cs0702052/main.tex
1: \documentclass[conference,twocolumn]{IEEEtran}
2: \usepackage{amsmath, amssymb, amsfonts, latexsym}
3: \usepackage{graphicx}
4: %\usepackage{float}
5: 
6: % Theorems
7: %-----------------------------------------------------------------
8: \newtheorem{corollary}{Corollary}
9: \newtheorem{proposition}{Proposition}
10: \newtheorem{theorem}{Theorem}
11: \newtheorem{lemma}{Lemma}
12: %\theoremstyle{definition}
13: \newtheorem{definition}{Definition}
14: %\theoremstyle{remark}
15: \newtheorem{remark}{Remark}
16: 
17: 
18: \def\RR{\mathbb{R}}
19: \def\ZZ{\mathbb{Z}}
20: \def\NN{\mathbb{N}}
21: \def\FF{\mathbb{F}}
22: \newcommand{\abs}[1]{\left\vert#1\right\vert}
23: \newcommand{\tran}[1]{#1^T}
24: \DeclareMathOperator{\Span}{span}
25: \DeclareMathOperator{\rank}{rank}
26: \DeclareMathOperator{\head}{head}
27: \DeclareMathOperator{\tail}{tail}
28: 
29: \newcommand{\edmonds}{Z}
30: \newcommand{\graph}{\mathcal{G}}
31: \newcommand{\edge}{e}
32: \newcommand{\node}{v}
33: \newcommand{\edges}{\mathcal{\MakeUppercase{\edge}}}
34: \newcommand{\nodes}{\mathcal{\MakeUppercase{\node}}}
35: \newcommand{\numedges}{\MakeUppercase{\edge}}
36: \newcommand{\numnodes}{\MakeUppercase{\node}}
37: \newcommand{\din}[1]{d_{\text{in}}\left(#1\right)}
38: \newcommand{\dout}[1]{d_{\text{out}}\left(#1\right)}
39: 
40: \newcommand{\event}{\mathsf{E}}
41: 
42: %-----------------------------------------------------------------
43: \title{On Random Network Coding for Multicast}
44: \author{
45: \authorblockN{Adrian Tauste Campo}
46: \authorblockA{Universitat Pompeu Fabra\\Barcelona, Spain}
47: \and
48: \authorblockN{Alex Grant}
49: \authorblockA{Institute for Telecommunications Research\\
50:   University of South Australia}
51: }
52: 
53: \begin{document}
54: \maketitle
55: 
56: \begin{abstract}
57:   Random linear network coding is a particularly decentralized
58:   approach to the multicast problem. Use of random network codes
59:   introduces a non-zero probability however that some sinks will not
60:   be able to successfully decode the required sources. One of the main
61:   theoretical motivations for random network codes stems from the
62:   lower bound on the probability of successful decoding reported by Ho
63:   et. al. (2003). This result demonstrates that all sinks in a
64:   linearly solvable network can successfully decode all sources
65:   provided that the random code field size is large enough. This paper
66:   develops a new bound on the probability of successful decoding.
67: \end{abstract}
68: 
69: \section{Introduction}
70: It has been recently proved that network layer coding can increase
71: throughput, particularly for multicast scenarios \cite{AhlCai00}. It
72: is also known that linear network codes \cite{LiYeu03} can achieve
73: max-flow upper bounds on the throughput in a single source multicast
74: network. The algebraic approach of \cite{KoeMed03} is particularly
75: useful in the design and analysis of linear network codes, and we
76: adopt the notation and terminology of that paper. 
77: 
78: Random networks codes \cite{HoMed03,HoKoe03} are linear network codes
79: in which the encoding coefficients are chosen randomly from a finite
80: field. The sink nodes can decode correctly if and only if the overall
81: transfer matrix from the sources to each sink is invertible. One of
82: the main theoretical results for random network codes consists of the
83: following lower bound on the probability of successful decoding
84: \cite{HoMed03}, assuming that the underlying network is linearly
85: solvable over $\FF_q$ (i.e. there exists a linear code which satisfies
86: the multicast requirements).  For a network code in which some of the
87: code coefficients are chosen independently and uniformly from a finite
88: field with cardinality $q$, the probability that all $d$ receivers can
89: decode the source processes is at least
90: \begin{equation}\label{eq:bound}
91:   \left(1-\frac{d}{q}\right)^{\nu}
92: \end{equation}
93: where $\nu$ is the maximum number of links receiving signals with
94: independent random coefficients in any set of links constituting a
95: flow solution from all sources to any receiver \cite{HoKoe03}. 
96: 
97: A looser bound (subject to the same conditions as above) which depends
98: only on $\eta$, the total number of edges receiving signals with
99: independent random coefficients is given by \cite{HoMed03,HoMed}
100: \begin{equation}
101:   \label{eq:bound2}
102:   \left(1-\frac{d}{q}\right)^\eta.
103: \end{equation}
104: 
105: Thus provided a linear solution over $\FF_q$ exists in the first
106: place, the probability of successful decoding can be made as close to
107: one as desired, by increasing the field size $q$. The bounds
108: (\ref{eq:bound}) and (\ref{eq:bound2}) rely on the special structure
109: of the determinant polynomial of the transfer matrix of the network.
110: 
111: This paper develops the following new lower bound.
112: 
113: \begin{theorem}
114:   Consider a network code in which $\eta$ edges receive signals with
115:   independent random coefficients chosen independently and uniformly
116:   from a finite field with cardinality $q$.  If there is some choice
117:   of coefficients for these $\eta$ edges that results in a solution
118:   over $\FF_q$ then the probability that all receivers can decode the
119:   source processes is at least
120:   \begin{equation}
121:     \label{eq:newbound}
122:     \left(1-\frac{1}{q}\right)^{\eta}.
123:   \end{equation}
124: \end{theorem}
125: 
126: Our approach for the proof of this theorem is to identify a critical
127: sub-matrix of the Edmonds matrix whose non-singularity is a necessary
128: and sufficient condition for decoding success. This critical matrix is
129: different for each sink in the network.  The new bound results
130: directly from a nesting property of the critical matrices.
131: 
132: In the new bound, the field size $q$ required to attain a given
133: probability of success depends only on the number of edges with random
134: coefficients, and not on the number of sinks. The resulting $d$-fold
135: reduction in the required $q$ could be significant. We emphasize that
136: (\ref{eq:newbound}), like (\ref{eq:bound}) applies only when the
137: underlying network is solvable over $\FF_q$. This is a consequence of
138: the conditions for applicability of the Schwartz-Zippel inequality,
139: which is used in the proof of both bounds. Thus (\ref{eq:newbound})
140: does \emph{not} imply the universal existence of binary solutions for
141: every network. The bounds (\ref{eq:bound}),  (\ref{eq:bound2}) and
142: (\ref{eq:newbound}) only provide lower bounds for a given $q$ when the
143: network is solvable over $\FF_q$.
144: 
145: We further conjecture that for large random networks satisfying
146: certain properties, the success probability behaves as
147: \begin{equation}
148:   \label{eq:conjecture}
149:   \prod_{i=1}^{\numedges} \left(1-\frac{1}{q^i}\right)
150: \end{equation}
151: where $\numedges$ is the total number of links in the network.
152: 
153: The paper is organized as follows: Section \ref{sec:model} presents
154: our model and introduces some algebraic notation.  Section
155: \ref{sec:newbound} develops the new bound (\ref{eq:newbound}), while
156: Section \ref{sec:random} discusses random graphs, leading to the
157: conjecture (\ref{eq:conjecture}).
158: 
159: \section{Network Coding Model}\label{sec:model}
160: We adopt the model from \cite{KoeMed03}. The network is represented by
161: a directed acyclic graph $\graph=(\nodes,\edges)$ with
162: $\numnodes=|\nodes|$ nodes and $\numedges=|\edges|$ edges. There are
163: $r$ independent, discrete source processes with messages
164: belonging to $\FF_q$, and $d\geq 1$ receivers. Each receiver
165: node has $L\geq r$ incoming edges. The multicast requirement is that
166: each receiver node can decode every source message from the signals on
167: its incident edges.
168: 
169: Each edge $\edge\in\edges$ is incident to node $\node\in\nodes$ if
170: $\node=\head(\edge)$, or is an outgoing edge if $\node=\tail(l)$. The
171: in-degree of a node $\node$ is $\din{\node}$ and the out-degree is
172: $\dout{\node}$. The time unit is chosen such that the capacity of each
173: link is one bit per unit time and edges with larger capacity are
174: modeled as parallel edges.  Without loss of generality, it can be
175: assumed that each source is associated with a source node
176: $s_\alpha\in\nodes$ with $\din{s_\alpha}=0$ and $\dout{s_\alpha}=1$,
177: $\alpha=1,2,\dots,r$ . Similarly, each sink node $t_\beta$ has
178: $\din{t_\beta}=r$ and $\dout{t_\beta}=0$, $\beta=1,2,\dots,d$ (it is
179: always possible to obtain such a graph by introducing auxiliary nodes
180: and edges). It will further be assumed that edges are labeled
181: ancestrally.
182: 
183: A \emph{scalar linear network code} for $G$ is an assignment of linear
184: encoding functions $f_v:\FF_q^{\din{v}}\mapsto\FF_q^{\dout{v}}$ to
185: each node $v\in\nodes$. Such codes are sufficient for the multicast
186: problem on acyclic delay networks.  Following \cite{KoeMed03}, define
187: the \emph{encoding matrix} $F\in\FF_q^{\numedges\times\numedges}$
188: where $F_{ij}$ is the coefficient applied to the symbol incoming on
189: edge $i\in\edges$ for contribution to outgoing edge
190: $j\in\edges$. According to the assumption of ancestral ordering, $F$
191: is strictly upper triangular.  Similarly, the \emph{source matrix}
192: $A\in\FF_q^{r\times\numedges}$ maps messages onto outgoing source
193: edges and the \emph{sink matrix}
194: $B_{\beta}\in\FF_q^{r\times\numedges}$ maps incoming sink edges onto
195: the sinks $t_\beta\in\nodes$, $\beta=1, 2,\dots, d$.
196: % Further define
197: % \begin{equation*}
198: % B =
199: % \begin{bmatrix}
200: %   B_1       \\
201: %   B_2
202: %   \vdots    \\
203: %   B_d
204: % \end{bmatrix}
205: % \end{equation*}
206: 
207: Let $x\in\FF_q^{1\times r}$ be a row vector representing the source
208: messages. Then the received vector of symbols
209: $y_\beta\in\FF_q^{1\times r}$ at sink 
210: $\beta=1,2,\dots,d$ is given by
211: \begin{displaymath}
212:   y_\beta = x M_\beta
213: \end{displaymath}
214: where 
215: \begin{equation*}
216:   M_\beta = A (I-F)^{-1} \tran{B_\beta}.  
217: \end{equation*}
218: Each sink can decode all sources if and only if $\det(A (I-F)^{-1}
219: \tran{B_\beta}) \neq 0$ for every $\beta=1,2,\dots,d$, or equivalently
220: if the Edmonds matrix
221: \begin{equation*}
222:   \edmonds_\beta=
223:   \begin{bmatrix}
224:     A & 0 \\
225:     I-F & \tran{B_\beta}
226:   \end{bmatrix}
227: \end{equation*}
228: is non-singular. 
229: 
230: Considering the entries of $A$, $F$ and $B_\beta$ as variables, the
231: Leibniz determinant formula provides a way of writing $\det
232: \edmonds_\beta$ as a multivariate polynomial $P_\beta$ in the $a_{ij},
233: f_{ij}, b_{ij}$. Furthermore, this multivariate polynomial has degree
234: at most $\nu$ but is linear in each variable individually. Therefore
235: the product
236: \begin{equation}
237:   P=\prod_\beta P_\beta\label{eq:P}
238: \end{equation}
239: has degree $d\nu$, with each variable of degree $d$ or less.
240: 
241: The lower bound (\ref{eq:bound}) results from a modified
242: Schwartz-Zippel bound, which takes into account the individual
243: variable degree constraint of $P_\beta$ \cite[Lemma 1]{HoKoe03}. We
244: reproduce this lemma for reference.
245: \begin{lemma}\label{lem:sz}
246:   Let $P$ be a multivariate polynomial of degree $d\nu$, with the
247:   exponent of any individual variable at most $d$. Let each variable
248:   be chosen uniformly from $\FF_q$. Then if $P$ is not identically
249:   zero,
250:   \begin{equation}\label{eq:sz}
251:     \Pr\left(P\neq0\right) \geq \left(1-\frac{d}{q}\right)^\nu.
252:   \end{equation}
253: \end{lemma}
254: 
255: We make two remarks on this approach. First, application of Lemma
256: \ref{lem:sz} to $P$ as defined in (\ref{eq:P}) implies an independence
257: of the events $P_{\beta_1}=0$ and $P_{\beta_2}=0$. Depending on the
258: structure of the network, these events may be strongly dependent. For
259: example, consider $P_{1}=P_{2}=\dots=P_d$, meaning all sinks have
260: identical incoming signals ($B_1=B_2=\dots=B_d$). Then Lemma
261: \ref{lem:sz} yields a lower bound $(1-d/q)^\nu$, rather than
262: $(1-1/q)^\nu$. Obviously this is an extreme example, yet it
263: illustrates the point that (\ref{eq:bound}) may be loose.
264: 
265: Secondly, the modified Schwartz-Zippel bound itself can be very loose,
266: as the following example shows. Let $H\in\FF_q^{m\times m}$ with each
267: entry $h_{ij}$ chosen independently with a uniform distribution on
268: $\FF_q$. Then it is well known that
269: \begin{equation}
270:   \label{eq:fullexact}
271:   \Pr\left(\det H \neq 0\right) = \pi_m(q) = \prod_{i=1}^m
272:   \left(1-q^{-i}\right). 
273: \end{equation}
274: In contrast, Lemma \ref{lem:sz} gives the lower bound
275: \begin{equation}\label{eq:badbound}
276:   \Pr\left(\det H \neq 0\right) \geq \left(1-q^{-1}\right)^m, 
277: \end{equation}
278: which also could be obtained from (\ref{eq:fullexact}) by lower
279: bounding each term in the product by the minimum term
280: $(1-q^{-1})$. 
281: 
282: We emphasize that (\ref{eq:sz}) applies only when $P$ is not
283: identically zero for every choice of variables (e.g. all coefficients
284: are zero). This precludes application of (\ref{eq:sz}) to non-solvable
285: networks, i.e. networks where every choice of $F$ makes $Z_\beta$
286: singular and hence $P=0$.
287: 
288: In Section \ref{sec:newbound} we partially address the dependency
289: between the $P_\beta$, while in Section \ref{sec:random} we consider
290: large random networks, where we also discuss the extent to which
291: (\ref{eq:fullexact}) improves (\ref{eq:badbound}).
292: 
293: \section{The New Bound} \label{sec:newbound} According to our
294: assumption regarding sources and sinks, and the ancestral ordering of edges,
295: we can further assume without loss of generality that
296: \begin{align*}
297:   A &=
298:   \begin{bmatrix}
299:     I_{r\times r} & 0_{r\times (\numedges-r)}
300:   \end{bmatrix}
301:   \\
302:   B_\beta &=
303:   \begin{bmatrix}
304:     0_{r\times k_\beta} & I_{r\times r} & 0_{r\times(\numedges-r-k_\beta)}
305:   \end{bmatrix}, \beta=1,2,\dots,d
306: \end{align*}
307: where $k_1>r$ and $k_\beta > r+k_{\beta-1}$, $\beta>1$. This means
308: that the sources inject messages into the network via edges
309: $1,2,\dots,r$ and that each sink observes signals on $r$ consecutively
310: numbered edges. No sink shares edges with any other sink or
311: source. See Figure \ref{fig:butterfly} for an example of how to
312: arrive at this formulation.
313: 
314: Then the Edmonds matrix for sink $\beta$ has the following structure:
315: \begin{equation}
316:   Z_\beta =
317:   \begin{bmatrix}
318:     I_{r} & 0   & 0   & 0 & 0      \\
319:     U_1           & W_{11} & W_{12} & W_{13} & 0 \\
320:     0             & U_2 & W_{21} & W_{22} & 0 \\
321:     0             & 0   & U_3     & W_{31} & I_{r} \\
322:     0             & 0   & 0       & U_4   & 0
323:   \end{bmatrix}
324: \end{equation}
325: where the $U_i$ are square, upper triangular with diagonal elements all
326: equal to $1$. The matrices $U_1$ and $U_3$ are  $r\times r$, $U_2$ is
327: $(k_\beta-2r)\times(k_\beta-2r)$ and $U_4$ is
328: $(\numedges-r-k_\beta)\times(\numedges-r-k_\beta)$. 
329: 
330: \begin{definition}
331:   The \emph{critical matrix} for sink $\beta$ is the following
332:   $(k_\beta-r)\times(k_\beta-r)$ principal sub-matrix of $Z_\beta$,
333:   \begin{equation}\label{eq:critical}
334:     C_\beta = \begin{pmatrix}
335:       W_{11} & W_{12} \\
336:       U_2 & W_{21}
337:     \end{pmatrix}.
338:   \end{equation}
339: \end{definition}
340: 
341: \begin{lemma}\label{lem:critical}
342: The determinant of the Edmonds matrix for sink $\beta$ has the same
343: magnitude as the determinant of its critical matrix. 
344: \begin{equation*}
345:   |\det Z_\beta| = |\det C_\beta| 
346: \end{equation*}
347: \end{lemma}
348: \begin{proof}
349:   Straightforward from either the Laplace expansion of $\det Z_\beta$,
350:   or repeated application of the partitioned matrix determinant formula.
351: \end{proof}
352: We can immediately apply Lemma \ref{lem:sz} to $\det C_\beta$ to bound
353: the probability for a given sink
354: \begin{equation}\label{eq:onesink}
355:   \Pr\left(\det Z_\beta\neq 0\right) = \Pr\left(\det C_\beta\neq
356:     0\right) \geq \left(1-\frac{1}{q}\right)^{\eta_\beta},
357: \end{equation}
358: where $\eta_\beta$ is the number of columns in $C_\beta$ with variable
359: terms, i.e. the number of edges in the subset $\{r+1,r+2,\dots,k_\beta\}$
360: receiving signals with random coefficients.
361: 
362: For the $d$ receiver problem, we have the following very useful
363: property of the critical matrices, which is guaranteed by their
364: construction.
365: \begin{lemma}[Nesting of critical matrices]\label{lem:nesting}
366:   $C_{\beta_1}$ is a principal
367:   sub-matrix of $C_{\beta_2}$ for $\beta_2 > \beta_1$.  
368: \end{lemma}
369: Hence each critical matrix $C_\beta$ has as nested principal
370: sub-matrices, all the critical matrices for sinks $1,2,\dots,\beta-1$.
371: 
372: \begin{proof}[Proof of main result (\ref{eq:newbound})]
373:   Let $\event_\beta$, $\beta=1,2,\dots,d$ be the event that sink $\beta$ can
374:   decode. By Lemma \ref{lem:critical}, $\event_\beta \iff
375:   \det Z_\beta \neq 0 \iff \det C_\beta \neq 0$. Now the probability
376:   that all sinks can decode is given by
377:   \begin{equation}\label{eq:chainrule}
378:     \Pr\left(\bigcap_{\beta=1}^d \event_\beta\right) =
379:     \Pr(\event_1) \Pr(\event_2\mid\event_1) \dots
380:     \Pr(\event_\beta\mid\event_1\dots\event_{\beta-1})   
381:   \end{equation}
382:   Now consider $\Pr(\event_m\mid\event_1,\dots,\event_{m-1}) = \Pr(\det
383:   C_m\neq 0 \mid \det C_1 \neq 0,\dots,\det C_{m-1}\neq0)$ for some $2\leq
384:   m\leq\beta$. By Lemma \ref{lem:nesting}, $C_m$ can be partitioned
385:   \begin{equation*}
386:     C_m =
387:     \begin{pmatrix}
388:       C_{m-1} & U \\
389:       V & W
390:     \end{pmatrix}
391:   \end{equation*}
392:   for appropriate choices of $U, V, W$. 
393: 
394:   Conditioned on $\det C_{m-1} \neq 0$, we can use the partitioned
395:   matrix determinant formula to write
396:   \begin{equation}\label{eq:partitiondet}
397:     \det C_m = \det(C_{m-1}) \det\left(W - V C_{m-1}^{-1} U\right),
398:   \end{equation}
399:   which (conditioned on $\det C_{m-1} \neq 0$) is zero if and only if
400:   $\det\left(W - V C_{m-1}^{-1} U\right)=0$.
401: 
402:   Let $\phi_m$ be the multivariate polynomial corresponding to $\det
403:   C_m$, and let $\sigma_{m-1}$ be the multivariate polynomial
404:   corresponding to $\det\left(W - V C_{m-1}^{-1} U\right)$. Then from
405:   (\ref{eq:partitiondet}) $\deg\phi_m = \deg\phi_{m-1} +
406:   \deg\sigma_{m-1}$. This relation also holds for the degree of any
407:   individual variable. From the Leibniz formula and the structure of
408:   the Edmonds matrix (as explained previously for $P_\beta$), we also
409:   know that the individual degree of any variable in $\phi_m$ or
410:   $\phi_{m-1}$ is zero or one.  Hence
411:   \begin{equation*}
412:     \deg\sigma_{m-1} = \deg\phi_m - \deg\phi_{m-1},
413:   \end{equation*}
414:   and the degree of any individual variable in $\sigma_{m-1}$ is at
415:   most 1.  Collecting results so far and applying Lemma \ref{lem:sz},
416:   \begin{align*}
417:     \Pr(\event_m\mid\event_1,\dots,\event_{m-1})
418:     &= \Pr\left(\det\left(W - V C_{m-1}^{-1} U\right)\neq 0 \right) \\
419:       &=
420:     \Pr\left(\sigma_{m-1}\neq 0\right) \\ &\leq
421:     \left(1-\frac{1}{q}\right)^{\deg\phi_m - \deg\phi_{m-1}} 
422:   \end{align*}
423:   
424:   Finally, substitution into (\ref{eq:chainrule}) results in a
425:   telescoping sum for the exponents, $\deg\phi_1 + \deg\phi_2 -
426:   \deg\phi_1 + \deg\phi_3 - \deg\phi_2 + \dots$, leaving only
427:   \begin{equation*}
428:     \Pr\left(\bigcap_{\beta=1}^d \event_\beta\right) \geq \left(1 -
429:       \frac{1}{q}\right)^{\deg\phi_d} 
430:   \end{equation*}
431: 
432:   This directly yields (\ref{eq:newbound}) via $d\nu \leq \eta
433:   \triangleq \deg\phi_d = \eta_d \leq \numedges$.
434: \end{proof}
435: 
436: Let 
437: \begin{equation*}
438:   z(d,q)=\frac{\log(1-d/q)}{\log(1-1/q)}.
439: \end{equation*}
440: Then (\ref{eq:newbound}) is tighter than (\ref{eq:bound}) whenever
441: \begin{equation*}
442:   \eta < \nu\, z(d,q).
443: \end{equation*}
444: Furthermore, $z(d,q)>d$ and
445: \begin{align*}
446:   \lim_{q\rightarrow d}z(d,q)&=\infty \\
447:   \lim_{q\rightarrow\infty}z(d,q)&=d.
448: \end{align*}
449: Roughly speaking, the new bound is tighter for networks with
450: $\numedges = O(\nu d)$ and sufficiently small $q$.
451: 
452: In some instances it may be useful to have a bound which depends only
453: on the total number of edges carrying signals with random
454: coefficients. Replacing $\nu$ with $\eta$ in (\ref{eq:bound}) results
455: in (\ref{eq:bound2}) which is looser than (\ref{eq:newbound}), since
456: \begin{equation*}
457:    \left(1-d/q\right)^\eta < \left(1-1/q\right)^\eta.
458: \end{equation*}
459: 
460: Note that successful decoding at a particular sink $\beta$ in general
461: depends on only part of $C_\beta$. There can be a much smaller
462: sub-matrix that determines singularity, for example, $C_\beta$ might
463: be block diagonal, with successful decoding of sink $\beta$ depending
464: only on one of the blocks (this case arises when there are disjoint
465: paths from the sources to each sink). Thus $C_\beta$ may be larger
466: than strictly required for analysis of sink $\beta$ alone, however
467: defining the critical matrix this way yields the nesting property that
468: results in the new bound.
469: 
470: \section{Example: The Butterfly Network}
471: Figure \ref{fig:butterfly} shows the well-known butterfly network,
472: with additional nodes and edges introduced in order to satisfy our
473: assumptions on sources and sinks. The source $s$ has $r=2$ messages,
474: and the edge labels indicate the edge ordering. Edges $1$ and $2$
475: carry the two messages from the source, while edges $12$ resp. $13$
476: duplicate the signals on edges $5$ resp. $10$, and edges $14$
477: resp. $15$ duplicate $8$ resp. $11$. Supposing that all other edges
478: carry random linear combinations of signals, $\nu=7$ and $\eta = 9$.
479: 
480: \begin{figure}[htbp]
481:   \centering
482:   \includegraphics*[scale=0.7]{butterfly}
483:   \caption{The butterfly network.}
484:   \label{fig:butterfly}
485: \end{figure}
486: 
487: Figure \ref{fig:critical} shows the structure of the Edmonds matrix
488: $Z_1$, and the nested critical matrices $C_1$ and $C_2$. To see how
489: the nesting arises, $B_2$ has been placed alongside. For clarity, most
490: of the zeros have been omitted from each matrix. The solid disks
491: represent random entries of $F$.
492: 
493: \begin{figure}[htbp]
494:   \centering
495:   \includegraphics*[width=0.8\columnwidth]{critical}
496:   \caption{Critical matrices for the butterfly network.}
497:   \label{fig:critical}
498: \end{figure}
499: 
500: Figure \ref{fig:plot} shows the empirically measured probability of
501: decoding success versus the field size $q$ for the network of Figure
502: \ref{fig:butterfly} (filled circles). This was achieved using monte-carlo
503: simulation, selecting each of the coefficients uniformly from $\FF_q$.
504: Results for the first ten prime fields are shown. Also shown are the
505: existing bounds (\ref{eq:bound}), dashed line, (\ref{eq:bound2}), solid
506: line, and the new bound (\ref{eq:newbound}), dot-dashed line. In this
507: case, the new bound is considerably tighter.
508: 
509:  
510: %  The matrices $A$,
511: % $B_1$, $B_2$ and $F$ have the following form.
512: % \begin{align*}
513: %   A &=
514: %   \begin{bmatrix}
515: %     I_2 & 0_{13}
516: %   \end{bmatrix} \\
517: %   B_1 &=
518: %   \begin{bmatrix}
519: %     0_{11} & I_2 & 0_{2}
520: %   \end{bmatrix} \\
521: %   B_2 &=
522: %   \begin{bmatrix}
523: %     0_{13} & I_2 
524: %   \end{bmatrix} \\
525: % I-F &=
526: % \left[
527: % \begin{array}{ccccccccccccccc}
528: %  1&&x&x&&&&&&&&&&& \\
529: %  &1&x&x&&&&&&&&&&& \\
530: %  &&1&&x&x&&&&&&&&& \\
531: %  &&&1&&&x&x&&&&&&& \\
532: %  &&&&1&&&&&&&x&&& \\
533: %  &&&&&1&&&x&&&&&& \\
534: %  &&&&&&1&&x&&&&&& \\
535: %  &&&&&&&1&&&&&&&x \\
536: %  &&&&&&&&1&x&x&&&& \\
537: %  &&&&&&&&&1&&&x&& \\
538: %  &&&&&&&&&&1&&&x& \\
539: %  &&&&&&&&&&&1&&& \\
540: %  &&&&&&&&&&&&1&& \\
541: %  &&&&&&&&&&&&&1& \\
542: % \end{array}\right]
543: % \end{align*}
544: 
545:  
546: \begin{figure}[htbp]
547:   \centering
548:   \setlength{\unitlength}{1mm}
549:   {\begin{picture}(85,55)
550:     \put(0,0){\includegraphics*[width=0.9\columnwidth]{plot}}
551:     \put(85,0){\makebox(0,0){$q$}}
552:     \put(0,55){\makebox(0,0){$p$}}
553:   \end{picture}}
554:   \caption{Success probability $p$ versus field size $q$ compared to
555:     bounds (\ref{eq:bound}), (\ref{eq:bound2}) and
556:     (\ref{eq:newbound}) for the butterfly network.}
557:   \label{fig:plot}
558: \end{figure}
559: 
560: \section{Random Graphs}\label{sec:random}
561: Successful decoding for a particular sink $\beta$ depends on the
562: non-singularity of its critical matrix $C_\beta$. To obtain
563: (\ref{eq:newbound}) we used Lemma \ref{lem:sz} to bound the
564: probability that this matrix is non-singular. It is interesting to
565: consider however circumstances under which (\ref{eq:fullexact}) might
566: be applicable, providing an even tighter bound.
567: 
568: There are two main obstacles to the application of
569: (\ref{eq:fullexact}) for determination of the probability that $\det
570: C_\beta\neq 0$. Firstly, (\ref{eq:fullexact}) applies to ``full''
571: matrices, with each element chosen independently and uniformly from
572: $\FF_q$.  In contrast, $C_\beta$ is of the form (\ref{eq:critical}),
573: with all elements below the $r$-th diagonal equal to zero (the
574: strictly lower triangular part of $U_2$). Secondly, the non-zero
575: elements in the upper portion (upper triangular part of $U_2$ and all
576: of $W_{11}$, $W_{12}$ and $W_{21}$) of $C_\beta$ are determined by the
577: topology of the network itself. For a sparsely connected network, the
578: proportion of zeros in this part of the matrix will greatly exceed
579: $1/q$.
580: 
581: Assuming that the random network code coefficients are chosen from the
582: non-zero elements of $\FF_q$, the total number of non-zero elements in
583: $F$ is
584: \begin{equation*}
585:   \sigma \triangleq \sum_{\node\in\nodes} \din{\node}\dout{\node} \leq
586:   \numedges^2. 
587: \end{equation*}
588: Let $\rho=\sigma/E^2$ be the proportion of non-zero elements. Ignoring
589: the structure required by (\ref{eq:critical}), generate a random
590: $m\times m$ matrix $C^{(m)}$ with elements identically distributed
591: according to
592: \begin{equation*}
593:   \Pr\left(c_{ij}=f\right) =
594:   \begin{cases}
595:     1-\rho & f=0 \\
596:     \frac{\rho}{q-1} & f \neq 0
597:   \end{cases}
598: \end{equation*}
599: 
600: It is a remarkable fact that provided $\rho$ does not tend to zero or one
601: too quickly with $m$, 
602: \begin{equation*}
603:   \lim_{m\rightarrow\infty} \Pr\left(\det C^{(m)}\neq 0\right) = \pi_m(q).
604: \end{equation*}
605: See \cite{Coo00} for a discussion of this threshold
606: effect. Conditioned on the event that $C^{(m)}$ has no all-zero rows
607: or columns (if it did, the network flow would anyway be infeasible
608: regardless of choice of code), the requirement is
609: \begin{equation*}
610:   \rho > \frac{1}{m}\left(\frac{1}{2}\log m + \log\log m\right).
611: \end{equation*}
612: This result even holds for independent, but non-identically
613: distributed entries, as discussed by Cooper \cite{Coo00}.  
614: 
615: Now for sufficiently small $\rho$, $C^{(m)}$ can be permuted with high
616: probability into the form (\ref{eq:critical}). This leads us to
617: conjecture that there exist conditions on $\sigma$ such that
618: $\pi_m(q)$ is the success probability for a large, randomly generated
619: network with a given degree distribution. The remainder of this
620: section analyzes some properties of $\pi_m(q)$, and demonstrates the
621: improvement that may be obtained compared to (\ref{eq:badbound}).
622: 
623: To guarantee a particular probability $p$ using (\ref{eq:badbound}),
624: the field size $q$ must satisfy
625: \begin{equation*}
626:   q \geq \frac{1}{1-p^{1/m}} =  \frac{1}{2}+ m \log \frac{1}{p} +
627:   O\left(\frac{1}{m}\right).  
628: \end{equation*}
629: Hence the required field size increases linearly with the size of the
630: matrix.
631: 
632: % \begin{equation*}
633: %   \lim_{m\rightarrow\infty}\left(1-1/q\right)^m = 0
634: % \end{equation*}
635: 
636: Let $\pi_\infty(q)=\lim_{m\rightarrow\infty}\pi_m(q)$ then
637: \begin{equation*}
638:   \pi_\infty(q) = \prod_{i=1}^\infty\left(1-q^{-i}\right) = q^{1/24}
639:   \left(\frac{1}{2} 
640:   \vartheta_1'\left(q^{-1/2}\right)\right)^{1/3},
641: \end{equation*}
642: where $\vartheta_1$ is the Jacobi theta function \cite[Equation
643: 8.181.3]{GraRyz94} and
644: \begin{align*}
645:   \vartheta_1'(q) &= \left.\frac{\partial}{\partial
646:       z}\vartheta_1(z,q)\right|_{z=0} \\
647:   &= 2\sum_{i=0}^\infty (-1)^i (1+2i) q^{-\frac{1}{2}(i+\frac{1}{2})^2}.
648: \end{align*}
649: Truncating the latter series gives the following lower bound,
650: \begin{equation*}
651:   \pi_\infty(q) \geq \left(1-\frac{3}{x}\right)^{1/3}.
652: \end{equation*}
653: This lower bound is compared to $\pi_\infty$ for the first 20 primes
654: in Figure \ref{fig:bound}. 
655: \begin{figure}
656:   \centering
657:   \setlength{\unitlength}{1mm}
658:   {\begin{picture}(80,55)
659:     \put(0,0){\includegraphics*[width=80mm]{fg}}
660:     \put(80,1){\makebox(0,0){$q$}}
661:     \put(1,50){\makebox(0,0){$\pi_\infty$}}
662:   \end{picture}}
663:   \caption{Lower bound (solid line) and $\pi_\infty(q)$ (dots).}
664:   \label{fig:bound}
665: \end{figure}
666: For a given probability $p$ in (\ref{eq:fullexact}), the required
667: field size $q$ for $m\rightarrow\infty$ satisfies
668: \begin{equation*}
669:   q \geq \frac{3}{1-p^3}.
670: \end{equation*}
671: which does not depend on $m$.
672: 
673: \section{Concluding remarks}
674: \label{sec:conclusion}
675: Random network coding is a promising decentralized approach for
676: multicast. One of the main implementation considerations is the size
677: of the finite field required to achieve a specified probability that
678: every sink can decode every source. This paper presented a new bound
679: on the success probability, which in certain circumstances is tighter
680: that the previous bound. We also presented a heuristic argument that
681: motivates the investigation of tighter bounds for large random
682: networks, based on the distribution of rank of large random finite
683: field matrices. 
684: 
685: \section*{Acknowledgments}
686: This work was performed while A. Tauste Campo was visiting the
687: Institute for Telecommunications Research.  This work was supported by
688: the Australian Government under grant DP0557310, and by the Defence
689: Science and Technology Organisation under contracts 4500485167 and
690: 4500550654. The authors would like to thank Ian Grivell and Terence
691: Chan and for helpful discussions.
692: 
693: \bibliographystyle{IEEEtran}
694: \bibliography{network,alex}
695: 
696: \end{document}
697: