiting
270: distribution from the Fourier coefficients.  If both constants are in
271: the support, then the limiting distribution is uniform over $\cL_n$.
272: If only one of the constants is present, then the distribution is
273: uniform over half of $\cL_n$, and if neither is present, then the
274: distribution will be uniform over a quarter of $\cL_n$.
275: \end{proof}
276: 
277: If the support of $\mu$ contains negations, then using the same proof
278: technique yields the following theorem.
279: 
280: \begin{theorem}\label{thm:lin_neg}
281: Let $\alpha(y) = c \oplus y_1 \oplus\cdots\oplus y_k$ be a linear
282: $k$-adic connective, as defined above, and assume that the support of
283: $\mu$ contains negations of the projections.
284: \begin{enumerate}
285: \item If $\{0,1\}\cap \Supp(\mu) = \emptyset$ and $k$ is odd then the
286:       limiting distribution is uniform over all linear functions of odd
287:       number of variables.
288: \item If $\{0,1\}\cap \Supp(\mu) = \emptyset$ and $k$ is even then the
289:       limiting distribution is uniform over all linear functions of
290:       even number of variables.
291: \item Otherwise, the limiting distribution is uniform over all of $\cL_n$.
292: \end{enumerate}
293: \end{theorem}
294: \begin{proof}
295: If $\{0,1\}\cap \Supp(\mu) \not= \emptyset$, then there is only one
296: coefficient of magnitude $1$, $\Delta_0(\zerobf) = 1$, implying the
297: last case.
298: 
299: Otherwise, there is one other magnitude $1$ coefficient,
300: $\Delta_0(\onebf \oplus w_1) = -1$.  If $k$ is odd, then
301: $\Delta_{i+1}(\onebf \oplus w_1) = \Delta_i(\onebf \oplus w_1)^k = -1$,
302: implying the first case of the theorem.  If $k$ is even, then
303: $\Delta_{i+1}(\onebf \oplus w_1) = \Delta_i(\onebf \oplus w_1)^k = 1$,
304: implying the second case.
305: \end{proof}
306: 
307: Note, that if negations are present, no alternating distribution can occur.
308: To bound the convergence of $\pi_i$ to $\pi$ we use the inverse Fourier
309: transform.
310: 
311: \begin{theorem}
312: Let $\alpha$ be a $k$-adic linear connective, $k > 1$, of a linear
313: process on $n$ variables that has a limiting distribution $\pi$.  If 
314: $i > \frac{2\log(n)}{\log(k)}$, then for any linear function $f$, 
315: $|\pi(f) - \pi_i(f)| < 2^{-n}$.
316: \end{theorem}
317: \begin{proof}
318: Let $D = \{w\ :\ |\Delta_0(w)| < 1\}$, then $\pi_i(f)$ may be written as:
319: \begin{eqnarray*}
320: \pi_i(f) & = & 2^{-n-1}\sum_{w \in B_{n+1}}(-1)^\iprod{w}{f}\Delta_i(w) \\
321:          & = & 2^{-n-1}\sum_{w \not\in D}(-1)^\iprod{w}{f}\Delta_i(w)
322:              + 2^{-n-1}\sum_{w \in D}(-1)^\iprod{w}{f}\Delta_i(w)\\
323:          & = & \pi(f)
324:              + 2^{-n-1}\sum_{w \in D}(-1)^\iprod{w}{f}\Delta_i(w).
325: \end{eqnarray*}
326: Thus, for any linear function $f$,
327: \[|\pi(f) - \pi_i(f)| = |2^{-n-1}\sum_{w \in D}(-1)^\iprod{w}{f}\Delta_i(f)|
328:    \leq   \max_{w \in D}|\Delta_0(w)|^{k^i} \leq (1-n^{-1})^{k^i}.\]
329: Solving inequality $(1-n^{-1})^{k^i} < 2^{-n}$, in terms of $i$, yields:
330: $i > \frac{2\log(n)}{\log(k)}$.
331: \end{proof}
332: 
333: 
334: \section{Growing Self-Dual Functions}\label{sec:self}
335: Savick\'y~\cite{Sa90} showed that if the connective is balanced (that
336: is, if it assumes the value $1$ for just one-half of the combinations
337: of argument values) and non-linear, and the support of $\mu$ is all of
338: $A_0$, then the limiting distribution will be uniform over all of
339: $\cF_n$.  If we remove the constants from the support of $\mu$ and
340: assume the connective $\alpha$ is self-dual (that is, satisfies
341: $\alpha(y_1, \ldots y_k) = \overline{\alpha(\bar{y}_1, \ldots,
342: \bar{y}_k)}$), then the support of the growth process is the set of all
343: self-dual functions.  In this case the limiting distribution of the
344: growth process is uniform over this support.
345: 
346: \begin{theorem}\label{thm:self_dual}
347: If the connective is non-linear and self-dual, and the support of $\mu$
348: comprises the projections and their negations, then the limiting
349: distribution will be uniform over the family of self-dual $n$-adic
350: functions.
351: \end{theorem}
352: \begin{proof}
353: Observe that there is a bijection between the set of all functions on $n$
354: variables and the set of self-dual functions on $n+1$ variables, for
355: example, the map
356: \[f(x_1,x_2,\ldots,x_n) \mapsto f(x_1,x_2,\ldots,x_n)x_{n+1}
357:   \vee \overline{f(\bar{x}_0,\bar{x}_1,\ldots,\bar{x}_n)}\bar{x}_{n+1}.\]
358: The result follows.
359: \end{proof}
360: 
361: 
362: 
363: \section{Growing Monotone Functions}\label{sec:mono}
364: We now focus on growth processes that use monotone connectives.  For
365: the rest of this section we assume that $\alpha$ is monotone and the
366: support of $\mu$ contains only monotone functions from $A_0$ (that is,
367: projections and possibly constants).  We first investigate unbalanced
368: connectives.
369: 
370: 
371: \subsection{Using Unbalanced Connectives}
372: Growth processes that use unbalanced monotone connectives concentrate
373: probability on a threshold function; the type of threshold function
374: depends on the connective and the support.  A threshold function
375: $T_k(x_1, \ldots, x_n)$ assumes the value $1$ if and only if at least
376: $k$ of its $n$ arguments assume the value $1$.  We consider constant
377: functions $T_{n+1} = 0$ and $T_0 = 1$ to be special cases of threshold
378: functions.  There are two cases to consider: first, when the
379: characteristic polynomial of $\alpha$, $A_\alpha(p)$, has no fixed-point
380: on the open interval $(0,1)$, and second, when $A_\alpha(p)$ has a
381: fixed-point on $(0,1)$.
382: 
383: \begin{proposition}\label{prop:no_fp}
384: If $\alpha$ is a monotone connective whose characteristic polynomial,
385: $A(p)$, has no fixed-point on the interval $(0,1)$, then the limiting
386: distribution will be concentrated on a threshold function.
387: \end{proposition}
388: \begin{proof}
389: Since $A(p)$ has no fixed-point on $(0,1)$, either $A_\alpha(p) < p$
390: throughout $(0,1)$, or $A_\alpha(p) > p$ throughout $(0,1)$.  If
391: $A_\alpha(p) > p$ throughout $(0,1)$, then by the standard
392: amplification argument, the limiting distribution is concentrated on
393: $T_1$ (disjunction of all variables) or, $T_0$  if $1$ is in the
394: support of $\mu$.  Similarly, if $A_\alpha(p) < p$ throughout $(0,1)$,
395: then the limiting distribution is concentrated on $T_n$ (conjunction of
396: all variables) or $T_{n+1}$ if $0$ is in the support of $\mu$.
397: \end{proof}
398: 
399: Furthermore, all connectives whose characteristic polynomials have no
400: fixed-point on $(0,1)$, are either of the form $\alpha(x) = x_i \vee
401: \alphap(x)$ (when $A_\alpha(p) > p$) or $\alpha(x) = x_i \wedge
402: \alphap(x)$ (when $A_\alpha(p) < p$).  If $\alpha(x) \not= x_i \vee
403: \alphap(x)$, then $A_\alpha(p) = O(p^2)$ which implies that there
404: exists a positive constant $\epsilon_0$ such that for all $0 < \epsilon
405: < \epsilon_0$, $A_\alpha(\epsilon) < \epsilon$.  Similarly, if
406: $\alpha(x) \not= x_i \wedge \alphap(x)$, then by duality, $1 -
407: A_\alpha(1-p) = O(p^2)$, which means that $A_\alpha(1-\epsilon) >
408: 1-\epsilon$ for all $0 < \epsilon < \epsilon_1$ for some $\epsilon_1 >
409: 0$.  Since $A_\alpha(p)$ is continuous, there must exist a
410: fixed-point in $(0,1)$, which is a contradiction.
411: 
412: In the second case, where $A_\alpha(p)$ has a fixed-point in $(0,1)$,
413: Moore and Shannon~\cite{MoSh56} have shown that this fixed-point is
414: unique.  Not surprisingly, the limiting distribution depends on the
415: fixed-point.  Thus, we first derive two facts about the fixed-point of
416: the characteristic polynomial, to deal with the second case.
417: 
418: \begin{lemma}\label{lem:bal}
419: The characteristic polynomial $A(p)$ has a fixed-point of $\Half$ if
420: and only if the connective $\alpha$ is balanced.
421: \end{lemma}
422: \begin{proof}
423: By definition $\sum_{i=0}^n \beta_i\Ch{n}{i}$ is the number of
424: assignments for which $\alpha$ is true.  If $A_\alpha(\Half) = \Half$, then
425: $A_\alpha(\Half) = \sum_{i=0}^n \beta_i\Ch{n}{i} (\Half)^i(\Half)^{n-i} =
426: \frac{1}{2^n} \sum_{i=0}^n \beta_i\Ch{n}{i} = \Half$.  Hence,
427: $\sum_{i=0}^n \beta_i \Ch{n}{i} = 2^{n-1}$ which means that $\alpha$ is
428: balanced.  Conversely, if $\alpha$ is balanced, then $A_\alpha(\Half) =
429: \Half$.
430: \end{proof}
431: 
432: \begin{lemma}\label{lem:equiv}
433: If $\alpha$ is a monotone, non-projection connective, then any fixed-point
434: of $A_\alpha(p)$ on $(0,1)$ is either irrational or $\Half$.
435: \end{lemma}
436: \begin{proof}
437: By contradiction; without loss of generality assume that the fixed-point
438: $p_0 = \frac{r}{s} < \frac{1}{2}$ and $\gcd{(r,s)} = 1$.  Hence,
439: \[A_\alpha\left(\frac{r}{s}\right) = \sum_{j=0}^k \beta_j\Ch{k}{j}
440:    \left(\frac{r}{s}\right)^j\left(\frac{s - r}{s}\right)^{k-j} =
441: \frac{r}{s}.\]
442: Multiplying both sides by $s^k$, noting that $\beta_k = \beta_{k-1} =
443: 1$, and evaluating the result modulo $(s-r)^2$ yields
444: \begin{eqnarray*}\label{eq:fp_half}
445: rs^{k-1} & \equiv & \sum_{j=0}^k \beta_j\Ch{k}{j} r^j(s - r)^{k-j}
446:           \equiv  r^k + kr^{k-1}(s-r) +
447:                 (s-r)^2\sum_{j=0}^{k-2}\beta_j\Ch{k}{j}r^j(s-r)^{k-j-2}\\
448:          & \equiv & r^k + kr^{k-1}(s-r) \bmod{(s-r)^2}.
449: \end{eqnarray*}
450: Evaluating the left side modulo $(s-r)^2$ yields
451: \begin{eqnarray*}
452: rs^{k-1} & \equiv & r(r + (s-r))^{k-1}
453:            \equiv  r\sum_{i=0}^{k-1} \Ch{k-1}{i}r^i(s-r)^{k-1-i} \\
454:          & \equiv & rr^{k-1} + r(k-1)r^{k-2}(s-r) +
455:                     r(s-r)^2\sum_{j=0}^{k-3}\Ch{k-1}{j}r^j(s-r)^{k-3-j} \\
456:          & \equiv & r^k + (k-1)r^{k-1}(s-r) \bmod{(s-r)^2}.
457: \end{eqnarray*}
458: Therefore,
459: \[r^{k-1}(s-r) \equiv 0 \bmod{(s-r)^2}.\]
460: Since $\gcd{(r,s)} = \gcd{(r,(s-r)^2)} = 1$, $r^{k-1} \not\equiv 0
461: \bmod{(s-r)^2}$; this is a contradiction.
462: \end{proof}
463: 
464: 
465: \begin{theorem}\label{thm:mono_unbal}
466: Let $\alpha$ be a monotone unbalanced connective whose characteristic
467: polynomial has a fixed-point $t \in (0,1)$, and let the support of
468: $\mu$ contain only the projections.  The limiting distribution of
469: the growth process is concentrated on the threshold function $T_{\lceil
470: tn\rceil}$.
471: \end{theorem}
472: \begin{proof}
473: Since $\alpha$ is unbalanced and has a fixed-point on $(0,1)$, by
474: Lemma~\ref{lem:bal}, the fixed-point is not $\Half$.  Hence, by
475: Lemma~\ref{lem:equiv}, the fixed-point is irrational.  Since the
476: fraction of variables set to true in any assignment is by definition
477: rational, the fraction will always be strictly greater or strictly less
478: than the fixed-point $t$.  Hence, by the standard amplification argument,
479: the limiting distribution will be concentrated on the threshold function
480: $T_{\lceil tn\rceil}$.
481: \end{proof}
482: 
483: Theorem \ref{thm:mono_unbal} can easily be modified to cover the cases
484: in which one or both constants are in the support of $\mu$.   Combining
485: proposition~\ref{prop:no_fp} and theorem~\ref{thm:mono_unbal} proves
486: the initial claim.
487: 
488: \begin{theorem}
489: If $\alpha$ is a monotone unbalanced connective and the support of
490: $\mu$ does not contain the negations of projections, then the limiting
491: distribution will be concentrated on a threshold function.
492: \end{theorem}
493: 
494: \subsubsection{Convergence Bounds}
495: Except in one case, all these growth processes converge very quickly to
496: their limiting distribution: in $O(\log(n))$ iterations.  In the
497: exceptional case the convergence requires $O(n^k)$ iterations where $k$
498: is the arity of the connective $\alpha$; we provide specific criteria
499: that determine whether a process will converge quickly or not.  There
500: are two main cases: either $A_\alpha(p)$ has a fixed-point, or not.  We
501: first derive bounds for the latter case, and then for the former.
502: Unless explicitly stated, we assume that constants are not in
503: $\Supp(\mu)$, however, the following analysis changes little if
504: constants are in $\Supp(\mu)$.
505: 
506: In the first case, either $A_\alpha(p) > p$ for $p \in (0,1)$, and
507: $A_\alpha(p) < p$, for $p \in (0,1)$.  Since, the two cases are
508: symmetric, the same bounds apply to both.  Hence, without loss of
509: generality assume that $A_\alpha(p) < p$ on the interval $(0,1)$.
510: 
511: \begin{lemma}\label{LEM:FAST-NFP}\label{lem:fast-nfp}
512: If $\alpha$ is a monotone connective such that $A_\alpha(p) < p$ on the
513: interval $(0,1)$ and, $A_\alpha(p)$ has degree $k > 2$ and $\beta_{k-1}
514: \leq \frac{k-2}{k}$, then for all positive $\epsilon < \epsilon_k =
515: \frac{1}{k2^{k+1}}$,
516: \[\frac{35}{24} < \Ap_\alpha(1-\epsilon)\]
517: \end{lemma}
518: \begin{proof}
519: See Appendix.
520: \end{proof}
521: 
522: \begin{lemma}\label{LEM:SLOW-NFP}\label{lem:slow-nfp}
523: If $\alpha$ is a monotone connective such that $A_\alpha(p) < p$ on the
524: interval $(0,1)$ and, $A_\alpha(p)$ has degree $k > 2$ and $\beta_{k-1}
525: = \frac{k-1}{k}$, then, for all positive $\epsilon < k^{-1}$,
526: \[1+\epsilon^k < \Ap_\alpha(1-\epsilon)
527:             \leq (1-\epsilon)^{k-2}(k(k-2)\epsilon+1)\]
528: \end{lemma}
529: \begin{proof}
530: See Appendix.
531: \end{proof}
532: 
533: \begin{lemma}\label{LEM:TRIV}\label{lem:triv}
534: If $\alpha$ is a monotone connective that is not of the form $\alpha(x)
535: = x_i \vee \alphap(x)$, then on the interval $(0,1)$, $A_\alpha(p) <
536: (\Ch{k}{2}+1)p^2$.
537: \end{lemma}
538: \begin{proof}
539: See Appendix.
540: \end{proof}
541: 
542: \begin{theorem}
543: Let $\alpha$ be a $k$-adic monotone connective such that $A_\alpha(p) <
544: p$ on the interval $(0,1)$, $k > 2$ and $\beta_{k-2} \leq \frac{k-2}{k}$.
545: There exists a constant $c_\alpha$, such that for all $n > 0$, if $i \geq
546: 3\log(n) + c_\alpha$, then for all $f$, $|\pi_i(f) - \pi(f)| < 2^{-n}$.
547: \end{theorem}
548: \begin{proof}
549: Let $\ft_i$ be a random variable with the distribution $\pi_i$.  Using
550: an argument similar to Valiant's~\cite{Va84}, we claim that if $i \geq
551: 3\log(n) + c_\alpha$, then for $|x| = n$, $P[\ft_i(x) = 0] = 0$, and
552: for all $x$ such that $|x| < n$, $P[\ft_i(x) = 1] < 2^{-2n}$.  The
553: former follows from the monotonicity of $\alpha$; regardless of the
554: number of iterations, a false negative will never occur.
555: 
556: In the latter case, assuming that all variables are independent, if
557: $|x| < n$, then $P[\ft_0(x) = 1] = |x|/n \leq 1 - n^{-1}$.  For $i > 0$,
558: $P[\ft_i(x) = 1] =  A_\alpha^i(p)$, where $A_\alpha^i$ denotes the $i$th
559: composition of $A_\alpha$ with itself.  Expanding $A_\alpha(p)$ around $1$,
560: \[A_\alpha(p) = A_\alpha(1) + \Ap_\alpha(1)(p-1) + O((p-1)^2),\]
561: yields:
562: \[A_\alpha(1 - \epsilon) = 1 - \epsilon\Ap_\alpha(1) + O(\epsilon^2).\]
563: From Lemma~\ref{lem:fast-nfp}, let $\gamma = 35/24$ and let $\epsilon_k =
564: \frac{1}{k2^{k+1}}$.  There exists an $\epsilon_0 < \epsilon_k$ such that
565: for all $\epsilon < \epsilon_0$
566: \[A_\alpha(1 - \epsilon) < 1 - \epsilon\gamma.\]
567: Since $P[\ft_0(x) = 1] \leq 1 - n^{-1}$, for $i \geq 2\log(n) +
568: 2\log(\epsilon_0) > (\log(n) + \log(\epsilon_0))/\log(35/24)$,
569: \[A_\alpha^i(1 - \epsilon) < 1 - \epsilon\gamma^i < 1 - \epsilon_0.\]
570: 
571: An additional constant number of iterations, say $d_\alpha$, yields
572: \[A_\alpha^{d_\alpha}(1 - \epsilon_0) < c.\]
573: By Lemma~\ref{lem:triv}, $A_\alpha(p) < k^2p^2$, thus we fix $c <
574: \frac{1}{2k^2}$ and let $j = \log(n) + 1$. Hence,
575: \[A_\alpha^j(c) < (k^2c)^{2^j} < 2^{-2^j} = 2^{-2n}.\]
576: 
577: Therefore, for $i \geq 3\log(n) + 2\log(\epsilon_0) + d_\alpha + 1$ and
578: all $x$ such that $|x| < n$,\ $P[\ft_i(x)=1]<2^{-2n}$, implying that
579: $|\pi_i(f) - \pi(f)| < 2^{-n}$.
580: \end{proof}
581: 
582: Unfortunately, if $\beta = \frac{k-1}{k}$, convergence takes time
583: polynomial in $n$.  If $|x| = n-1$ then $P[\ft_0(x) = 1] = 1 - n^{-1}$.
584: Furthermore, by Lemma~\ref{lem:slow-nfp}, for sufficiently large $n$,
585: $\Ap_\alpha(1-n^{-1}) < (1-n^{-1})^{k-2}(k(k-2)n^{-1} + 1)$.  Since
586: $\gamma < \Ap_\alpha(1-n^{-1})$, therefore
587: \[\log(\gamma) < (k-2)\log(1-n^{-1}) + \log(k(k-2)n^{-1} + 1),\]
588: implying that
589: \[\log(\gamma)^{-1} > \left((k-2)\log(1-n^{-1})+\log(k(k-2)n^{-1}+1)\right)^{-1}
590:                     > \frac{n}{k^2-3k+2} + O(1).\]
591: Thus, if $A_\alpha^i(1-n^{-1}) < 1-\epsilon_0$, then for sufficiently
592: large $n$, $A_\alpha^i(1-n^{-1}) < \epsilon_0$ implies that $i >
593: (k-1)(k-2)2n(\log(n) + \log(\epsilon_0))$.  In fact, this is the best
594: case.  If $\alpha(x) = \vee_{i=2}^k (x_1\wedge x_i)$, then $A_\alpha(p)
595: = p - p(1-p)^{k-1}$.  By Lemma~\ref{lem:slow-nfp}, $\gamma < 1 +
596: n^{2-k}(k-1-kn^{-1})$, implying that $\log(\gamma) < \log(1 +
597: n^{2-k}(k-1-kn^{-1}))$, and
598: \[\log(\gamma)^{-1} > \left(\log( 1 + n^{2-k}(k-1-kn^{-1}))\right)^{-1}
599:                     > \frac{n^{k-2}}{k-1}.\]
600: Thus, if $A_\alpha^i(1-n^{-1}) < 1-\epsilon_0$, then for sufficiently
601: large $n$, $i > \frac{n^{k-2}}{k-1}(\log(n) + \log(\epsilon_0))$.
602: Consequently, connectives whose characteristic polynomial has no
603: fixed-point can be classified as either quickly converging or slowly
604: converging, with the value of the second last coefficient,
605: $\beta_{k-1}$, determining rate of convergence!
606: 
607: When the characteristic polynomial $A_\alpha(p)$ does have a
608: fixed-point on the interval $(0,1)$, a similar analysis is used.
609: 
610: 
611: \begin{lemma}\label{LEM:MINSLOPE}\label{lem:minslope}
612: Let $A_\alpha(p)$ be the characteristic polynomial of any $k$-adic
613: monotone connective. If $A_\alpha(p)$ has a fixed-point $s \in (0,1)$,
614: then $\Ap_\alpha(s) \geq 1 + \frac{k-2}{2^{k-2}}$.
615: \end{lemma}
616: \begin{proof}
617: See Appendix.
618: \end{proof}
619: 
620: \begin{theorem}\label{thm:bound-fp}
621: Let $\alpha$ be a $k$-adic monotone connective such that $A_\alpha(s) =
622: s \in (0,1)$.  There exists a constant $c_\alpha$, such that for all $n
623: > 0$, if $i \geq k2^k\log(n) + c_\alpha$, then for all functions
624: $f$, $|\pi_i(f)-\pi(f)|< 2^{-n}$.
625: \end{theorem}
626: \begin{proof}
627: Let $\ft_i$ be a random variable with the distribution $\pi_i$.  Using
628: an argument similar to Valiant's~\cite{Va84}, we claim that if $i \geq
629: k2^k\log(n) + c_\alpha$ then for all $x$ such that $|x| < sn$,
630: $P[\ft_i(x) = 1] < 2^{-2n}$, and for all $x$ such that $|x| > sn$,
631: $P[\ft_i(x) = 0] < 2^{-2n}$.  We first argue the former.
632: 
633: Assuming that all variables are independent, if $|x| < sn$, then $P[\ft_0(x)
634: = 1] \leq s - n^{-1}\epsilon_\alpha(n)$, where $\epsilon_\alpha(n) =
635: \min_{j\in\Zf}|s - \frac{j}{n}| = |s - \frac{j_0}{n}|$.  Since $s$ is
636: an algebraic of degree $k$, by Liouville's Approximation
637: Theorem~\cite{Ap97}
638: \[\epsilon_\alpha(n) = \left|s - \frac{j_0}{n}\right| > \frac{e_\alpha}{n^k},\]
639: where the constant $e_\alpha$ depends only on the connective.
640: 
641: For $i > 0$, $P[\ft_i(x) = 1] = A_\alpha^i(p)$, where $A_\alpha^i$
642: denotes the $i$th composition of $A_\alpha$ with itself.  Expanding
643: $A_\alpha(p)$ around $s$,
644: \[A_\alpha(p) = A_\alpha(s) + \Ap_\alpha(s)(p-s) + O((p-s)^2),\]
645: yields:
646: \[A_\alpha(s - \epsilon) = s - \epsilon\Ap_\alpha(s) + O(\epsilon^2).\]
647: By Lemma~\ref{lem:minslope}, fix $\gamma = 1+2^{-k+1}$; there
648: exists an $\epsilon_0$ such that for all $\epsilon < \epsilon_0$,
649: $A_\alpha(s - \epsilon) < s - \epsilon\gamma$.  Since $P[\ft_0(x) = 1]
650: \leq s - n^{-1}\epsilon_\alpha(n)$, if
651: \[i \geq \log(n\ \epsilon_\alpha(n)^{-1}\epsilon_0)/\log(\gamma) \geq
652:          \log(n^{k+1}e_\alpha^{-1}\epsilon_0)/\log(\gamma),\]
653: then
654: \[A_\alpha^i(s - \epsilon) < s - \epsilon\gamma^i < s - \epsilon_0.\]
655: 
656: An additional constant number of iterations, say $d_\alpha$, yields
657: \[A_\alpha^{d_\alpha}(s - \epsilon_0) < c;\]
658: By Lemma~\ref{lem:triv}, $A_\alpha(p) < k^2p^2$, thus, we fix $c <
659: \frac{1}{2k^2}$ and let $j = \log(n) + 1$. Hence,
660: \[A_\alpha^j(c) < (k^2c)^{2^j} < 2^{-2^j} = 2^{-2n}.\]
661: Therefore, if
662: \[i \geq k2^k\log(n) + \frac{\log(e_\alpha^{-1}\epsilon_0)}{\log(\gamma)}
663:                             + d_\alpha + 1,\]
664: for all $x$ such that $|x| < sn$, $P[\ft_i(x) = 1] < 2^{-2n}$.
665: 
666: By the same argument, if $|x| > sn$, $P[\ft_i(x) = 0] < 2^{-2n}$. Since
667: $|x| > sn$, for $P[\ft_0(x) = 0] \leq 1 - s - n^{-1}\epsilon_\alpha(n)$,
668: $P[\ft_1(x) = 0] = \bar{A}_\alpha(p) = 1 - A_\alpha(1 - p)$, and
669: $P[\ft_j(x) = 0] = \bar{A}_\alpha^j(p)$, $j > 0$.  Just as in the
670: preceding case, the composition of $\bar{A}_\alpha$ with itself first
671: yields a first order divergence from $1 - s$, followed by a second order
672: convergence towards zero.  Therefore, $|\pi_i(f) - \pi(f)| < 2^{-n}$.
673: \end{proof}
674: 
675: To reduce the constant in front of the log term, one solution is to use
676: a non-uniform initial distribution, as was done by Valiant~\cite{Va84}.
677: 
678: 
679: \subsection{Using Balanced Connectives}
680: In this subsection, it will be convenient to start by assuming that the
681: support of $\mu$ contains both constants, as well as the projections,
682: and to deal later with the cases in which one or both constants are
683: missing from the support of $\mu$.  If the connective is balanced, then
684: by Lemma~\ref{lem:bal}, its characteristic polynomial has a fixed-point
685: of $\Half$.  If the number $n$ of variables is odd, then the fraction
686: of inputs that are true for any assignment is bounded away from
687: $\Half$, that is, for any $j \in \{1,2,..,n+1\}, |\Half -
688: \frac{j}{n+2}| \geq \frac{1}{2n+4}$.  Hence, by the standard
689: amplification argument, the limiting distribution will be concentrated
690: on the $n$-adic majority function $T_{\lceil n/2\rceil}$.  In fact, the
691: convergence to the majority function is logarithmic in $n$; by
692: Theorem~\ref{thm:bound-fp}, if $i \geq k2^k\log(n) + O(1)$, the
693: $|\pi_i(f) - \pi(f)| < 2^{-n}$.  When the number of variables is even,
694: however, something completely different happens.
695: 
696: The family of slice functions, denoted $\cS_{m,n}$ and defined by
697: Berkowitz~\cite{Ber82}, are monotone $n$-adic functions that assume the
698: value $1$ for all assignments of weight greater than $m$, assume the
699: value $0$ for all assignments of weight less than $m$, and may take on
700: either value for assignments of weight $m$.  Unlike other growth
701: processes where the distribution is either concentrated on a single
702: function or is uniform on the support of the growth process, the growth
703: processes we are about to deal with have a limiting distribution that
704: is uniform on  $\cS_{n/2,n}$.  This set includes a large number,
705: $2^{n\choose n/2}$, of functions; but according to a result of
706: Korshunov~\cite{Ko80}, includes only a tiny fraction, less than $\exp
707: -({n\choose n/2+1} 2^{-n/2})$, of the support of the growth process,
708: which is the set $\cM_n$ of all monotone functions.
709: 
710: Define the $n$-adic functions
711: \[\chih(x) = \left\{\begin{array}{lr}
712:                     1, & |x| = \frac{n}{2} \\
713:                     0, & \mathit{otherwise}
714:                     \end{array}\right.
715: \mathrm{\ \ \ \ and\ \ \ \ }
716: \upsh(x) = \left\{\begin{array}{lr}
717:                     1, & |x| > \frac{n}{2} \\
718:                     0, & \mathit{otherwise}
719:                     \end{array}\right..\]
720: 
721: \begin{claim}\label{thm:Fourier_slice}
722: The Fourier coefficients of the probability distribution $\pi$ that is
723: uniform on the slice functions in $\cS_{n/2,n}$ are given by
724: \[\Delta(f) = \left\{\begin{array}{lr}
725:                      0,    & \iprod{f}{\chih} \not=0 \\
726:                      (-1)^\iprod{f}{\upsh}, & \iprod{f}{\chih} = 0
727:                      \end{array}\right.\]
728: \end{claim}
729: \begin{proof}
730: Let $c = {|\Shalf|}^{-1} = 2^{-{n\choose n/2}}$.  If $\iprod{f}{\chih} = 0$,
731: then
732: \[\Delta(f)   =   \sum_{g \in \cF_n} (-1)^\iprod{f}{g} \pi(g)
733:               =   c\sum_{g \in \Shalf} (-1)^\iprod{f}{g}
734:               =   c\sum_{g \in \Shalf} (-1)^\iprod{f}{\upsh}
735:               =   (-1)^\iprod{f}{\upsh}.\]
736: Otherwise let $w$ be a singleton such that $w \leq f \wedge \chih$ and
737: let $W = \{g \in \Shalf\ |\ g \geq w\}$.  Then
738: \[\Delta(f) = c\sum_{g \in \Shalf} (-1)^\iprod{f}{g}
739:             = c\sum_{g \in W} (-1)^\iprod{f}{g}+(-1)^\iprod{f}{g \oplus w}
740:             = 0.\]
741: \end{proof}
742: 
743: We shall need to combine amplification with Fourier methods to obtain
744: our result in this case.  The following ``Restriction Lemma'' is the
745: key to doing this.
746: 
747: \begin{claim}
748: Let $\alpha$ be a balanced monotone connective.  Then if $f(x) = 1$ for
749: some $x$ with $|x| < n/2$, or if $f(x) = 0$ for some $x$ with $|x| > n/2$,
750: then $\lim_{i\rightarrow\infty} \pi_i(f) = 0$.
751: \end{claim}
752: \begin{proof}
753: This follows from Theorem~\ref{thm:bound-fp}.
754: \end{proof}
755: 
756: \begin{lemma}[The Restriction Lemma]\label{lem:restrict}\ \\
757: If $\alpha$ is a balanced monotone connective, then for all $w \in \cF_n$,
758: \[\lim_{i\rightarrow\infty} \Delta_i(w) =
759:    (-1)^\iprod{\upsh}{w} \lim_{i\rightarrow\infty} \Delta_i(w \wedge \chih).\]
760: \end{lemma}
761: \begin{proof}
762: \newcommand{\wcxh}{{w\cap\chih}}
763: We begin with the definition
764: \[\Delta_i(w) = \sum_{v \in B_n} (-1)^\iprod{v}{w} \pi_i(v),\]
765: then rewrite the equation as
766: \[\Delta_i(w) = \sum_{v \in B_n} (-1)^\iprod{v}{w}\pi_i(v)
767:               = \sum_{t \leq \chih}\sum_{u \leq \notchih}
768:                    (-1)^\iprod{t\vee u}{w} \pi_i(t \vee u),\]
769: and consider the restriction of $w$ to the slice $\frac{n}{2}$, that is,
770: $w \wedge \chih$.
771: Since $\lim_{i\rightarrow\infty} \pi_i(t \vee u) = 0$ if $u \not= \upsh$,
772: $\lim_{i\rightarrow\infty} \Delta_i(w\wedge \chih)$ can be rewritten as
773: \begin{eqnarray*}
774: \lim_{i\rightarrow\infty} \Delta_i(w\wedge \chih)
775:  &=& \lim_{i\rightarrow\infty} \sum_{v \in B_n}
776:              (-1)^\iprod{v}{w\wedge \chih}\pi_i(v) \\
777:  &=& \lim_{i\rightarrow\infty} \sum_{t\leq\chih}\sum_{u \leq \notchih}
778:                    (-1)^\iprod{t\vee u}{w\wedge \chih} \pi_i(t \vee u) \\
779:  &=& \sum_{t\leq\chih}\sum_{u \leq \notchih}(-1)^\iprod{t\vee u}{w\wedge \chih}
780:                                    \lim_{i\rightarrow\infty}\pi_i(t \vee u) \\
781:  &=& \sum_{t \leq \chih} (-1)^\iprod{t \vee \upsh}{w\wedge \chih}
782:                                  \lim_{i\rightarrow\infty}\pi_i(t \vee \upsh) \\
783:  &=& \sum_{t \leq \chih} (-1)^\iprod{t}{w\wedge \chih}
784:                                  (-1)^\iprod{\upsh}{w\wedge \chih}
785:                                  \lim_{i\rightarrow\infty}\pi_i(t \vee \upsh)\\
786:  &=& \sum_{t \leq \chih} (-1)^\iprod{t}{w\wedge \chih}
787:                                  \lim_{i\rightarrow\infty}\pi_i(t \vee \upsh) \\
788:  &=& \sum_{t \leq \chih} (-1)^\iprod{t}{w}
789:                                    \lim_{i\rightarrow\infty}\pi_i(t \vee \upsh).
790: \end{eqnarray*}
791: This, in conjunction with
792: \begin{eqnarray*}
793: \lim_{i\rightarrow\infty} \Delta_i(w)
794:  &=& \lim_{i\rightarrow\infty} \sum_{t\leq\chih}\sum_{u \leq \notchih}
795:                    (-1)^\iprod{t\vee u}{w} \pi_i(t \vee u)
796:   =  \sum_{t\leq\chih}\sum_{u \leq \notchih} (-1)^\iprod{t\vee u}{w}
797:                                  \lim_{i\rightarrow\infty} \pi_i(t \vee u) \\
798:  &=& \sum_{t \leq \chih} (-1)^\iprod{t \vee \upsh}{w}
799:                                  \lim_{i\rightarrow\infty} \pi_i(t \vee \upsh)
800:   =  \sum_{t \leq \chih} (-1)^\iprod{t}{w}(-1)^\iprod{\upsh}{w}
801:                                  \lim_{i\rightarrow\infty}\pi_i(t \vee \upsh) \\
802:  &=& (-1)^\iprod{\upsh}{w} \sum_{t \leq \chih} (-1)^\iprod{t}{w}
803:                                  \lim_{i\rightarrow\infty} \pi_i(t \vee \upsh)
804:   =  (-1)^\iprod{\upsh}{w} \lim_{i\rightarrow\infty} \Delta_i(w\wedge \chih)
805: \end{eqnarray*}
806: yields the identity.
807: \end{proof}
808: 
809: Hence, all we need to show is that $\lim_{i\rightarrow\infty}\Delta_i(w)
810: = 0$ for $w$ such that $0 < w \leq \chih$. To do this we use
811: Savick\'y's~\cite{Sa90} argument, which uses induction on the weight of $w$
812: together with the recurrence
813: \[\Delta_{i+1}(w) = \sum_{j=1}^k a_j(w)\Delta_i(w)^j + y_i(w)\]
814: where
815: \begin{eqnarray*}
816: a_j(w) & = & \sum_\Sarg{t \in B_k}{|t| = j} S_\alpha(t)^{|w|}, \\
817: y_i(w) & = & \sum_\Sarg{\vbf \in \cF_n^k}{0 < \vbf_j < w}
818:                 \prod_\Sarg{a \in B_n}{w(a) = 1} S_\alpha(\vbf(a))
819:                 \prod_{j=1}^k \Delta_i(\vbf_j), \\ % \mathrm{and} \\
820: S_\alpha(t)
821:        & = & \frac{1}{2^k}\sum_{r \in B_k} (-1)^\iprod{r}{t}(-1)^{\alpha(r)}.
822: \end{eqnarray*}
823: Since the connective is not linear, this recurrence is much more
824: complicated than the one in proposition~\ref{prop:lin_rec}.  The
825: result is the following proposition.
826: 
827: \begin{proposition}\label{thm:result}
828: Let $\alpha$ be a monotone balanced non-projection connective, $n$ be even
829: and the support of $\mu$ comprise the projection functions and constants.
830: If $0 < w \leq \chih$, then $\lim_{i \rightarrow \infty} \Delta_i(w) = 0$.
831: \end{proposition}
832: \begin{proof}
833: Let $w \leq \chih$ and recall equation~\ref{eqn:fft}:
834: \[\Delta_0(w) = \sum_{f \in F_n} \pi_0(f)(-1)^\iprod{f}{w}
835:               = \frac{1}{n +2}\sum_{f \in A_0}(-1)^\iprod{f}{w}.\]
836: To prove this proposition we need only show that $\Delta_0(w) = 0$ if
837: $|w| = 1$, and that $|\Delta_0(w)| < 1$ if $|w| = 2$.  In the first
838: case, since $w \leq \chih$, $w$ is true on a single assignment of
839: weight $n/2$.  Hence, $\iprod{w}{x_i} = 1$ for exactly half the
840: projections, where $x_i$ is the $i$th projection function.   Hence, the
841: projections cancel each other out.  Similarly, the two constants
842: annihilate one another.  Hence, $\Delta_0(w) = 0$ if $|w| = 1$.
843: 
844: The latter case, $|w| = 2$, is only slightly harder.  Since the
845: constant $0$ is part of the support, there will at least one positive
846: contribution, $-1^\iprod{\zerobf}{w}\pi_0(\zerobf) = \frac{1}{n+2}$.
847: Hence, in order for $|\Delta_0(w)| = 1$ all other contributions must
848: also be positive, specifically, $\iprod{w}{x_i} = 0$ for all $x_i$; by
849: the pigeonhole principle this is not possible.  Hence, $|\Delta_0(w)| <
850: 1$ if $|w| = 2$.  Substituting these base cases into Theorem 5.3
851: of~\cite{Sa90} yields the result.  
852: \end{proof}
853: 
854: This proposition, together with Claim~\ref{thm:Fourier_slice}, yields the
855: one of our main results.
856: 
857: \begin{theorem}
858: Let $\alpha$ be a monotone balanced non-projection connective, $n$ be even
859: and the support of $\mu$ comprise the projection functions and constants.
860: Then the limiting distribution is uniform on the functions in $\cS_{n/2,n}$.
861: \end{theorem}
862: 
863: 
864: \subsubsection{Convergence Bounds}
865: To bound the convergence within the slice we use a theorem of
866: Savick\'y~\cite{Sa95}; the conditions of the theorem are verified in
867: Theorem~\ref{thm:result}.
868: 
869: \begin{theorem}[(Savick\'y, 4.8 in \cite{Sa95})] \label{thm:t48}
870: If $\alpha$ is balanced and nonlinear, $\Delta_0(w) = 0$ for every $w$
871: such that $|w| = 1$, $\Delta_0(w) < 1$ for every $w$ such that $|w| =
872: 2$, and there exists a $w$ such that $|w| = 2$ and $\Delta_0(w) > 0$,
873: then
874: \[\max_{f \in \cF_n}{|\pi_i(f) - \pi(f)|} = e^{-\Omega(i)}\]
875: \end{theorem}
876: 
877: A more explicit bound, in terms of the number of arguments and the
878: arity of the connective, is possible.  We use a more explicit version
879: of Lemma~\ref{lem:restrict} and bound the convergence of the growth
880: processes characterized by Theorem 5.3 in~\cite{Sa90}.   A corollary of
881: Lemma~\ref{lem:bound_rl} is that the same bound also applies to the
882: growth processes on monotone formulas whose limiting distribution is
883: uniform over the slice functions.
884: 
885: \begin{lemma}\label{LEM:BOUND_RL}\label{lem:bound_rl}
886: If $\alpha$ is a balanced monotone connective, then for all $w \in \cF_n$,
887: \[\Delta_i(w)=O(\epsilon^{2^i})+(-1)^\iprod{\upsh}{w}\Delta_i(w \wedge \chih).\]
888: \end{lemma}
889: \begin{proof}
890: See appendix.
891: \end{proof}
892: 
893: \begin{theorem}[(Savick\'y, 5.3 in \cite{Sa90})]\label{thm:t53}
894: Let $\zerobf < w \in \cF_n$, let $\alpha$ be a $k$-adic nonlinear
895: balanced connective, and assume that the initial distribution is
896: uniform over the projections, negations, and constants.  The
897: $\lim_{i\rightarrow\infty} \Delta_i(w) = 0$.
898: \end{theorem}
899: 
900: \begin{lemma}\label{LEM:W3P_EQ}\label{lem:w3p_eq}
901: Assume that the conditions of Theorem~\ref{thm:t53} are satisfied and
902: let $a = \sum_{t\in B_k}|S_\alpha(t)|^3 < 1 - 2^{-k}$.  If $|w| = d
903: \geq 2$ and
904: \begin{equation}\label{equ:id}
905: i_d = n2^k\log(a^{-1}) + \sum_{j=3}^d \frac{(k+1)^j j}{\log(a^{-1})},
906: \end{equation}
907: then $|\Delta_i(w)| \leq a^{i - i_d} b_d(i)$, where $b_d(i) =
908: (i - i_2 + 2)^{(k+1)^{d-3}}$, and $b_2(i) = 1$.
909: \end{lemma}
910: \begin{proof}
911: See appendix.
912: \end{proof}
913: 
914: \begin{theorem}\label{thm:savbounds}
915: Assume that the conditions of Theorem~\ref{thm:t53} are satisfied, let 
916: $a$ be as in Lemma~\ref{lem:w3p_eq}, and let
917: \[I = n2^k\log(a^{-1}) + \frac{2^{2n}(k+1)^{2^n}}{\log(a^{-1})}.\]
918: For any positive $c < 1$, if $w \not= \zerobf$ and 
919: \begin{equation}\label{eqn:final}
920: i \geq \frac{\log(c)}{\log(a)} +
921:           \frac{\log(i - I + 2)}{\log(a^{-1})}(k+1)^{2^n} + I,
922: \end{equation}
923: then $|\Delta_i(w)| \leq c$.
924: \end{theorem}
925: \begin{proof}
926: By Lemma~\ref{lem:w3p_eq} the coefficient of weight $2^n$ has the greatest
927: converging bound:
928: \[|\Delta_{i+1}(w)| \leq a^{i-i_{2^n}}(i - n2^k\log(a^{-1}) + 2)^{(k+1)^{2^n}}\]
929: where
930: \begin{eqnarray*}
931: i_{2^n} & = & n2^k\log(a^{-1}) +\sum_{j=3}^{2^n}\frac{(k+1)^j j}{\log(a^{-1})}\\
932:         & \leq & n2^k\log(a^{-1}) + \frac{2^{2n}(k+1)^{2^n}}{\log(a^{-1})} = I
933: \end{eqnarray*}
934: Solving for $i$ in the inequality
935: \[a^{i-i_{2^n}}(i - n2^k\log(a^{-1}) + 2)^{(k+1)^{2^n}} < c\]
936: completes the proof.
937: \end{proof}
938: 
939: Thus, by equation~\ref{eqn:inv}
940: \begin{eqnarray*}
941: \pi_i(g) 
942:   &=& \frac{1}{2^{2^n}}\sum_{f \in \cF_n} (-1)^\iprod{f}{g} \Delta_i(f)\\
943:   &=& \frac{1}{2^{2^n}}+\frac{1}{2^{2^n}}\sum_{f\in\cF_n\backslash\zerobf}
944:                   (-1)^\iprod{f}{g} \Delta_i(f)\\
945:   &\leq& \frac{1}{2^{2^n}} +
946:          \frac{1}{2^{2^n}}\sum_{f \in \cF_n\backslash \zerobf} |\Delta_i(f)|\\
947:   &\leq& \frac{1}{2^{2^n}}+\max_{f\in\cF_n\backslash\zerobf}|\Delta_i(f)|,
948: \end{eqnarray*}
949: implying that for all $g \in \cF_n$, $|\pi(g) - \pi_i(g)| \leq c$ if $i$ 
950: satisfies equation~\ref{eqn:final}.
951: 
952: 
953: \subsubsection{Varying the Initial Distribution}
954: Theorem~\ref{thm:result} can easily be modified to cover the cases in
955: which one of the constants is missing from the support of $\mu$:  in
956: these cases there is concentration on a single function when $n$ is
957: even and uniform distribution on a set of slice functions when $n$ is
958: odd.  When the support of $\mu$ consists only of the projection
959: functions, however, the situation can be more complicated.  If $\alpha$
960: is not self-dual or $n$ is odd, the result is the same as when both
961: constants are present.  If $\alpha$ is self-dual and $n$ is even,
962: however, the result is given by the following theorem.
963: 
964: \begin{theorem}
965: Let $\alpha$ be a monotone self-dual non-projection connective, $n$ be
966: even and the support of $\mu$ comprise the projection functions.  Then
967: the limiting distribution is uniform on the self-dual functions in
968: $\cS_{n/2,n}$.
969: \end{theorem}
970: \begin{proof}
971: Same as theorem~\ref{thm:self_dual}.
972: \end{proof}
973: 
974: We note that there are $2^{\Half{n\choose n/2}}$ self-dual functions in
975: $\cS_{n/2,n}$.  According to a result of Sapozhenko~\cite{Sa89}, this
976: is only a tiny fraction, less than $\exp -({n\choose n/2+1}
977: 2^{-n/2-1})$, of the support of the growth process, which is the set of
978: self-dual monotone functions.
979: 
980: 
981: 
982: \section{Growing Other Functions}\label{sec:other}
983: We can use the same method to analyze other growth processes.  For
984: example, the uniform distribution on the set of bi-preserving functions
985: (that is, those functions satisfying $f(0, \ldots, 0) = 0$ and $f(1,
986: \ldots, 1) = 1$) can be generated by a growth process that uses the
987: bi-preserving selection connective $\alpha(x,y,z) = xy \vee \bar{x}z$
988: and an initial distribution that is uniform on the projection
989: functions.  The same technique as in the monotone case is sufficient to
990: prove this; the corresponding restriction lemma yields the identity
991: \[\lim_{i\rightarrow\infty}\Delta_i(w) = (-1)^\iprod{w}{\eta_n}
992:                                         \Delta(w \wedge \kappa_n)\]
993: where $\eta_n = \wedge_{j=1}^n x_j$ and $\kappa_n = \vee_{j=1}^n x_j
994: \bigwedge \overline{\wedge_{j=1}^n x_j}$.  Similar analysis for the
995: $0$-preserving and $1$-preserving functions follows easily.
996: 
997: 
998: \section{Conclusion}\label{sec:conc}
999: In this paper we have developed techniques for analyzing growth processes
1000: when the limiting distribution is uniform over a set of Boolean functions.
1001: In particular, we can deal with situations in which this set comprises
1002: neither a single function nor all Boolean functions.
1003: 
1004: We believe that straightforward extensions of the techniques used here
1005: will yield a classification of all situations leading to uniform
1006: distributions over sets of functions.  The step that remains to be
1007: taken is the classification of all sets of functions that can be
1008: computed by formulas that are complete $k$-ary trees built from a
1009: single connective.  There is some work by Kudryavtsev~\cite{Ku60,Ku60b}
1010: on this problem, but it stops short of a complete classification.
1011: 
1012: A more adventurous direction for further work is to deal with
1013: situations leading to non-uniform distributions.  Empirical
1014: computations show that these situations can be quite complicated, and
1015: we are not yet able to formulate a conjecture that covers all our
1016: data.
1017: