1:
2: \documentclass[11pt,dvips,draftcls,onecolumn]{IEEEtran}
3:
4: \usepackage{psfig,amsfonts,amsmath,amssymb,color,hyperref}
5: \hypersetup{colorlinks=true,linkcolor=black,urlcolor=dblue,pdfpagemode=None,pdfstartview=FitH}
6: \definecolor{gray}{cmyk}{.2,0.2,.3,.1}
7: \definecolor{dred}{cmyk}{0,0.9,0.4,0.3}
8: \definecolor{dblue}{rgb}{0,0,0.5}
9: \definecolor{dgreen}{rgb}{0,0.3,0}
10: \definecolor{dgray}{rgb}{0.3,0.3,0}
11: %\renewcommand{\familydefault}{cmss}
12: \DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm}
13: \DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf}
14: \DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt}
15: \DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf}
16: \DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit}
17: \DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl}
18: \DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc}
19: \newtheorem{theorem}{Theorem}
20: \newtheorem{lemma}{Lemma}
21: \newtheorem{proposition}{Proposition}
22: \newtheorem{example}{Example}
23: \newtheorem{definition}{Definition}
24: \newcommand{\rend}{\hfill$\square$}
25: \newcommand{\tend}{\hfill$\blacksquare$}
26: \newtheorem{remark}{Remark}
27:
28:
29: \title{On Multiflows in Random Unit-Disk Graphs, and the Capacity of
30: Some Wireless Networks
31: \footnote{The authors are with
32: the School of Electrical and Computer Engineering, Cornell University,
33: Ithaca, NY. URL:
34: \href{http://cn.ece.cornell.edu/}{{\tt http://cn.ece.cornell.edu/}}.
35: Work supported by the National Science Foundation, under awards
36: CCR-0238271 (CAREER), CCR-0330059, and ANR-0325556. Parts of this
37: work were presented at the 2003 edition of ACM MobiHoc, in Annapolis,
38: MD~\cite{PerakiS:03}; and at the 2004 edition of the Information
39: Theory Workshop, in San Antonio, TX~\cite{PerakiS:04}.}}
40: \author{Christina Peraki \hspace{2cm} Sergio D.\ Servetto}
41: \date{March 20, 2005.}
42:
43:
44: \begin{document}
45: \maketitle
46:
47: \begin{picture}(0,0)
48: \put(-10,210){\tt\small Submitted to the IEEE Transactions on Information
49: Theory, March 2005.}
50: \end{picture}
51:
52: \begin{abstract}
53: \noindent\it
54: We consider the capacity problem for wireless networks. Networks
55: are modeled as random unit-disk graphs, and the capacity problem is
56: formulated as one of finding the maximum value of a multicommodity
57: flow. In this paper, we develop a proof technique based on which
58: we are able to obtain a tight characterization of the solution to
59: the linear program associated with the multiflow problem, to within
60: constants independent of network size. We also use this proof method
61: to analyze network capacity for a variety of transmitter/receiver
62: architectures, for which we obtain some conclusive results. These
63: results contain as a special case (and strengthen) those of Gupta
64: and Kumar for random networks, for which a new derivation is provided
65: using only elementary counting and discrete probability tools.
66: \end{abstract}
67:
68:
69: \section{Introduction}
70: \label{sec:intro}
71:
72: \subsection{The Capacity of Wireless Networks -- Five Years Later}
73:
74: In March 2000 (exactly five years ago as of the writing of this paper),
75: Gupta and Kumar published a landmark piece of work, where they presented
76: a thorough study on the capacity of wireless networks~\cite{GuptaK:00}.
77: For {\em random} networks, this problem was formulated as one of
78: forming tessellations of a sphere, then defining routes in between
79: cells, for which tight upper and lower bounds were obtained on their
80: capacity. The main finding in~\cite{GuptaK:00} was actually a rather
81: negative one: under a variety of very reasonable scenarios, in all
82: cases the throughput available to each node in the network was found
83: to be of the form $\Theta\big(\frac{1}{\sqrt{n}}\big)$ {\em at most},
84: for a network with $n$ nodes -- that is, this throughput becomes
85: vanishingly small for large networks.
86:
87: The results of~\cite{GuptaK:00} generated a flurry of activity in
88: this area (surveyed below, in subsection~\ref{sec:related-work}).
89: However, five years later, although some progress has been made
90: towards understanding the capacity of large networks in a regime
91: in which the minimum distance among nodes remains fixed and the
92: area covered grows unbound with the number of nodes, some questions
93: related to the original setup in~\cite{GuptaK:00}, dealing with
94: {\em high-density} networks (meaning, networks with a growing
95: number of nodes covering a fixed finite area) still remain, at
96: best, only partially answered:
97: \begin{itemize}
98: \item The ability to generate directed beams of energy in a wireless
99: network could potentially change its behavior rather drastically,
100: making the network ``look like'' a wired one. What exactly
101: is the impact of directional antennas then on network capacity?
102: \item Despite some attempts, a pure information theoretic analysis
103: on the capacity of high-density wireless networks still remains
104: elusive.
105: \end{itemize}
106: In this paper, we revisit the problem of capacity for random
107: networks considered in~\cite{GuptaK:00}. We consider an entirely
108: different problem formulation: our formalization of the network
109: capacity problem consists of finding the value of a multicommodity
110: flow problem defined on a random graph, for which we are able to
111: obtain a number of results that contain those of~\cite{GuptaK:00}
112: as a special case, generalizing them in a number of interesting
113: directions.
114: %In particular, we provide complete answers to the
115: %question of directional antennas, and we provide partial but very
116: %promising answers to the question of how to formulate the setup
117: %of~\cite{GuptaK:00} in pure information-theoretic terms.
118:
119: \subsection{Problem Formulation}
120:
121: Consider the following network communication problem. $n$ nodes are
122: uniformly distributed on the closed set $[0,1]\times[0,1]$, forming
123: a random graph $G=(V,E)$. Each node $s_i$ can only send messages to
124: and receive messages from nodes within distance $d_n$, where $d_n$,
125: in order for the graph to be connected with probability 1 (as
126: $n \rightarrow \infty$), has to satisfy
127: \begin{equation}
128: \pi d_n^2 = \frac{\log n+\xi_n}{n},
129: \label{eq:min-conn-radius}
130: \end{equation}
131: for some $\xi_n\rightarrow\infty$~\cite{GuptaK:98}. Source-destination
132: pairs are formed randomly: for each source node $s_i$ one destination
133: node $t_i$ is chosen by sampling uniformly (without replacement) from
134: the set of network nodes ($1 \leq i \leq n$) -- each node is both a
135: source, a destination for some other node, and a relay for other nodes.
136: All links have the same fixed finite capacity $c$, independent of network
137: size. This scenario is illustrated in Fig.~\ref{fig:problem-setup}.
138:
139: \begin{figure}[ht]
140: \centerline{\psfig{file=problem-setup.eps,width=12cm,height=2.5cm}}
141: \caption{\small Problem setup. $n$ randomly located transmitters
142: send data to $n$ randomly chosen receivers, all nodes act as
143: sources/destinations/relays, and nodes can only exchange messages
144: with nearby nodes (within range $d_n$).}
145: \label{fig:problem-setup}
146: \end{figure}
147:
148: Our goal in this paper is to determine the rate of growth of the {\em
149: maximum stable throughput} (MST) for the network~\cite{TsybakovM:79}---the
150: rate at which all sources can inject packets, while maintaining stability
151: for the system---and provided all sources inject data at the same rate.
152:
153: The problem of determining MST under a fairness constraint is an
154: instance of a {\em multicommodity flow} problem~\cite[Ch.\ 29]{CormenLRS:01}:
155: \begin{itemize}
156: \item There are $n$ commodities: the packets available for transmission
157: from transmitter $s_i$ to receiver $t_i$.
158: \item The load on a single link contributed by all sources that use that
159: link cannot exceed its capacity.
160: \item Subject to these constraints, we want to find the largest number
161: of packets per unit of time that can be injected simultaneously by all
162: sources.
163: \end{itemize}
164: Representing our network by a graph $G=(V,E)$, the capacity of an
165: edge $e=(u,v)$ by $c(u,v)$, and letting our optimization variables be
166: $f_i(u,v)$ (the flow along edge $(u,v)$ for the $i$-th commodity), then
167: the maximum multiflow problem above can be formulated as a
168: linear program, as shown in Table~\ref{tab:lp-multicommodity}.
169:
170: \begin{table}[ht]
171: \caption{\small\rm Linear programming formulation of the multicommodity
172: flow problem with a fairness constraint.}
173: \begin{center}
174: \fbox{\begin{minipage}{8.6cm}
175: max \hspace{2cm} $\lambda_n$ \\
176: $\;$ subject to:
177: \[\begin{array}{rll}
178: & \lambda_n = \sum_{(s_i,v)\in E} f_i(s_i,v), & 1 \leq i \leq n \\
179: & \sum_{i=1}^n f_i(u,v) \leq c(u,v), & (u,v) \in E \\
180: & f_i(u,v) = -f_i(v,u), & (u,v)\in E, 1\leq i\leq n \\
181: & \sum_{v\in V} f_i(u,v) = 0, & u\in V-\{s_i,t_i\}, 1\leq i\leq n
182: \end{array}\]
183: \end{minipage}}\end{center}
184: \label{tab:lp-multicommodity}
185: \end{table}
186:
187: Our main task in this paper is to provide a characterization of the
188: optimal value $\lambda^*_n$. Note that since the graph is random, and
189: the LP is a function of the random graph, $\lambda^*_n$ is a random
190: variable itself.
191:
192: \subsection{Asymptotically Tight Bounds}
193:
194: Not much is known about the structure of optimal solutions to the
195: maximum multiflow problem---the only technique we are aware
196: of for deciding whether a particular amount of flow of each commodity
197: can be supported by the network consists of formulating this problem
198: as a linear program, and then answering the non-emptyness question
199: for its polytope of optimization using a standard LP solver (e.g.,
200: the Ellipsoid method~\cite{GroetschelLS:88}), or some of the efficient
201: algorithms for maximum multiflow such as that of Karger and
202: Plotkin~\cite{KargerP:95}. Hence, we will not be able to use those
203: formulations to do much more than obtain numerical results for
204: our problem. We are thus motivated to search for an alternative
205: formulation of the problem. And one such possible alternative is
206: illustrated in Fig.~\ref{fig:special-case}.
207:
208: \begin{figure}[ht]
209: \centerline{\psfig{file=special-case.eps,width=12cm,height=3cm}}
210: \caption{\small In this formulation, we only consider the traffic
211: generated by sources on the left-half of the network, with destination
212: on the right-half---the traffic generated by all other
213: source/destination pairs is discarded.}
214: \label{fig:special-case}
215: \end{figure}
216:
217: Note that doing this amounts to introducing a restriction in the domain
218: of optimization of the linear program from Table~\ref{tab:lp-multicommodity}:
219: instead of considering all possible network flows, we only consider those
220: which satisfy the constraints of Fig.~\ref{fig:special-case}. But what
221: is crucial in this case is that, different from the problem of
222: Fig.~\ref{fig:problem-setup}, this new problem involving flow going
223: from the left to the right admits a regular {\em single commodity} flow
224: formulation. The resulting linear program is shown in
225: Table~\ref{tab:lp-regflow}.
226:
227: \begin{table}[ht]
228: \caption{\small\rm Linear program for the single commodity flow problem.}
229: \begin{center}\fbox{\begin{minipage}{7cm}
230: max \hspace{1.5cm} $n\nu_n$ \\
231: subject to:
232: \[\begin{array}{rll}
233: & n\nu_n = \sum_{u\in V} f(s,u), & u \in V \\
234: & f(u,v) \leq c(u,v), & (u,v) \in E \\
235: & f(u,v) = -f(v,u), & (u,v)\in E \\
236: & \sum_{v\in V} f(u,v) = 0, & u\in V-\{s,t\} \\
237: \end{array}\]
238: \end{minipage}}\end{center}
239: \label{tab:lp-regflow}
240: \end{table}
241:
242: The interest in this new linear program is due to the fact that,
243: since it corresponds to a classical single commodity problem, we
244: can try to solve it analytically using the max-flow/min-cut
245: theorem~\cite{FordF:62}. However, the relationship between the
246: optimal value $\lambda^*_n$ for the ``difficult'' multicommodity
247: problem, and $n\nu^*_n$ for the ``easier'' single commodity problem,
248: is not entirely straightforward.
249: %\footnote{Note: the value of the
250: %single commodity problem is scaled by a factor or $n$, so as to
251: %keep $\nu_n$ and $\lambda_n$ in the same units, and hence directly
252: %comparable.}
253: On one hand, the linear program in
254: Table~\ref{tab:lp-regflow} is a {\em restriction} of that in
255: Table~\ref{tab:lp-multicommodity}, since in the former some flow
256: variables are constrained to 0 (that is how we incorporate the
257: constraint of flow going only from left to right). On the other
258: hand, the linear program in Table~\ref{tab:lp-regflow} is a
259: {\em generalization} of that in Table~\ref{tab:lp-multicommodity},
260: since the latter removes the multicommodity constraints (all
261: commodities are treated as a single commodity). Thus, an
262: important question is that of giving a precise relationship between
263: $\lambda^*_n$ and $\nu^*_n$.
264:
265: \subsection{Related Work}
266: \label{sec:related-work}
267:
268: This work is primarily motivated by our struggle to understand the
269: results of Gupta and Kumar on the capacity of wireless
270: networks~\cite{GuptaK:00}. And the main idea behind our approach
271: is simple: the transport capacity problem posed in~\cite{GuptaK:00},
272: in the context of random networks, is essentially a throughput
273: stability problem---the goal is to determine how much data can be
274: injected by each node into the network while keeping the system
275: stable---, and this throughput stability problem admits a very
276: simple formulation in term of flow networks. Note also that because
277: of the mechanism for generating source/destination pairs, all
278: connections have the same average length (one half of one network
279: diameter), and thus we do not need to deal with the bit-meters/sec
280: metric considered in~\cite{GuptaK:00}.
281:
282: As mentioned before,~\cite{GuptaK:00} sparked significant interest
283: in these problems. Follow up results from the same group were reported
284: in~\cite{GuptaK:01, XieK:02}. Some information theoretic bounds
285: for large-area networks were obtained in~\cite{LevequeT:05}. When
286: nodes are allowed to move, assuming transmission delays proportional
287: to the mixing time of the network, the total network throughput is
288: $O(n)$, and therefore the network can carry a non-vanishing rate per
289: node~\cite{GrossglauserT:02}. Using a linear programming formulation,
290: non-asymptotic versions of the results in~\cite{GuptaK:00} are given
291: in~\cite{ToumpisG:02}; an extended version of that work can be found
292: in~\cite{Toumpis:PhD}. An alternative method for deriving transport
293: capacity was presented in~\cite{KulkarniV:04}. The capacity of large
294: Gaussian relay networks was found in~\cite{GastparV:05}. Preliminary
295: versions of our work based on network flows have appeared
296: in~\cite{PerakiS:03, PerakiS:04}; and network flow techniques have
297: been proposed to study network capacity problems (cf.,
298: e.g.,~\cite{AhlswedeCLY:00}, \cite[Ch.\ 14.10]{cover-thomas:it-book}),
299: and network coding problems~\cite{KoetterM:03}. From the network
300: coding literature, of particular relevance to this work is the work
301: on multiple unicast sessions~\cite{LiL:04b}.
302:
303: \subsection{Main Contributions and Organization of the Paper}
304:
305: Let $\lambda^*_n$ denote an optimal solution to the linear program
306: in Table~\ref{tab:lp-multicommodity}, and let $\nu^*_n$ denote an
307: optimal solution to the linear program in Table~\ref{tab:lp-regflow}.
308: Our first result consists of finding the asymptotic value of
309: $\lambda^*_n$: with probability $1$ as $n\to\infty$,
310: \begin{equation}
311: \Theta(\lambda^*_n) \;\; = \;\; \Theta(\nu^*_n)
312: \;\; = \;\; \Theta\left(\frac{\log^{\frac 3 2}(n)}{\sqrt{n}}\right).
313: \label{eq:main-result}
314: \end{equation}
315: This result formally establishes the equivalence between the two
316: linear programs, in a well defined sense: they both have solutions
317: which differ by at most a constant factor, independent of network
318: size.
319:
320: A second important contribution is to show an application of the
321: proof methods developed to establish~(\ref{eq:main-result}), to
322: obtain the maximum stable throughput for various transmitter/receiver
323: architectures:
324:
325: \begin{itemize}
326:
327: \item We consider first the case of omnidirectional antennas, where
328: we show that the scaling laws obtained based on our proof method
329: are identical to those of~\cite{GuptaK:00}: per-node throughput is
330: $\Theta\big(\frac{1}{\sqrt{n\log n}}\big)$.
331:
332: \item Then we apply the same proof techniques to the determination
333: of scaling laws for a new architecture, in which transmitter nodes
334: can generate a single and arbitrarily narrow directed beam, and in
335: which receivers can successfully decode multiple transmissions as
336: long as the transmitters are not co-linear. And in this case we
337: find that:
338: \begin{itemize}
339: \item If only enough power to maintain the network connected is radiated
340: at each node, the maximum stable throughput of this network is
341: $\Theta\left(\!\sqrt{\frac{\log n}{n}}\,\right)$.
342: \item If now enough power is radiated to achieve MST
343: linear in network size (certainly feasible with narrow beams), then
344: the number of {\em resolvable} beams that
345: each node must generate is $\Theta(n)$.
346: \end{itemize}
347:
348: \item Finally, we consider a node architecture in which each node
349: is able to generate {\em multiple and arbitrarily narrow} directed
350: beams, simultaneously to all nodes within its transmission range,
351: and receivers operate as above. In this case we find that:
352: \begin{itemize}
353: \item If only enough power to maintain the network connected is
354: radiated at each node, the maximum stable throughput of this network
355: is $\Theta\big(\!\log^{\frac 3 2}n \big/ \sqrt{n}\big)$.
356: \item If now enough power is radiated to achieve MST
357: linear in network size (certainly feasible with
358: narrow beams), then the number of {\em resolvable} beams
359: that each node must generate is $\Theta(n^{\frac 1 3})$.
360: \end{itemize}
361:
362: \end{itemize}
363:
364: Essentially, our results show that both directional antennas, as well
365: as the ability to communicate simultaneously with multiple nodes, can
366: only achieve modest improvements in terms of achievable MST. While
367: some performance gains are certainly feasible at reasonable complexities
368: (in the order of a low-degree polynomial in $\log n$), the number of
369: {\em resolvable} beams that need to be generated to increase the
370: achievable MST by more than a polylog factor is polynomial in network
371: size, and thus exponential in the minimum number of beams required to
372: keep the
373: network connected. How many beams need to be resolved is a reasonable
374: measure of complexity, since the higher this number, the narrower
375: these beams need to be made, and hence the higher the complexity
376: of a practical implementation.
377:
378: We also believe another original contribution is given by our proof
379: techniques:
380: \begin{itemize}
381: \item Our results are obtained using only elementary network flow
382: concepts, and the calculations involved require only basic probability
383: theory, calculus and combinatorics. By formulating the
384: problem of~\cite{GuptaK:00} as an elementary problem of flows
385: in random graphs, we obtain what we believe is a number of interesting
386: insights into the nature of this problem which were not obvious
387: to us from their proof technique, as well as a set of meaningful
388: generalizations to deal with the case of directional antennas.
389: \item Except for some elements of the protocol model considered
390: in~\cite{GuptaK:00}, most of the work we are aware of on this
391: subject (e.g.,~\cite{GastparV:05, GrossglauserT:02, GuptaK:01,
392: LevequeT:05, ToumpisG:02, XieK:02}), has focused on the use
393: of ``continuous'' tools, dealing with Gaussian signals, power
394: constrained channels, etc. Our work instead takes a ``discrete''
395: approach to the network capacity problem, tackling it primarily
396: using flow, counting and discrete probability tools. Thus, we
397: believe our proof technique, while using only elementary tools,
398: has some novelty in the context of the problem considered here.
399: \end{itemize}
400: An added benefit of our proof method is that we are able to prove
401: {\em strong} convergence (meaning, convergence with probability 1)
402: in all cases. In particular, when considering the specialization
403: of our results to the setup of~\cite{GuptaK:00}, our results are
404: stronger, in that only {\em weak} convergence is established there.
405: %A discussion on a pure information-theoretic interpretation of our
406: %results, exploiting recently discovered connections between network
407: %flow theory and network information theory~\cite{BarrosS:03e},
408: %is included in Section~\ref{sec:conclusions}.
409:
410: The rest of this paper is organized as follows. In
411: Section~\ref{sec:bounds-formulation} we formulate upper and lower
412: bounds on the value of $\lambda_n^*$: the upper bound is evaluated
413: in Section~\ref{sec:eval-upper-bound}, and the lower bound is
414: evaluated in Section~\ref{sec:eval-lower-bound}. Then,
415: applications of these results in the context of wireless networking
416: problems follow: in Section~\ref{sec:omnidirectional} we present
417: an alternative derivation for the results of Gupta and Kumar in
418: the context of random networks~\cite{GuptaK:00}, and in
419: Section~\ref{sec:directional} these results are extended to deal
420: with two different cases involving directional antennas. The
421: paper concludes with Section~\ref{sec:conclusions}.
422:
423:
424: \section{Asymptotically Tight Bounds on the Value of the Linear Program}
425: \label{sec:bounds-formulation}
426:
427: In this section we start with some preliminaries presenting the
428: tools used to carry out our analysis, to then go on to formulate
429: upper and lower bounds on the value of $\lambda_n^*$.
430:
431: \subsection{Tools}
432: \label{sec:tools}
433:
434: To compute the maximum value of a single commodity flow in our network,
435: we use a standard result in flow networks: the max-flow/min-cut
436: theorem of Ford and Fulkerson~\cite{FordF:62}. We solve this problem
437: by counting how many edges can be constructed so that they all
438: simultaneously straddle a minimum cut.
439:
440: \subsubsection{The Max-Flow/Min-Cut Theorem}
441:
442: $f$ is a flow of maximum value {\em iff} $|f| = c(S,T)$ (for some cut $(S,T)$).
443: We focus our attention on one particular cut (shown also in
444: Fig.~\ref{fig:our-cut}):
445: \begin{eqnarray*}
446: S & = & { (x_i,y_i)\in V \cap [0,\mbox{\small $\frac 1 2$})\times[0,1]}, \\
447: T & = & { (x_i,y_i)\in V \cap [\mbox{\small $\frac 1 2$},1] \times [0,1] }.
448: \end{eqnarray*}
449:
450: \begin{figure}[ht]
451: \vspace{-2mm}
452: \centerline{\psfig{file=our-cut.eps,height=3cm,width=12cm}}
453: \vspace{-2mm}
454: \caption{\small To illustrate the choice of a cut to derive bounds.
455: $L$ and $R$ are sections of the network on each side of the cut
456: boundary, of width $d_n$, the transmission range.}
457: \label{fig:our-cut}
458: \end{figure}
459:
460: In this way, to compute the value of a maximum flow we need to
461: determine how many edges straddle the $x=\frac 1 2$ cut. To do
462: that, we proceed in two steps. First, we compute the {\em expected}
463: number of edges that straddle this cut, with this expectation taken
464: as an ensemble average over all possible network realizations. Then,
465: we derive a sharp concentration result: given an arbitrary network
466: realization, with probability 1 as $n\rightarrow\infty$, we show
467: that in this network, the actual number of edges that straddle the
468: cut has the exact same rate of growth (in the $\Theta$ sense
469: of~\cite{GrahamKP:94}) as the ensemble mean does.
470:
471: \subsubsection{Mean Values}
472:
473: What is the average number of nodes in a subset
474: $A\subseteq[0,1]\times[0,1]$? A simple calculation shows
475: that
476: \begin{equation}
477: E(\textsl{Number of nodes in }A) = n P(A)
478: = n \int_A f_{XY} \mbox{d}A = n|A|,
479: \label{eq:num-nodes}
480: \end{equation}
481: where $|A|$ denotes the area of $A$.
482:
483: \subsubsection{Chernoff Bounds}
484:
485: In addition, to prove sharp concentration results, we need to bound
486: the probability of deviations from its mean by sums of independent
487: random variables:
488: \begin{itemize}
489: \item Consider $n$ points $X_{1} \ldots X_{n}$ iid and uniformly distributed
490: on $[0,1] \times [0,1]$. We have a number of
491: subsets $A_j \subset [0,1] \times [0,1]$, for $j=1...f(n)$ (the number
492: of subsets may depend on the number of points $n$), and denote the area
493: of any such subset by $|A_j|$. Now we define some random variables:
494: \[ N_{ij} = \left\{ \begin{array}{ll}
495: 1, & X_i \in A_j \\
496: 0, & \textrm{otherwise.}
497: \end{array} \right.
498: \]
499: Since the $X_i$'s are independent, the $N_{ij}$'s are also independent.
500: \item Now let $N_j$ be another random variable defining the number of points
501: in $A_j$, i.e., $N_j=\sum_{i=1}^n N_{ij}$. We see in this case that the
502: $N_j$'s, $j=1 \ldots f(n)$ are random variables where each is the sum of
503: $n$ iid binary random variables (but not necessarily independent among
504: the $N_j$'s themselves).
505: \item The expected number of points in $A_{j}$ is
506: \[ E(N_j) \;\; = \;\; E\left(\sum_{i=1}^n N_{ij}\right)
507: \;\; = \;\; \sum_{i=1}^n E(N_{ij}). \]
508: But, since $P(X_i \in A_j) = |A_j|$, we have that
509: $E(N_{ij}) = 1|A_j|+0(1-|A_j|) = |A_j|$, and hence
510: $E(N_j) = n|A_j|$.
511: \end{itemize}
512:
513: For the family of variables $N_j$, we have the following standard
514: results, known as the {\em Chernoff} bounds (see, e.g.,~\cite[Ch.\
515: 4]{MotwaniR:95}):
516: \begin{enumerate}
517: \item For any $\delta >0$:
518: \[ P\big[N_j>(1+\delta)n|A_j|\big]
519: \;\; < \;\; \left(\frac{e^{\delta}}{(1+\delta)^{1+\delta}}\right)^{n|A_j|}.
520: \]
521: \item For any $0<\delta <1$:
522: \[ P\big[N_j<(1-\delta)n|A_j|\big] \;\; < \;\; e^{-\frac{1}{2}n|A_j|\delta^{2}}.
523: \]
524: \end{enumerate}
525: With a few simple calculations we can rewrite the first bound as
526: \begin{eqnarray*}
527: \lefteqn{P\big[(N_j-n|A_j|)>\delta n|A_j|\big]
528: \;\; < \;\;
529: \left(\frac{e^{\delta}}{e^{(1+\delta)\ln(1+\delta)}}\right)^{n|A_j|}} \\
530: & = & \left(e^{\delta - (1+\delta)\ln(1+\delta)}\right)^{n|A_j|}
531: \;\; = \;\; e^{(\delta - (1+\delta)\ln(1+\delta))n|A_j|}
532: \;\; = \;\; e^{-\theta_{1} n|A_j|},
533: \end{eqnarray*}
534: where $-\theta_{1} \triangleq \delta-(1+\delta)\ln(1+\delta)$. We can
535: also rewrite the second bound as
536: \[ P\big[(N_j - n|A_j|)<-\delta n|A_j|\big]
537: \;\; < \;\; e^{(-\frac{1}{2}\delta^{2})n|A_j|}
538: \;\; = \;\; e^{ -\theta_{2} n|A_j|},
539: \]
540: where $-\theta_2 \triangleq -\frac{1}{2} \delta^{2}$. Consider
541: now the case of $0<\delta<1$: restricted to this range, we have
542: that $\theta_1 > 0$; and $\theta_2$ is clearly positive as well.
543: Thus, by defining $\theta(\delta) = \min(\theta_1,\theta_2)$,
544: we have
545: \begin{equation}
546: P\big[\;|N_j-n|A_j|| > \delta n|A_j|\;\big] \;\; < \;\; e^{-\theta n|A_j|}.
547: \label{eq:chernoff}
548: \end{equation}
549: Our interest in~(\ref{eq:chernoff}) is because, if we can prove
550: probability bounds of that form, then we can claim that
551: $\frac{N_j}n = \Theta(|A_j|)$ with probability 1, in the limit
552: as $n\to\infty$. In other words, for the random variables $N_j$,
553: as $n\to\infty$, there exist constants such that deviations from
554: their mean by more than these constants occur with probability 0.
555: Note that as $n\to\infty$, $e^{-\theta n|A_j|} \rightarrow 0$, so
556: \[\lim_{n\to\infty} P\big[\;|N_j-n|A_j|| > \delta n|A_j|\;\big] \;\;=\;\; 0, \]
557: or equivalently,
558: \begin{eqnarray*}
559: 1 & = & \lim_{n\to\infty} P\big[\;|N_j-n|A_j|| \leq \delta n|A_j|\;\big] \\
560: & = & \lim_{n\to\infty} P\big[\;0 \leq (1-\delta)n|A_j| \leq N_j
561: \leq (1+\delta) n|A_j|\;\big] \\
562: & = & \lim_{n\to\infty} P\big[N_j=\Theta(n|A_j|)\;\big],
563: \end{eqnarray*}
564: where $\Theta$ is defined in~\cite{GrahamKP:94} as:
565: \[
566: \Theta(g(n)) \;\; = \;\;
567: \big\{ f(n) : \exists c_1>0, c_2>0, n_0, \textrm{ for which } 0 \leq
568: c_1 g(n) \leq f(n) \leq c_2 g(n) \textrm{, } \forall n \geq n_0. \big\}.
569: \]
570:
571: \subsection{An Equivalent Linear Program}
572: \label{sec:ssd}
573:
574: As suggested in the Introduction, we will not work directly with
575: the original linear program from Table~\ref{tab:lp-multicommodity},
576: but instead we will work with a new linear program, one in which
577: flow is constrained to move from left to right. The new LP is
578: formally stated in Table~\ref{tab:demand-zero}.
579:
580: \begin{table}[!ht]
581: \caption{\small\rm Linear programming formulation of the multicommodity
582: flow problem, in which the supply of connections other than those going
583: from left to right are set to 0.}
584: \begin{center}\fbox{\begin{minipage}{12cm}
585: max \hspace{2cm} $\ell_n$ \\
586: $\;$ subject to:
587: \[\begin{array}{rll}
588: & \ell_n = \sum_{(s_i,v)\in E} f_i(s_i,v), & 1\leq i\leq n \textrm{ and }\\
589: && s_i \in S=\{u \in V: u \in [0,\frac{1}{2})\times[0,1]\},\\
590: && t_i \in T=\{u \in V: u \in [\frac{1}{2},1]\times[0,1]\},\\
591: & \sum_{i=1}^n f_i(u,v) \leq c(u,v), & (u,v) \in E \\
592: & f_i(u,v) = -f_i(v,u), & (u,v)\in E, 1\leq i\leq n \\
593: & \sum_{v\in V} f_i(u,v) = 0, & u\in V-\{s_i,t_i\}, 1\leq i\leq n \\
594: & f_i(s_i,v) = 0,
595: & \forall s_i\in T=\{u\in V:u\in[\frac{1}{2},1]\times[0,1]\},\\
596: && t_i\in S=\{u\in V:u\in[0,\frac{1}{2})\times[0,1]\}.
597: \end{array}\]
598: \end{minipage}}\end{center}
599: \label{tab:demand-zero}
600: \end{table}
601:
602: Considering sources on the left half of the network and destinations
603: on the right, essentially says that in our linear programming formulation
604: in Table~\ref{tab:lp-multicommodity} we must add the constraint of
605: setting to 0 the demands of commodities such that either the source
606: is located on the right or the sink is located on the left. But this
607: constraint changes the result of the linear program only by a constant
608: factor, and therefore asymptotically we get the same values from
609: Table~\ref{tab:lp-multicommodity} and Table~\ref{tab:demand-zero}.
610: Intuitively, the reason is simple: since nodes are uniformly distributed,
611: we should have about $n/2$ nodes in $S$ and about $n/2$ nodes in $T$
612: with high probability; at the same time, since the source/destination
613: pairs are uniformly distributed, about $n/4$ of the sources are placed
614: on the left side of the network with destinations on the right side;
615: therefore, by considering only traffic generated by sources in $S$
616: for destinations only in $T$ the value of the multicommodity problem
617: should at most decrease by a factor of 4, and hence remains of the
618: same order. To see this more formally, consider the following indicator
619: variables:
620: \[ I^{(n)}_i = \left\{ \begin{array}{ll}
621: 1, & s_i \in S \wedge t_i \in T \\
622: 0, & \textrm{otherwise,}
623: \end{array} \right.
624: \]
625: where $S$ and $T$ are given in Table~\ref{tab:demand-zero}. Then,
626: $I^{(n)}=\sum_{i=1}^{n} I^{(n)}_i$ is another random variable whose
627: value is equal to the number of pairs with the source on the left half
628: and the sink on the right half. We would like to compute how many
629: are these pairs, to calculate the difference between the values of
630: the two linear programs. To do this, we first compute the mean of
631: $I^{(n)}$, then we use the Chernoff bounds to prove a sharp concentration
632: of this variable around its mean.
633:
634: We start by computing $E(I^{(n)})$. We have:
635: \[ E(I^{(n)}) \;\; = \;\; E\left(\sum_{i=1}^n I^{(n)}_i\right)
636: \;\; = \;\; \sum_{i=1}^n E(I^{(n)}_i), \]
637: due to linearity of expectation. Now,
638: \[ E(I^{(n)}_i)
639: \;\; = \;\; 1 \cdot P(s_i \in S \wedge t_i \in T)
640: + 0 \cdot P(s_i \in T \vee t_i \in V)
641: \;\; = \;\; P(s_i \in S \wedge t_i \in T). \]
642: Since the nodes in our network are uniformly and independently
643: distributed, we have that the events $\{s_i\in S\}$ and $\{t_i\in T\}$
644: are independent events, and therefore:
645: \[ E(I^{(n)}_i) \;\; = \;\; P(s_i \in S \wedge t_i \in T)
646: \;\; = \;\; P(s_i \in S) \cdot P(t_i \in T)
647: \;\; = \;\; \frac{1}{2} \times \frac{1}{2}
648: \;\; = \;\; \frac{1}{4},\]
649: which finally gives us:
650: \[ E(I^{(n)}) \;\; = \;\; \sum_{i=1}^n E(I^{(n)}_i)
651: \;\; = \;\; \frac{n}{4}.\]
652: This expected number occurs with high probability as $n\to\infty$ because,
653: according to the Chernoff bound,
654: \[
655: P\big[\;|I^{(n)}-E(I^{(n)})| > \delta E(I^{(n)})\;\big]
656: \;\; < \;\; e^{-\theta E(I^{(n)})}
657: \;\; = \;\; e^{-\theta \frac{n}{4}} \;\; \to \;\; 0,
658: \]
659: as $n \rightarrow \infty$. Thus, with high probability, there
660: are about $\frac{n}{4}$ sources in $S$ with destinations in $T$.
661: As a result, from the fairness constraint we have that
662: $P(\ell_n^* = \lambda_n^*/4)\to 1$, and herefore,
663: $\Theta(\lambda_n^*) = \Theta(\ell_n^*)$.
664:
665: \subsection{Formulation of the Bounds}
666:
667: With these tools, it is easy to describe asymptotic upper and lower bounds
668: on $\lambda_n^*$:
669: \begin{itemize}
670: \item To obtain an upper bound, we eliminate the multicommodity
671: constraints from the linear program in Table~\ref{tab:demand-zero},
672: and use the max-flow/min-cut theorem to compute the value of a maximum
673: flow. Thus, we have that $\Theta(\ell_n^*) \leq \Theta(\nu_n^*)$,
674: for $\nu_n$ as defined in Table~\ref{tab:lp-regflow}.
675: \item To obtain a lower bound, we construct a feasible point for the
676: linear program in Table~\ref{tab:demand-zero}, to obtain a value
677: $\gamma_n$ for the LP for which, clearly, $\gamma_n \leq \ell_n^*$.
678: \end{itemize}
679: In the next two sections we evaluate these bounds, to show that
680: \[
681: \Theta\left(\frac{\log^{\frac 3 2}n}{\sqrt{n}}\right)
682: \;\; = \;\; \gamma_n
683: \;\; \leq \;\; \ell_n^*
684: \;\; \leq \;\; \Theta(\nu_n^*)
685: \;\; = \;\; \Theta\left(\frac{\log^{\frac 3 2}n}{\sqrt{n}}\right),
686: \]
687: and thus conclude that
688: $\ell_n^* = \Theta\left(\frac{\log^{\frac 3 2}n}{\sqrt{n}}\right)$,
689: and thus
690: $\lambda_n^* = \Theta\left(\frac{\log^{\frac 3 2}n}{\sqrt{n}}\right)$
691: as well.
692:
693:
694: \section{Evaluation of the Upper Bound}
695: \label{sec:eval-upper-bound}
696:
697: In this section, our goal is to show that $\nu_n^* =
698: \Theta\left(\frac{\log^{\frac 3 2}n}{\sqrt{n}}\right)$, based on
699: the methods outlined in the previous section.
700:
701: \subsection{Counting Edges Across a Minimum Cut}
702: \label{sec:upperb-average}
703:
704: Fix a particular node on the left side of the minimum cut, $L$.
705: The number of edges that cross the cut for that one node is exactly
706: the number of nodes in the right side of the cut, $R$, that are
707: within distance $d_{n}$. Therefore, for an arbitrary point $p=(x,y)$
708: in $L=[\frac 1 2-d_n,\frac 1 2)\times [0,1]$, we draw a circle of
709: radius $d_n$ and center $(x,y)$. The points $q=(u,v)$ in $R=[\frac
710: 1 2,\frac 1 2+d_n]\times [0,1]$ that are inside the circle are equal
711: to the number of edges we want to count. These points $p$ and $q$
712: for which an edge exists satisfy the following conditions: (1)
713: $\frac 1 2-d_n \leq x \leq \frac 1 2$; (2) either (a) $0 \leq y \leq
714: 1$, or (b) $d_n \leq y \leq 1-d_n$; (3) $\frac 1 2 < u$; and (4)
715: $(u-x)^2 + (v-y)^2 \leq d_n^2$.
716: The situation is illustrated in Fig.~\ref{fig:multiple-beams-count}.\\
717:
718: \begin{figure}[ht]
719: \centerline{\psfig{file=count1.eps,width=12cm,height=2.7cm}}
720: \caption{\small To illustrate constraints on edges.}
721: \label{fig:multiple-beams-count}
722: \end{figure}
723:
724: For each $p=(x,y)$, we get the average number of points $q=(u,v)$
725: within the shaded arc $Q_p$ in Fig.~\ref{fig:multiple-beams-count}
726: using eqn.~(\ref{eq:num-nodes}): $E(${\sl Number of points in} $Q_p)
727: = n|Q_p|$.
728:
729: To compute the area of $Q_p$ (denoted $|Q_p|$), we let $\vartheta$
730: denote the angle of the arc illustrated in
731: Fig.~\ref{fig:multiple-beams-count}. Then, it follows from
732: elementary trigonometric identities that
733: $\sin{\frac{\pi-\vartheta}{2}}=\frac{\frac{1}{2}-x}{d_{n}}$, and so
734: $\cos{\frac{\vartheta}{2}}=\frac{\frac{1}{2}-x}{d_{n}}$. So, $|Q_p|
735: = \mbox{\small$\frac 1 2$} \vartheta d_{n}^2
736: -\mbox{\small$\frac 1 2$}
737: d_n\cos{\frac{\vartheta}{2}}2d_{n}\sin{\frac{\vartheta}{2}}
738: = \mbox{\small$\frac 1 2$} \vartheta d_n^2
739: - \mbox{\small$\frac 1 2$}d_n^2\sin{\vartheta}
740: = \mbox{\small$\frac 1 2$}d_n^2 (\vartheta - \sin{\vartheta})$.
741: And plugging this expression into $n|Q_p|$, we get $n|Q_p| = n
742: \frac{1}{2} d_n^2 (\vartheta - \sin{\vartheta})$. But from the
743: trigonometric identities above, we have that $\vartheta = 2
744: \arccos{\frac{\frac{1}{2}-x}{d_{n}}}$ and hence, $\sin{\vartheta} =
745: 2 \sin{\frac{\vartheta}{2}} \cos{\frac{\vartheta}{2}}$, which
746: implies $\sin^2{\vartheta} =
747: 4\sin^2{\frac{\vartheta}2}\cos^2{\frac{\vartheta}2}$, which again
748: implies $\sin^2{\vartheta}
749: = 4 (1 -\cos^2{\frac{\vartheta}{2}}) \cos^2{\frac{\vartheta}{2}}$.
750: Now, since $0 \leq \vartheta \leq \pi$, $\sin{\vartheta}$ $=$ $2
751: \cos{\frac{\vartheta}{2}}
752: \sqrt{1 -\cos^2{\frac{\vartheta}{2}}}$ $\geq$ $0$,
753: and so, finally, we get an expression for $n|Q_p|$ in terms of $n$,
754: $d_n$, and the coordinates of the transmitter $p=(x,y)$:
755: \begin{eqnarray*}
756: n|Q_p|
757: & = & \frac{1}{2} n d_n^2 (\vartheta - \sin{\vartheta}) \\
758: & = & \frac{1}{2} n d_n^2 \left(2 \arccos{\frac{\frac{1}{2}-x}{d_{n}}} -
759: 2\cos{\frac{\vartheta}{2}}\sqrt{1-\cos^2{\frac{\vartheta}{2}}}\right)\\
760: & = & n d_{n}^2 \left(\arccos{\frac{\frac{1}{2}-x}{d_{n}}} -
761: \frac{\frac{1}{2}-x}{d_{n}}
762: \sqrt{1 -\frac{(\frac{1}{2}-x)^2}{d_{n}^2}}\right).
763: \end{eqnarray*}
764:
765: The result above is the average number of edges that cross
766: the cut, starting at a fixed point $p=(x,y)$ in $L$. To calculate
767: the total number of edges $S$ that cross the cut on average, we need
768: to add up $n|Q_p|$ over all nodes $p$, (i.e., compute $S =
769: \sum_{p\in L} n|Q_p|$).
770: And our plan to do this is to approximate this sum by an integral.
771:
772: The value of $|Q_p|$ is clearly dependent on the location of $p$:
773: for $p$'s in $L$ near the boundary of the cut ($x<\approx\frac 1
774: 2$), $\vartheta\approx\pi$ and hence the shaded area is large; for
775: $p$'s still in $L$ but far from the boundary of the cut
776: ($x>\approx\frac 1 2-d_n$), $\vartheta\approx 0$ and hence the
777: shaded area is small. Furthermore, except near the top and bottom
778: boundaries, the area of $Q_p$ is independent of $y$. Therefore, to
779: obtain a simple expression for the sought sum, our first step
780: consists of dividing $L$ into $\frac{d_n}\Delta$ thin strips of
781: height 1 and width $\Delta$ (for $\Delta\ll d_n$), and expanding
782: $\sum_{p\in L} n|Q_p|$ in two different ways:
783: \begin{eqnarray*}
784: S_a & = & \sum_{k=1}^{d_n/\Delta} n|Q_{xy}| \cdot\underbrace{|\{
785: p=(x,y)\in L:
786: 0\leq y\leq 1 \}|}_{s_a}; \\
787: S_b & = & \sum_{k=1}^{d_n/\Delta} n|Q_{xy}| \cdot\underbrace{|\{
788: p=(x,y)\in L:
789: d_n\leq y\leq 1-d_n \}|}_{s_b};
790: \end{eqnarray*}
791: (in both cases, we take $\frac 1 2-d_n+(k-1)\Delta\leq x\leq \frac 1
792: 2-d_n+k\Delta$). $S_a$ is an upper bound on $\sum_{p\in L} n|Q_p|$,
793: since we may count edges that end up outside the network; $S_b$ is a
794: lower bound, since we may not count some valid edges close to the
795: network boundary; but as long as $d_n \rightarrow 0$ as
796: $n\rightarrow\infty$, both bounds become tight
797: and equal to $\sum_{p\in L} n|Q_p|$.
798:
799: The next step is to observe that once again we can approximate the
800: size estimates $s_a$ and $s_b$ using eqn.~(\ref{eq:num-nodes}): $s_a
801: = n\Delta$ and $s_b = n(1-2d_n)\Delta$. Hence we get:
802: \begin{eqnarray*}
803: S_a & = & \sum_{k=1}^{d_n/\Delta} n|Q_{xy}|\cdot n\Delta
804: \;\;\; = \;\;\; n^2\Delta\sum_{k=1}^{d_n/\Delta} |Q_{xy}| \\
805: & \approx & n^2\int_{x=\frac 1 2-d_n}^{\frac 1 2}\int_{y=0}^1
806: |Q_{xy}|\mbox{d}x\mbox{d}y; \\
807: S_b & = & \sum_{k=1}^{d_n/\Delta} n|Q_{xy}|\cdot n(1-2d_n)\Delta
808: \;\;\; = \;\;\; n^2(1-2d_n)\Delta\sum_{k=1}^{d_n/\Delta} |Q_{xy}| \\
809: & \approx & n^2\int_{x=\frac 1 2-d_n}^{\frac 1 2}\int_{y=d_n}^{1-d_n}
810: |Q_{xy}|\mbox{d}x\mbox{d}y,
811: \end{eqnarray*}
812: since $\Delta\sum_{k=1}^{d_n/\Delta} |Q_{xy}|$ is a Riemann sum
813: that, as we let $\Delta\rightarrow 0$, converges to the integral
814: over an
815: appropriate region of $|Q_p|$.
816:
817: And now we are almost done. Since $S_b \leq \sum_{p\in L} n|Q_p|
818: \leq S_a$, and we have that for $n$ large, $S_a \approx S_b \approx
819: n^2\int_L |Q_p|$, we finally get:
820: \begin{eqnarray*}
821: \lefteqn{\sum_{p\in L} n|Q_p| \;\;\approx\;\; n^2\int_L |Q_p|\mbox{ d}p} \\
822: & = & n^2 \int_{\frac{1}{2}-d_{n}}^{\frac{1}{2}}
823: \int_{0}^{1} d_{n}^2 \left\lbrack \arccos{\frac{\frac{1}{2}-x}{d_{n}}}
824: -
825: \frac{\frac{1}{2}-x}{d_{n}} \sqrt{1 -\frac{(\frac{1}{2}-x)^2}{d_{n}^2}}
826: \right\rbrack \mbox{d}y \mbox{d}x \\
827: & = & n^2 d_{n}^2 \int_{\frac{1}{2}-d_{n}}^{\frac{1}{2}} \int_{0}^{1}
828: \arccos{\frac{\frac{1}{2}-x}{d_{n}}}\mbox{ d}y \mbox{d}x
829: - n^2 d_{n}^2
830: \int_{\frac{1}{2}-d_{n}}^{\frac{1}{2}}
831: \int_{0}^{1} \frac{\frac{1}{2}-x}{d_{n}}
832: \sqrt{1 -\frac{(\frac{1}{2}-x)^2}{d_{n}^2}}\mbox{ d}y \mbox{d}x \\
833: & = & n^2 d_{n}^2 \int_{\frac{1}{2}-d_{n}}^{\frac{1}{2}}
834: \arccos{\frac{\frac{1}{2}-x}{d_{n}}} \mbox{ d}x
835: - n^2 d_{n}^2
836: \int_{\frac{1}{2}-d_{n}}^{\frac{1}{2}} \frac{\frac{1}{2}-x}{d_{n}}
837: \sqrt{1 -\frac{(\frac{1}{2}-x)^2}{d_{n}^2}} \mbox{ d}x \\
838: & \stackrel{(a)}{=} & -n^2 d_{n}^3 \int_{1}^{0} \arccos{u} \mbox{ d}u +
839: n^2 \int_{1}^{0} u \sqrt{ 1 - u^2} \mbox{ d}u \\
840: & = & n^2 d_{n}^3 \int_{0}^{1} \arccos{u} \mbox{ d}u -
841: n^2 \int_{0}^{1} u \sqrt{1 - u^2} \mbox{ d}u \\
842: & = & n^2 d_{n}^3 - \mbox{\small $\frac 1 3$}n^2d_n^3
843: \;\; = \;\; \mbox{\small $\frac 2 3$}n^2d_n^3,
844: \end{eqnarray*}
845: where $(a)$ follows from the change of variable
846: $\frac{\frac{1}{2}-x}{d_{n}}=u$.
847:
848: \subsection{Sharp Concentration Results}
849:
850: Our next goal is to show that the actual number of edges
851: straddling the cut in any realization of the network is sharply
852: concentrated around its mean. That is, in almost all networks, the
853: number of edges across the cut is $\Theta(n^2d_n^3) =
854: \Theta(\sqrt{n} \log^\frac{3}{2}(n)).$
855:
856: Define a binary random variable $N_{ij}$, which takes the value 1
857: if the $i$-th node is within the transmission range of a node at
858: coordinates $(x_j,y_j)$ on the other side of the cut, as illustrated
859: in Fig.~\ref{fig:multiple-beams-count}:
860: \[ N_{ij} = \left\{ \begin{array}{rl}
861: 1, & X_i \in Q_{(x_j,y_j)} \\
862: 0, & \textrm{otherwise.}
863: \end{array}
864: \right.
865: \]
866: Let $p$ denote the probability that $X_i$ is in $Q_{(x_j,y_j)}$
867: (i.e., that $N_{ij} = 1$). Then, $p = |Q_{(x_j,y_j)}| = \frac 1 2
868: d_n^2(\vartheta-\sin(\vartheta))$, with $0\leq\vartheta\leq\pi$ is
869: as in Fig.~\ref{fig:multiple-beams-count}. Therefore, defining
870: $\kappa_\vartheta$ as $\frac 1 2 (\vartheta-\sin(\vartheta))$, we
871: have $p = |Q_{(x_j,y_j)}| = \kappa_\vartheta d_n^2 =
872: \kappa_\vartheta\frac{\log n}n$.
873:
874: Define $N_j = \sum_{i=1}^n N_{ij}$ as the number of points in
875: $Q_{(x_j,y_j)}$. In this case, we have $E(N_j)$ = $E\big(\sum_{i=1}^n
876: N_{ij}\big)$ = $\sum_{i=1}^n p\cdot 1 + (1-p)\cdot 0$ = $np$ =
877: $\kappa_\vartheta\log(n)$. Now, by eqn.~(\ref{eq:chernoff}), we have
878: that
879: \[
880: P\big(|N_j-\kappa_\vartheta\log(n)| > \delta\kappa_\vartheta\log(n)\big)
881: \;\; < \;\; e^{-\theta\kappa_\vartheta\log(n)}
882: \;\; = \;\; n^{-\theta\kappa_\vartheta},
883: \]
884: As $n\rightarrow\infty$ this probability tends to zero, and
885: therefore, in almost all network realizations, a node on the left
886: side of the cut is connected to $\Theta(\log(n))$ nodes on the
887: right side.\footnote{Observe that $\kappa_\vartheta=0$ only over
888: a set of measure zero (the set of network locations such that
889: $x=\frac 1 2 - d_n$), and thus the exponent can be assumed
890: strictly positive.}
891: By an analogous argument, we have $\Theta(nd_n) =
892: \Theta\big(\sqrt{n\log n}\big)$ nodes on the left half.
893: Therefore, the actual number of edges across
894: the cut is $\Theta(\log^2 n)\cdot\Theta\big(\sqrt{n\log n}\big)$,
895: so $n\nu_n^* = \Theta\big(\sqrt{n}\log^{\frac 3 2} n\big)$.
896:
897:
898: \section{Evaluation of the Lower Bound}
899: \label{sec:eval-lower-bound}
900:
901: To give a lower bound for $\ell_n^*$, we construct one feasible
902: point: this is accomplished by giving a specific routing algorithm,
903: and finding how much traffic this scheme can carry:
904: \begin{enumerate}
905: \item We start by proving that, with probability 1 as $n\to\infty$,
906: there is a subgraph of the random graphs under consideration with
907: a clear, regular structure.
908: \item We then develop a (very simple) routing technique that makes
909: use of the links in the regular subgraph only.
910: \item Finally, we determine the throughput achieved in this way.
911: \end{enumerate}
912:
913: \subsection{Existence of a Regular Subgraph}
914:
915: Consider a partition of the network area (the closed set
916: $[0,1]\times[0,1]$) into {\em square} cells, each one of area
917: $c\frac{\log n}{n}$. To determine $c$, we observe that
918: the side of a cell is $\sqrt{\frac{c\log n}{n}}$, and so $c$
919: is chosen such that the inverse of this number,
920: $\sqrt{\frac{n}{c\log n}}$, is an integer -- this is to
921: guarantee that the cells form a partition of the whole
922: network, as illustrated in Fig.~\ref{fig:regular-subgraph}.
923:
924: \begin{figure}[!ht]
925: \centerline{\psfig{file=regular-subgraph.eps,height=6cm}}
926: \caption{\small To illustrate the presence of a structured subgraph
927: for large $n$. Consider the shaded center cell: all nodes within
928: that cell are connected by an edge to every node in the cells above,
929: below, left and right. To guarantee that all such edges can be
930: formed, from Pithagoras, the transmission range $d_n$ must be
931: chosen as $d_n = \sqrt{\frac{5c\log n}{n}}$. So, provided $c>0$,
932: connectivity of the network is guaranteed~\cite{GuptaK:98}.}
933: \label{fig:regular-subgraph}
934: \end{figure}
935:
936: With the construction of grid and choice of connectivity radius
937: $d_n$ shown in Fig.~\ref{fig:regular-subgraph}, each node within
938: a cell will have an edge connecting it to all nodes in the four
939: adjacent cells.
940:
941: \subsection{A Routing Algorithm}
942:
943: To describe the routing algorithm, we define first some notation:
944: \begin{itemize}
945: \item Given a cell $(i,j)$
946: ($1\leq i,j\leq \sqrt{\frac 1 c \frac{n}{\log n}}$), $v_{ij}$ denotes
947: any arbitrary node $v\in V$ contained in that cell.
948: \item Given a node $v\in V$, $(i(v),j(v))$ denotes the cell that contains $v$.
949: \item $v^{\textrm{\tiny curr}}$: current node; $v^{\textrm{\tiny dest}}$:
950: destination node.
951: \end{itemize}
952: The algorithm executed at each node to decide the next hop of a message
953: is as follows:
954: \begin{enumerate}
955: \item If $j(v^{\textrm{\tiny curr}}) < j(v^{\textrm{\tiny dest}})$,
956: send message to $(i(v^{\textrm{\tiny curr}}),j(v^{\textrm{\tiny curr}})+1)$.
957: \item Else, if $i(v^{\textrm{\tiny curr}}) > i(v^{\textrm{\tiny dest}})$,
958: send message to $(i(v^{\textrm{\tiny curr}})-1,j(v^{\textrm{\tiny curr}}))$.
959: \item Else, if $i(v^{\textrm{\tiny curr}}) < i(v^{\textrm{\tiny dest}})$,
960: send message to $(i(v^{\textrm{\tiny curr}})+1,j(v^{\textrm{\tiny curr}}))$.
961: \item Else, $v^{\textrm{\tiny curr}}$ and $v^{\textrm{\tiny dest}}$ are in
962: the same cell (so $v^{\textrm{\tiny dest}}$ is reachable in one hop from
963: $v^{\textrm{\tiny curr}}$), hence stop.
964: \end{enumerate}
965: These mechanics are illustrated in Fig.~\ref{fig:routing-mechanics}.
966: \begin{figure}[!ht]
967: \centerline{\psfig{file=routing-mechanics.eps,height=5cm}}
968: \caption{\small To illustrate routing mechanics. A source in a cell
969: left of the center cut sends messages to a destination on the right
970: side by first forwarding data horizontally, then vertically. Under
971: the assumption that all cells are non-empty, there is always a next
972: hop.}
973: \label{fig:routing-mechanics}
974: \end{figure}
975:
976: For this algorithm to work properly, we must insure that no cells
977: are empty: if this condition holds, then we can be sure that an L-shaped
978: path as shown in Fig.~\ref{fig:routing-mechanics} will always deliver
979: packets to destination. But the fact that no cells are empty is not
980: obvious, and requires proof. In fact, we will prove something stronger:
981: the number of nodes contained in any arbitrary cell is $\Theta(\log n)$,
982: with high probability as $n\to\infty$.
983:
984: Define $X_{ij}$ as the number of nodes within a cell
985: $(i,j)$ ($1\leq i,j\leq \sqrt{\frac 1 c \frac{n}{\log n}}$). Then,
986: \begin{itemize}
987: \item Mean value of $X_{ij}$:
988: \begin{equation}
989: E(X_{ij}) = n \cdot \frac{c\log n}{n} = c\log n.
990: \label{mean_sensors_per_node}
991: \end{equation}
992: \item Chernoff bound on deviations from the mean for $X_{ij}$:
993: \[
994: P(|X_{ij}-c\log n| > \delta c\log n)
995: \;\; \leq \;\; e^{-\theta c\log n}
996: \;\; = \;\; \frac 1{n^{c\theta}}.
997: \]
998: \item Probability that the occupancy of {\em none} of the cells
999: deviates significantly from its mean:
1000: \begin{eqnarray*}
1001: P\left(\bigcap_{i,j} |X_{ij}-c\log n| < \delta c\log n\right)
1002: & = & 1 - P\left(\bigcup_{i,j}
1003: |X_{ij}-c\log n| > \delta c\log n\right) \\
1004: & \geq & 1 - \sum_{i,j}
1005: P\big(|X_{ij}-c\log n| > \delta c\log n\big).
1006: \end{eqnarray*}
1007: Consider now any $\epsilon>0$; there is a value $n_0(\epsilon)$ such that,
1008: for all $n>n_0(\epsilon)$,
1009: \[
1010: \sum_{i,j} P\big(|X_{ij}-c\log n| > \delta c\log n\big)
1011: \;\; < \;\; \sum_{i,j} \frac 1{n^{c\theta}}
1012: \;\; \stackrel{(a)}{=} \;\; \frac 1 c \frac 1{n^{c\theta-1}\log n}
1013: \;\; < \;\; \epsilon,
1014: \]
1015: where $(a)$ follows from the fact that there are
1016: $\frac 1 c \frac n{\log n}$ cells, and provided $c > \frac 1 \theta$.
1017: Therefore,
1018: \begin{eqnarray*}
1019: P\left(\bigcap_{i,j} |X_{ij}-c\log n| < \delta c\log n\right)
1020: & \geq & 1-\epsilon,
1021: \end{eqnarray*}
1022: and thus all cells contain $\Theta(\log n)$ nodes almost surely, as
1023: $n\to\infty$.
1024: \end{itemize}
1025: With this, we see that all cells {\em simultaneously} will be non-empty
1026: in almost all networks. Thus, the routes defined by the proposed routing
1027: algorithm will always deliver data to destination. We still need to
1028: determine how much though.
1029:
1030: \subsection{Computation of the Achievable Throughput}
1031:
1032: The last step is to determine the throughput available to a
1033: source/destination pair constructed by the routing algorithm above.
1034:
1035: We start by stating the relatively straightforward fact that, since
1036: the routes constructed by the algorithm above do not split the flow
1037: at any intermediate node, the throughput of a connection is determined
1038: by the capacity available to that connection at the link with highest
1039: load.\footnote{This intuitive fact can be formalized based on Robacker's
1040: decomposition theorem for multiflows~\cite{PerakiS:04, Robacker:56}.}
1041: Now, since links have a fixed finite capacity, and since in our problem
1042: we work under a {\em fairness} constraint that forces all source/destination
1043: pairs to inject the same amount of data, we have that the capacity allocated
1044: to a commodity on any link is $\frac {\textrm{\tiny raw link capacity}}
1045: {\textrm{\tiny \# of commodities using that link}}$. Thus, our problem
1046: reduces to finding the maximum number of commodities sharing a link.
1047:
1048: We claim that no link in the network is shared by more connections
1049: than the links which straddle the center cut:
1050: \begin{itemize}
1051: \item The number of commodities sharing a link across the center cut is
1052: exactly equal to the number of nodes within a horizontal strip, as
1053: illustrated in Fig.~\ref{fig:horizontal-links}.
1054: \begin{figure}[!ht]
1055: \centerline{\psfig{file=horizontal-links.eps,width=15cm,height=3.4cm}}
1056: \caption{\small Each node left of the center cut generates traffic that
1057: must cross that cut.}
1058: \label{fig:horizontal-links}
1059: \end{figure}
1060: \item Clearly, the number of commodities on horizontal links (meaning,
1061: links going from one cell to another cell either left or right) decreases
1062: as we move away from the center cut:
1063: \begin{itemize}
1064: \item Moving left, the number of sources decreases.
1065: \item Moving right, once a connection reaches the column on which the
1066: cell containing its destination lies, it starts moving along vertical
1067: links and never goes back to horizontal ones.
1068: \end{itemize}
1069: \item The number of commodities on vertical links is at most the same
1070: as the number on links in the center cut -- but this requires proof.
1071: \end{itemize}
1072: To prove this last point, we need to count the number of commodities
1073: sharing a horizontal link crossing the center cut, and we have to give
1074: an upper bound on the number of commodities sharing an arbitrary vertical
1075: link.
1076:
1077: In terms of the number of commodities sharing a horizontal link across
1078: the center cut:
1079: \begin{itemize}
1080: \item The average number of commodities across the center cut is just
1081: \[
1082: n \cdot \mbox{$\frac 1 2$}\sqrt{\frac{c\log n}{n}}
1083: \;\; = \;\; \mbox{$\frac{\sqrt{c}}2$}\sqrt{n\log n},
1084: \]
1085: and again from the Chernoff bounds, we have that this is not only the
1086: ensemble average, but that in almost all networks, this number is
1087: $\Theta\big(\!\sqrt{n\log n}\big)$.
1088: \item By a similar argument, we have that the number of edges across
1089: the center cut in between two adjacent cells is $\Theta(\log^2 n)$ --
1090: with high probability, $\Theta(\log n)$ nodes in each cell, by construction
1091: there is a link between any two of those.
1092: \end{itemize}
1093: Thus, the number of commodities sharing a link across the center cut is
1094: $\frac{\Theta(\sqrt{n\log n})}{\Theta(\log^2 n)}$
1095: = $\Theta\left(\frac{\sqrt{n}}{\log^{\frac 3 2}(n)}\right)$.
1096:
1097: In terms of the number of commodities sharing any vertical link,
1098: an upper bound on this number is given by
1099: $\frac{\textrm{\tiny \# of nodes in a vertical strip}}
1100: {\textrm{\tiny \# of links between adjacent cells}}$.
1101: Why this is an upper bound is illustrated in Fig.~\ref{fig:vertical-links}.
1102: \begin{figure}[!ht]
1103: \centerline{\psfig{file=vertical-links.eps,height=7cm,width=5cm}}
1104: \caption{\small Upper bound on the number of commodities sharing a vertical
1105: link. Clearly, not all commodities that use a vertical link share a link
1106: across the thick horizontal line: some will switch from horizontal to
1107: vertical above the line and reach their destination before crossing that
1108: line, and the same will happen below. By estimating the number of
1109: commodities sharing a vertical link by the number of nodes in a vertical
1110: strip, we effectively say that all commodities in that strip use {\em all}
1111: vertical edges. This is certainly an overestimate, based on which we
1112: obtain only an upper bound.}
1113: \label{fig:vertical-links}
1114: \end{figure}
1115: But then, since the number of nodes in a vertical strip is twice the
1116: number of nodes in $\frac 1 2$ of a horizontal strip, from an argument
1117: entirely analogous to the count of commodities sharing a link across
1118: a center cut in the paragraph above, we have that the number of commodities
1119: sharing a vertical link is {\em at most}
1120: $\Theta\left(\frac{\sqrt{n}}{\log^{\frac 3 2}(n)}\right)$.
1121:
1122: In summary, we have that the link sharing the largest number of
1123: commodities is shared by
1124: $\Theta\left(\frac{\sqrt{n}}{\log^{\frac 3 2}(n)}\right)$ of them.
1125: Therefore, by the fairness constraint, the capacity of this link
1126: is shared equally among all commodities, and thus this capacity
1127: is the sought $\gamma_n$ value, i.e.,
1128: $\gamma_n = \Theta\left(\frac{\log^{\frac 3 2}(n)}{\sqrt{n}}\right)$ is
1129: achievable for the linear program in Table~\ref{tab:demand-zero}.
1130:
1131: \subsection{Remark}
1132:
1133: Note: the ``spirit'' of this proof is very similar to the proof
1134: in~\cite[Sec.\ IV]{GuptaK:00}: in both cases, the goal is to give
1135: an explicit construction to show the achievability of certain
1136: throughput values. However, the methods employed to analyze the
1137: throughput achieved by the routing strategies proposed differ
1138: significantly --~\cite{GuptaK:00} relies heavily on VC
1139: theory~\cite{VapnikC:71}, whereas we only use properties of flows
1140: and the Chernoff bound.
1141:
1142:
1143: \section{Applications to Wireless Networking Problems I: the Gupta-Kumar Setup}
1144: \label{sec:omnidirectional}
1145:
1146: Before considering more general node architectures in
1147: Section~\ref{sec:directional}, we show in this section how, for the case
1148: of nodes equipped with omnidirectional antennas, using our proof techniques
1149: we obtain scaling laws identical to those reported in~\cite{GuptaK:00},
1150: but under strong convergence.
1151:
1152: \subsection{Transmitter/Receiver Model}
1153:
1154: In~\cite{GuptaK:00}, transmissions were omnidirectional, and described
1155: based on a pure collision model: for a transmission to be successfully
1156: decoded, no other transmission has to be in progress within the range
1157: of the receiver under consideration. This setup is illustrated in
1158: Fig.~\ref{fig:txrx-model3}.
1159:
1160: \begin{figure}[ht]
1161: \centerline{\psfig{file=omnidirectional.eps,width=12cm,height=2.5cm}}
1162: \caption{\small A transmission model based on omnidirectional antennas
1163: and pure collisions.}
1164: \label{fig:txrx-model3}
1165: \end{figure}
1166:
1167: \subsection{Average Number of Edges Across the Cut}
1168:
1169: Our first task is to determine the {\em average} number of edges that
1170: can be simultaneously supported across the cut, average taken over all
1171: possible network realizations.
1172:
1173: \subsubsection{An Upper Bound}
1174:
1175: For a fixed receiver location $(x,y)$ in $R$, there can only be one
1176: active transmitter within distance $d_n$ of the receiver, for that
1177: transmission to be successfully received. Since to obtain an upper
1178: bound we only need worry about edges that cross the cut, we first
1179: consider all possible locations of one such transmitter in $L$, by
1180: drawing a circle of radius $d_n$ and center $(x,y)$. This region
1181: is illustrated in Fig.~\ref{fig:reverse-count}.
1182:
1183: \begin{figure}[ht]
1184: \centerline{\psfig{file=count2.eps,width=12cm,height=2.5cm}}
1185: \caption{\small For a receiver at location $(x,y)$, at most one transmitter
1186: in the shaded region $T_{xy}$ can send a message (if this message is to be
1187: successfully decoded on the other side of the cut).}
1188: \label{fig:reverse-count}
1189: \end{figure}
1190:
1191: Denoting by $|T_{xy}|$ the area of the shaded region $T_{xy}$ in
1192: Fig.~\ref{fig:reverse-count}, we use eqn.~(\ref{eq:num-nodes}) to
1193: estimate the number of transmitters located in $T_{xy}$ as $n|T_{xy}|$.
1194: However, since only one transmitter located within $T_{xy}$ can transmit
1195: successfully to a receiver at $(x,y)$, the number of nodes that are
1196: able to transmit at the same time from $L$ to $R$ is upper bounded by
1197: \[
1198: \frac{E(\textsl{Number of nodes in $L$})}
1199: {E(\textsl{Number of nodes in $T_{xy}$})}
1200: = \frac{n L}{n T_{xy}}
1201: \]
1202: This is an {\em upper bound}, because we are assuming that it is possible
1203: to find a set of locations $(x,y)$ in $R$ such that no area in $L$ is
1204: wasted---showing that this bound is indeed tight requires proof.
1205:
1206: Now, the area of $L$ is $d_n$. To compute the area of $T_{xy}$, we
1207: have to determine the area of an arc of a circle with angle $\vartheta$, as
1208: shown in Fig.~\ref{fig:reverse-count}, and in a computation entirely
1209: analogous to that of the calculation of $|Q_p|$ in
1210: Section~\ref{sec:multiple-directed-beams}. In this case, we have that
1211: $\sin(\frac 1 2(\pi-\vartheta))=\frac{x-\frac 1 2}{d_n}=\cos(\frac 1 2\vartheta)$,
1212: and since $\frac{1}{2} \leq x < \frac{1}{2}+d_{n}$ it is clear that
1213: we must have $0 < \vartheta \leq \pi$ and also $\sin{\vartheta} \geq 0$. Then,
1214: we get $|T_{xy}|
1215: = \mbox{\small$\frac 1 2$} \vartheta d_n^2
1216: - \mbox{\small$\frac 1 2$} d_n \cos(\mbox{\small$\frac 1 2$}\vartheta)
1217: 2 d_n \sin(\mbox{\small$\frac 1 2$}\vartheta)
1218: = \frac{1}{2} \vartheta d_{n}^2 - \frac{1}{2} d_{n}^2 \sin{\vartheta}$,
1219: and therefore, $|T_{xy}|
1220: = \mbox{\small$\frac 1 2$} d_n^2 (\vartheta - \sin{\vartheta})$. Hence,
1221: for each possible value of $\vartheta$, an upper bound on the number of
1222: nodes that are able to transmit at the same time from $L$ to $R$ is
1223: \[
1224: \frac{nL}{nT_{xy}}=\frac{nd_n}{n\frac 1 2d_n^2(\vartheta-\sin\vartheta)}
1225: =\frac 2{d_n(\vartheta-\sin\vartheta)}.
1226: \]
1227: Since this upper bound depends on the choice of receiver location
1228: (through the angle $\vartheta$), we will make this bound as small as possible
1229: by an appropriate choice of $\vartheta$. As noted above, $0<\vartheta\leq\pi$,
1230: and $\sin\vartheta\geq 0$. Hence, the number of transmitters in $L$ is
1231: smallest when $\vartheta=\pi$ and $\sin\vartheta=0$, i.e., when the receivers
1232: are located close to the cut boundary (as it should be, since it is in this
1233: case when receivers ``consume'' the maximum amount of transmitter area).
1234: In this case, we get
1235: \[ \min_{0<\vartheta\leq\pi}\left[\frac{2}{d_n(\vartheta-\sin\vartheta)}\right]
1236: = \frac 2{\pi d_n}
1237: \]
1238: as an upper bound on the number of edges across the cut. Furthermore,
1239: in this case we see immediately that to maximize capacity we must keep
1240: $d_n$ as small as possible---and we know from eqn.~(\ref{eq:min-conn-radius})
1241: that the smallest possible $d_n$ that will still maintain the network
1242: connected is $\Theta(\sqrt{\log n/n})$. Therefore, replacing for the
1243: optimal $d_n$, we finally get an upper bound of
1244: $\Theta\left(\sqrt{n/\log n}\right)$.
1245:
1246: \subsubsection{The Upper Bound is Asymptotically Tight}
1247:
1248: To verify that the upper bound is tight, we give an explicit flow
1249: construction. Consider the placement of disks shown in
1250: Fig.~\ref{fig:explicit-flow-construction}.
1251:
1252: \begin{figure}[ht]
1253: \centerline{\psfig{file=flow-construction.eps,width=12cm,height=2.5cm}}
1254: \caption{\small An explicit flow construction.}
1255: \label{fig:explicit-flow-construction}
1256: \end{figure}
1257:
1258: Since the height of the square is $1$, and we are placing nodes at
1259: distance $2d_n$ from each other, this guarantees that {\em if there
1260: are nodes in each of the circles to create valid tx/rx pairs}, then
1261: the number of successful simultaneous transmissions across the cut
1262: is $\frac 1{2d_n}=\Theta\left(\sqrt{n/\log n}\right)$. Whether all such pairs
1263: of nodes can be created simultaneously or not is the issue addressed
1264: next.
1265:
1266: \subsection{Uniform Convergence Issues}
1267:
1268: Next we prove that when $n$ points are dropped uniformly over the
1269: square $[0,1] \times [0,1]$, we have that simultaneously (i.e.,
1270: uniformly) over all $\frac 1{2d_n}$ circles from
1271: Fig.~\ref{fig:explicit-flow-construction}, each one of the circles
1272: contains $\Theta(\log(n))$ points in almost all network realizations.
1273: From this, we conclude that the distribution of the number of edges
1274: across the cut is sharply concentrated around its mean, and hence
1275: that in a randomly chosen network, with probability approaching 1 as
1276: $n\rightarrow\infty$, the actual number of straddling edges is indeed
1277: $\Theta\left(\sqrt{n/\log(n)}\right)$.
1278:
1279: \subsubsection{Statement of the Result}
1280:
1281: Consider we have $\frac{1}{2d_{n}}$ circles centered along the
1282: $x=\frac 1 2$ cut as shown in Fig.~\ref{fig:explicit-flow-construction},
1283: with centers $y_{j}=(2j-1)d_{n}$, $j=1 \ldots \frac{1}{2d_{n}}$ and radius
1284: $d_{n}$. Then, we have the following uniform convergence result:
1285:
1286: \begin{proposition}
1287: Define $B_j := [|N_{j}- \pi \log n| < \delta \pi \log n]$.
1288: Then, as $n \rightarrow \infty$, and for any $\delta \in (x,1)$
1289: ($x \approx 0.6$), we have that
1290: \[ \lim_{n\rightarrow\infty}
1291: P\left[\bigcap_{j=1}^{\sqrt{\frac{n}{\log n}}} B_j\right] = 1.
1292: \]
1293: \label{prop:uniformity}
1294: \end{proposition}
1295: Essentially what this proposition says is that with very high probability
1296: and uniformly over $j$, all $A_j$'s contain $\Theta(\log n)$ nodes.
1297:
1298: \subsubsection{Proof}
1299:
1300: Note that the area of a circle in Fig.~\ref{fig:explicit-flow-construction}
1301: is $\pi d_n^2 = \pi \frac{\log n}{n}$. Then, from the Chernoff bound,
1302: we have that for any $0<\delta<1$ we can find a $\theta > 0$ such that
1303: \begin{equation}
1304: P[|N_j-\pi \log n| > \delta \pi \log n]
1305: < e^{-\theta \pi \log n} = n^{-\theta\pi}.
1306: \label{eq:ch1}
1307: \end{equation}
1308: Thus, we can conclude that the probability that the number of nodes in
1309: a circle deviates by more than a constant factor from the mean tends to
1310: zero as $n\rightarrow\infty$. This is a key step in showing that all
1311: the events $B_j := \big[|N_j-\pi\log(n)| < \delta\pi\log(n)\big]$ occur
1312: {\em simultaneously}. Now, from the union bound, we have that
1313: \[
1314: P \left[ \bigcap_{j=1}^{\frac{1}{2d_{n}}} B_{j} \right]
1315: \;\; = \;\; 1 - P \left [\bigcup_{j=1}^{\frac{1}{2d_{n}}} B_{j}^{c} \right]
1316: \;\; \geq \;\; 1 - \sum_{j=1}^{\frac{1}{2d_{n}}} P[B_{j}^{c}].
1317: \]
1318: But, from eqn.~(\ref{eq:ch1}), $P[B_j^c] < n^{-\theta\pi}$, and therefore,
1319: \[
1320: \sum_{j=1}^{\frac 1{2d_n}} P[B_j^c]
1321: \;\; < \;\; \sum_{j=1}^{\frac{1}{2d_{n}}} n^{-\theta\pi}
1322: \;\; = \;\; \frac{n^{-\theta\pi}}{2d_n}
1323: \;\; = \;\; \frac{n^{\frac{1}{2} - \pi \theta}}{2 \sqrt{\log n}}
1324: \]
1325: Putting everything together, and letting $n\rightarrow\infty$, we have
1326: \[
1327: P \left[\bigcap_{j=1}^{\frac{1}{2d_n}}B_j\!\right]
1328: \;\;\geq\;\; 1\!-\!\frac{n^{\frac{1}{2} - \pi \theta}}{2 \sqrt{\log n}}
1329: \;\;\longrightarrow\;\; 1,
1330: \]
1331: if and only if $\pi \theta > \frac{1}{2}$. And this is true for
1332: $\delta \approx 0.6$ and above (this follows from the definition of $\theta$
1333: and a simple numerical evaluation).
1334:
1335:
1336: \section{Applications to Wireless Networking Problems II: Directional Antennas}
1337: \label{sec:directional}
1338:
1339: \subsection{On Directional Antennas and MST Issues}
1340:
1341: We consider now an application of the techniques that were used
1342: so far to analyze the network capacity problem in the context of
1343: directional antennas.
1344:
1345: Why the interest in directional antennas? Because there is a
1346: question about wireless networks equipped with such antennas which
1347: we believe is very important, and for which we could not find a
1348: satisfactory answer in the literature. We discussed in
1349: Section~\ref{sec:omnidirectional} the vanishing throughput problem
1350: identified in~\cite{GuptaK:00}. But in a different segment of the
1351: research community, the use of {\em directional} antennas has also
1352: received a fair amount of attention in recent times. The rationale
1353: is that with omnidirectional antennas, existing MAC protocols require
1354: all nodes in the vicinity of a transmission to remain silent. With
1355: directional antennas however, it should be possible to achieve higher
1356: overall throughput, by means of a higher degree of spatial reuse of
1357: the shared medium, and a smaller number of hops visited by a packet
1358: on its way to destination (see, e.g.,~\cite{ChoudhuryYRV:02}).
1359: Furthermore, in the context of energy-efficient broadcast/multicast,
1360: it has been argued that the ability of a transmitter to reach multiple
1361: receivers is an important source of gains to take advantage of in
1362: the development of suitable protocols, such as BIP~\cite{WieselthierNE:02}.
1363:
1364: If we take a step back, careful reading of these previous results
1365: raises an important question: how much exactly is there to gain from the
1366: use of directional antennas? Could directional antennas (in which the
1367: width of the beams tends to zero as $n$ gets large) be used to effectively
1368: overcome the vanishing maximum throughput of~\cite{GuptaK:00}? Although
1369: we have not been able to find answers to this question in the literature
1370: (and that motivated us to start working on this problem in the first
1371: place), we have found a couple of related results based on which we can
1372: say a-priori that the answer is probably {\em no}:
1373: \begin{itemize}
1374: \item In~\cite{GuptaK:00}, the authors claim that their result holds
1375: irrespective of whether transmissions are omnidirectional or directed,
1376: provided that in the case of directed antennas there is some lower
1377: bound (independent of network size) on how narrow the beams can be made.
1378: \item In~\cite{MergenT:02,TongZM:01}, for some {\em regular} networks,
1379: it is shown that enabling nodes with Multi-Packet Reception (MPR)
1380: capabilities~\cite{GhezVS:89} can only increase the total
1381: throughput of the network by a constant factor ($\approx 1.6$),
1382: independent of network size.
1383: \end{itemize}
1384: Given this state of affairs, it seems to us that deciding exactly how
1385: much there is to be gained by using directional antennas, and giving
1386: some measure of how complex the transmitters/receivers need to be made
1387: to achieve those gains, is indeed a topic worth being studied.
1388:
1389: \subsection{A Single Directed Beam}
1390: \label{sec:single-directed-beam}
1391:
1392: \subsubsection{Transmitter/Receiver Model}
1393:
1394: In this section we consider the first model based on directional
1395: antennas: transmitters can generate a beam of arbitrarily narrow width
1396: aimed at any particular receiver, and receivers can accept any number
1397: of incoming messages, provided the transmitters are not in the same
1398: straight line. This results in a significant increase in the complexity
1399: of the signal processing algorithms required at each node, and in this
1400: section our goal is to determine if and how much it is possible to
1401: increase the achievable MST, compared to the omnidirectional case.
1402: This model is illustrated in Fig.~\ref{fig:txrx-model2}.
1403:
1404: \begin{figure}[ht]
1405: \centerline{\psfig{file=single-beam.eps,width=12cm,height=2.5cm}}
1406: \caption{\small A single beam model for communication between nodes.}
1407: \label{fig:txrx-model2}
1408: \end{figure}
1409:
1410: Our goal in this subsection is to evaluate $\Theta(\nu_n^*)$ and
1411: $\Theta(\gamma_n)$, for this particular architecture.
1412:
1413: \subsubsection{Average Number of Edges Across the Cut}
1414:
1415: Since at most one edge per transmitter can be active at any point
1416: in time, the average number of edges going across the cut can be no
1417: larger than $n d_n$, the average number of transmitters on its left
1418: side. Since $L$ and $R$ have the same area, the average number of
1419: nodes on each side of the cut is the same (and equal to $nd_n$), and
1420: hence the maximum of $nd_n$ transmissions can actually be received,
1421: by ``pairing up'' every node from one side of the cut with every
1422: node on the other side. The pairing of nodes on each side of the
1423: cut is illustrated in Fig.~\ref{fig:cut-singlebeam}.
1424:
1425: \begin{figure}[ht]
1426: \centerline{\psfig{file=cut-single-beam.eps,width=12cm,height=2.5cm}}
1427: \caption{\small Pairing up one transmitter in $L$ with one receiver in
1428: $R$: at most $n|L| = n|R| = nd_n$ such pairs can be formed.}
1429: \label{fig:cut-singlebeam}
1430: \end{figure}
1431:
1432: Finally we note that, under the assumption of arbitrarily narrow and
1433: perfectly aligned beams, the only way in which we could have multiple
1434: receivers blocked out by a single transmission is by having them all
1435: lying in a nearly straight line (i.e., a set of vanishing measure)
1436: under the beam of a single transmitter. But then, to have an actual
1437: edge count lower than $\Theta(nd_n)$, we would require an increasingly
1438: large number of nodes falling in a decreasingly small area: under our
1439: statistical model for node placement, this event occurs with vanishing
1440: probability, and therefore the average edge count is $\Theta(nd_n)$.
1441:
1442: \subsubsection{Sharp Concentration Results}
1443:
1444: \paragraph{Number of Transmitters in $L$ and Receivers in $R$}
1445: Again, consider $n$ points $X_1...X_n$ uniformly distributed over
1446: the $[0,1]\times [0,1]$ plane, and consider the area $L$ on the left
1447: side of the cut, as shown in Fig.~\ref{fig:cut-singlebeam}. We define
1448: variables
1449: \[ N_i = \left\{\begin{array}{rl}
1450: 1, & X_i \in L \\
1451: 0, & \textrm{otherwise.}
1452: \end{array}\right.
1453: \]
1454: and $N = \sum_{i=1}^n N_i$. The probability $p$ of $X_i\in L$ is
1455: $p = |L| = 1\cdot d_n$. Hence,
1456: $E(N_i) = 1\cdot p + 0\cdot(1-p) = p = d_n$,
1457: and $E(N) = \sum_{i=1}^n E(N_i) = nd_n$.
1458: From the Chernoff bound, we know that
1459: \[ P\left(|N-nd_n| > \delta nd_n\right) \;\; < \;\; e^{-\theta nd_n}.
1460: \]
1461: Since $\theta>0$, we have that as $n\rightarrow\infty$, deviations of
1462: $N$ from its mean by a constant fraction (independent of $n$) occur
1463: with low probability, provided $d_n$ does not decay too fast.\footnote{Note
1464: that the fastest possible decay for $d_n$, according to
1465: eq.~(\ref{eq:min-conn-radius}), is when $d_n \approx \sqrt{\frac{c\log n}n}$.
1466: And in this case, $e^{-\theta nd_n} = e^{-\theta\sqrt{cn\log n}}\to 0$
1467: as $n\to\infty$. If $d_n$ is any bigger, this probability goes to zero
1468: even faster. So the Chernoff bound applies for any connected network.}
1469: Therefore, we conclude that in almost all realizations of the network,
1470: the number of transmitters in $L$ and the number of receivers in $R$ is
1471: $\Theta(nd_n)$.
1472:
1473: \paragraph{Number of Edges Across the Cut}
1474: Knowing that we have $\Theta(nd_n)$ transmitters and receivers within
1475: range of each other on each side of the cut is not enough to claim that
1476: the number of edges that cross the cut is $\Theta(nd_n)$. This is because,
1477: in our model for directional antennas, a receiver can successfully decode
1478: two simultaneous incoming transmissions provided the angle formed by the
1479: receiver and the two transmitters is strictly positive: if all three are
1480: on the same straight line, collisions still occur, and those edges are
1481: destroyed. Therefore, we still need to show that the actual number of
1482: edges is $\Theta(nd_n)$. And to do this, we need to say something about
1483: the location of points that end up in $L$, and not just count how many.
1484: To proceed, we cut the area of $L$ into $nd_n$ rectangles of height
1485: $\frac 1{nd_n}$ and width $d_n$, as illustrated in Fig.~\ref{fig:cutting-L}.
1486: Our goal then becomes to show that in ``most'' of these rectangles (meaning,
1487: in all but a constant fraction of them) we will have nodes capable of
1488: forming straddling edges.
1489: \begin{figure}[ht]
1490: \centerline{\psfig{file=cutting-L.eps,width=12cm,height=2.5cm}}
1491: \caption{\small Cutting $L$ and $R$ into rectangles of size $d_n \times
1492: \frac 1{nd_n}$.}
1493: \label{fig:cutting-L}
1494: \end{figure}
1495:
1496: Counting how many of the $nd_n$ rectangles in Fig.~\ref{fig:cutting-L}
1497: contain at least one of the $\Theta(nd_n)$ nodes that are dropped in $L$
1498: is an instance of a classical {\em occupancy} problem, in which $k$ balls
1499: are thrown uniformly onto $m$ bins, in the case where
1500: $k=m=nd_n$~\cite[Ch.\ 4]{MotwaniR:95}. Since $\frac 1 m$ is the
1501: probability that a ball falls in any particular bin, the probability $p$
1502: of an empty bin after throwing all $m$ balls is $p=(1-\frac 1 m)^m$ which,
1503: for $m$ large, becomes approximately $\frac 1 e$. Therefore, the average
1504: number of empty bins is $mp \approx \frac 1 e \sqrt{n\log(n)}$. And by
1505: the Chernoff bound, again we have that
1506: \[ P\left(Y- nd_n/e > \delta nd_n/e\right)
1507: < e^{-\theta nd_n/e},
1508: \]
1509: where $Y$ is the number of empty bins. So, the probability that the
1510: number of empty bins is a constant factor away from its mean is small
1511: (again, provided $d_n$ does not decay too fast), and hence, for $n$
1512: large, almost all network realizations will have $\Theta(nd_n)$ non-empty
1513: rectangles. But since transmitter/receiver pairs in different rectangles
1514: are not collinear, the number of edges across the cut is $\Theta(nd_n)$,
1515: qed.
1516:
1517: \subsubsection{Remarks}
1518:
1519: \paragraph{MST in a Minimally Connected Network}
1520: In this section, we found that the MST achievable by the type of
1521: tx/rx pairs considered here depends on the connectivity radius $d_n$.
1522: If we replace $d_n$ with $\sqrt{\frac{c\log n}{n}}$ (the minimum radius
1523: required to maintain a connected network, from~\cite{GuptaK:98}), we
1524: get
1525: \[ n d_n \approx n \sqrt{\frac{c\log n}{n}}
1526: = \Theta\left(\!\sqrt{n \log n}\right).
1527: \]
1528: Comparing this expression with its equivalent from
1529: Section~\ref{sec:omnidirectional}, we see that all we gain over the
1530: case of omnidirectional antennas is an increase in MST by a
1531: factor of $\Theta(\log n)$.
1532:
1533: \paragraph{Minimum Connectivity Radius Resuting in MST = $\Theta(n)$}
1534: In this tx/rx architecture we are considering the use of arbitrarily
1535: narrow and perfectly aligned directed beams. Therefore, it does make
1536: sense to consider the use of a possibly larger transmission range than
1537: the minimum required to keep the network connected, since in this case
1538: a large range does not force other tx/rx pairs to remain
1539: silent while a given transmission is in progress. And since by increasing
1540: the transmission range now we can increase throughput, our next goal
1541: is to determine the minimum range that would be required to achieve
1542: MST = $\Theta(n)$.
1543:
1544: Solving for $d_n$ in $\Theta(n) = \Theta(nd_n)$, we see that
1545: trivially, $d_n = \Theta(1)$. That is, to achieve MST linear in the
1546: number of nodes using a single beam in each transmission, the radius
1547: of each transmission has to be a constant independent of $n$.
1548:
1549: \paragraph{Minimum Number of Simultaneous Beams}
1550: From a practical point of view, does it matter that to achieve linear
1551: MST we need to keep the transmission radius constant? In this section
1552: we argue that yes it does, very much. To see why this is so, next we
1553: count the minimum number $\beta$ of narrow beams that a transmitter
1554: would have to generate simultaneously, if MST linear in the size of
1555: the network is to be achieved: this number gives a measure of the
1556: {\em complexity} of the beamforming transmitter, since $2\pi/\beta$ is
1557: an upper bound on the maximum angle of dispersion of the beam.
1558:
1559: Since a node can generate a beam to any receiver within its transmission
1560: range (see Fig.~\ref{fig:txrx-model2}), again using eqns.~(\ref{eq:num-nodes})
1561: and~(\ref{eq:chernoff}), we have that for $n$ large, the number of points
1562: within a circle of radius $d_n$ is $\Theta(n\cdot\pi d_n^2)$. In the case
1563: of $d_n$ only satisfying the requirement of keeping the network connected,
1564: \[
1565: \beta \;\; = \;\; n\cdot\pi d_n^2
1566: \;\; = \;\; n\left(\frac{\pi c\log n}n\right)
1567: \;\; = \;\; \Theta(\log n).
1568: \]
1569: This fact was known already---see~\cite{XueK:02} for a more complete
1570: analysis (constants hidden by the $\Theta$-notation included), including
1571: also a number of interesting references on the history of this problem.
1572: But if now we consider a larger $d_n$ satisfying the requirement of
1573: achieving linear MST, then
1574: \[
1575: \beta \;\; = \;\; n\cdot\Theta(1)^2 \;\; = \;\; \Theta(n).\]
1576: Therefore, we see $\beta$ has an {\em exponential} increase relative
1577: to the number required to maintain minimum connectivity---it is on this
1578: fact that we base our claim about directional antennas not being able
1579: to provide an effective means of overcoming the issue with per-node
1580: vanishing throughputs.
1581:
1582: \subsection{Multiple Directed Beams}
1583: \label{sec:multiple-directed-beams}
1584:
1585: \subsubsection{Transmitter/Receiver Model}
1586:
1587: In this section we consider another model based on directional
1588: antennas: transmitters can generate an arbitrary number of beams, of
1589: arbitrarily narrow width, aimed at any particular receiver; and receivers
1590: can accept any number of incoming messages, provided the transmitters are
1591: not in the same straight line. This is perhaps the most complex scheme
1592: that could be envisioned based on directed beams. Our goal is to determine
1593: if and how much it is possible to increase the achievable MST, compared
1594: to the previous two cases. This model is illustrated in
1595: Fig.~\ref{fig:txrx-model1}.
1596:
1597: \begin{figure}[ht]
1598: \centerline{\psfig{file=nondeg-bcast.eps,width=12cm,height=2.5cm}}
1599: \caption{\small A non-degraded broadcast channel model for communication
1600: between nodes: each node is able to send simultaneously a different packet
1601: to each one of the nodes within his transmission range. Furthermore,
1602: multiple broadcasts (from different transmitters) do not collide, unless
1603: the transmitters are perfectly aligned.}
1604: \label{fig:txrx-model1}
1605: \end{figure}
1606:
1607: Again, our goal in this subsection is to evaluate $\Theta(\nu_n^*)$ and
1608: $\Theta(\gamma_n)$, for this particular architecture.
1609:
1610: \subsubsection{Average Number of Edges Across the Cut}
1611:
1612: With minor variations, this calculation is essentially identical
1613: to that presented in Section~\ref{sec:upperb-average}, and the
1614: final result is the same: the ensemble average number of edges
1615: straddling the center cut is $\Theta(n^2d_n^3)$. See
1616: Section~\ref{sec:upperb-average} for details.
1617:
1618: \subsubsection{Sharp Concentration Results}
1619:
1620: Our next goal is to show that the actual number of edges straddling
1621: the cut in any realization of the network is sharply concentrated around
1622: its mean. That is, in almost all networks, the number of edges across
1623: the cut is $\Theta(n^2d_n^3)$,
1624:
1625: \paragraph{Number of Receivers per Transmitter}
1626: Define a binary random variable $N_{ij}$, which takes the value 1 if
1627: the $i$-th node is within the transmission range of a node at coordinates
1628: $(x_j,y_j)$ on the other side of the cut, as illustrated in
1629: Fig.~\ref{fig:multiple-beams-count}:
1630: \[ N_{ij} = \left\{ \begin{array}{rl}
1631: 1, & X_i \in Q_{(x_j,y_j)} \\
1632: 0, & \textrm{otherwise.}
1633: \end{array}
1634: \right.
1635: \]
1636: Let $p$ denote the probability that $X_i$ is in $Q_{(x_j,y_j)}$ (i.e.,
1637: that $N_{ij} = 1$). Then, $p = |Q_{(x_j,y_j)}| =
1638: \frac 1 2 d_n^2(\vartheta-\sin(\vartheta))$, with $0\leq\vartheta\leq\pi$
1639: is as in Fig.~\ref{fig:multiple-beams-count}. Therefore, defining
1640: $\kappa_\vartheta$ as $\frac 1 2 (\vartheta-\sin(\vartheta))$, we have
1641: $p = |Q_{(x_j,y_j)}| = \kappa_\vartheta d_n^2 =
1642: \kappa_\vartheta\frac{\log n}n$.
1643:
1644: Define $N_j = \sum_{i=1}^n N_{ij}$ as the number of points in
1645: $Q_{(x_j,y_j)}$. In this case, we have $E(N_j) = \sum_{i=1}^n N_{ij}
1646: = \sum_{i=1}^n p\cdot 1 + (1-p)\cdot 0 = np = \kappa_\vartheta\log(n)$.
1647: Now, again from the Chernoff bound, we have that
1648: \[
1649: P(|N_j-\kappa_\vartheta\log(n)| > \delta\kappa_\vartheta\log(n))
1650: < e^{-\theta\kappa_\vartheta\log(n)} = n^{-\theta\kappa_\vartheta},
1651: \]
1652: for $\theta$ defined as in previous applications. As $n\rightarrow\infty$
1653: this probability tends to zero, and therefore, in almost all network
1654: realizations, a transmitter on the left side of the cut will be able to
1655: reach $\Theta(\log(n))$ receivers on the right side.
1656:
1657: \paragraph{Total Number of Edges}
1658: In a manner analogous to the situation discussed in
1659: Section~\ref{sec:single-directed-beam}, knowing that there are
1660: $\Theta(nd_n)$ transmitters on the left side of the cut, and that each
1661: transmitter can reach $\Theta(nd_n^2)$ receivers on the other side, is
1662: not enough to conclude that the total number of edges going across the
1663: cut must be $\Theta(n^2d_n^3)$. This is because of our requirement that
1664: multiple transmitters not be perfectly aligned with a receiver for this
1665: receiver to decode all these messages simultaneously. Therefore, we
1666: still need to show that the actual number of edges is $\Theta(n^2d_n^3)$.
1667: And to do this, we need to say something about the location of points in
1668: $R$ that can be reached from $L$, and not just count how many. To
1669: proceed then, we cut the area of $Q_p$ into $\kappa_\vartheta\log(n)$
1670: slices, each slice of area $\frac{|Q_p|}{\kappa_\vartheta\log(n)}
1671: = \frac 1 n$, as illustrated in Fig.~\ref{fig:cutting-Qp}.
1672: \begin{figure}[ht]
1673: \centerline{\psfig{file=cutting-Qp.eps,width=12cm,height=2.5cm}}
1674: \vspace{-1mm}
1675: \caption{\small Cutting the shaded arc $Q_{xy}$ into regions of area
1676: $\frac 1 n$, to formulate this as an occupancy problem analogous to
1677: that of Fig.~\ref{fig:cutting-L}.}
1678: \label{fig:cutting-Qp}
1679: \end{figure}
1680:
1681: As in the occupancy problem considered in
1682: Section~\ref{sec:single-directed-beam}, our goal is to show that in
1683: ``most'' of these arc slices (most meaning, in all but a constant fraction
1684: of them) we will have nodes capable of forming straddling edges. This
1685: is again a problem of throwing $k$ balls uniformly into $m$ bins, where
1686: $k = m = \kappa_\vartheta\log(n)$. And again, we have that with probability
1687: that tends to 1 as $n\rightarrow\infty$, the number of empty bins is
1688: $\kappa_\vartheta\log(n)/e$, and hence the number of occupied bins is
1689: $\Theta(\log(n))$.
1690:
1691: Consider now a fixed transmitter located at some coordinates $(x,y)$.
1692: Any other transmitter located at coordinates $(x',y')\neq(x,y)$ defines
1693: a unique straight line that goes through $(x,y)$ and $(x',y')$. If
1694: there is a receiver on the other side of the cut along this line, within
1695: reach of both transmitters, then those two edges will be lost---and those
1696: will be the {\em only} lost edges, from among the $\kappa_\vartheta\log(n)$
1697: that each transmitter has. This situation is illustrated in
1698: Fig.~\ref{fig:manybeams-lost-edges}.
1699: \begin{figure}[ht]
1700: \centerline{\psfig{file=interference.eps,width=12cm,height=2.5cm}}
1701: \vspace{-1mm}
1702: \caption{\small To illustrate how we could end up losing edges: if the
1703: two black transmitters attempt simultaneously to communicate with the
1704: gray receiver, a collision will occur, and none of the edges will be
1705: created.}
1706: \label{fig:manybeams-lost-edges}
1707: \end{figure}
1708:
1709: And then we are done. We have established that in almost all network
1710: realizations, there are $\Theta(nd_n)$ transmitters within each side of
1711: the cut, that each transmitter can reach $\Theta(d_n^2)$
1712: receivers on the other side of the cut, and that integrating out
1713: $\kappa_\vartheta$ we obtain exactly $\Theta(n^2d_n^3)$ edges going
1714: across the cut. Therefore, the actual number of edges across the cut
1715: is sharply concentrated around its mean, qed.
1716:
1717: \subsubsection{Remarks}
1718:
1719: \paragraph{MST in a Minimally Connected Network}
1720: Substituting for $d_n = \sqrt{\frac{c\log n}{n}}$ in $\frac 2 3n^2d_n^3$,
1721: we get
1722: \[
1723: \mbox{\small $\frac 2 3$}n^2\left(\frac{\log n}{\pi n}\right)^\frac{3}{2}
1724: = \mbox{\small $\frac 2 3$}\sqrt{n} \log^\frac{3}{2}n
1725: = \Theta\left(\sqrt{n}\log^\frac{3}{2}(n)\right)
1726: \]
1727:
1728: Comparing this expression to the ones obtained in
1729: Sections~\ref{sec:omnidirectional} and~\ref{sec:single-directed-beam},
1730: we see that the MST gain due to the use of multiple simultaneous,
1731: arbitrarily narrow beam is, at most, $\Theta\left(\log^2(n)\right)$.
1732:
1733: \paragraph{Minimum Connectivity Radius Resuting in MST = $\Theta(n)$}
1734: The minimum $d_n$ resulting in linear MST is obtained by
1735: solving for $d_n$ in $\Theta(n^2 d_n^3) = \Theta(n)$. Now, for $n$ large
1736: enough, there exist constants $c_1<c_2\in\mathbb{R}$
1737: ($c_1>0$ and $c_2<\infty$), such that $c_1 n < \frac 2 3n^2 d_n^3 < c_2 n$,
1738: or equivalently,
1739: $c_1\frac 3 2 n^{-\frac 1 3} < d_n < c_2\frac 3 2 n^{-\frac 1 3}$.
1740: Therefore,
1741: \[ d_n = \Theta(n^{-\frac{1}{3}}). \]
1742:
1743: \paragraph{Minimum Number of Simultaneous Beams}
1744: In Section~\ref{sec:single-directed-beam}, we said that keeping the
1745: transmission range constant resulted in an impractically large number
1746: of beams that the receiver needed to generate, if linear MST was to
1747: be achieved by increasing the complexity of the signal processing
1748: algorithms. But if we generate multiple beams, we have just shown
1749: that this minimum radius now is no longer a constant, but instead
1750: tends to zero as $ \Theta(n^{-\frac{1}{3}})$. However, the situation
1751: is not much better compared to the single beam case, and to see this
1752: again we compute the minimum number of simultaneous beams that a
1753: transmitter would have to generate.
1754:
1755: If now we consider the larger $d_n$ satisfying the requirement of
1756: achieving maximum stable throughput linear in network size, then
1757: \[
1758: \beta = n\cdot\pi d_n^2 = n\Theta(n^{-\frac 2 3}) = \Theta(n^{\frac 1 3}).
1759: \]
1760: Therefore, we see that while $\beta$ is smaller than in the case of the
1761: single beam, we still have an {\em exponential} increase relative
1762: to the number required to maintain minimum connectivity---so again,
1763: we claim that directional antennas are not able to provide an effective
1764: means of overcoming the issue with per-node vanishing throughputs.
1765:
1766:
1767: \section{Conclusions}
1768: \label{sec:conclusions}
1769:
1770: \subsection{Summary of Contributions}
1771:
1772: In this paper, we have showed how network flow methods can be used
1773: to determine (to within constants) the maximum stable throughput
1774: achievable in a wireless network. This was done by formulating MST
1775: as a maximum multicommodity flow problem, for which tight upper and
1776: lower bounds were found. In the process, the difficult multicommodity
1777: problem was proved equivalent to a simpler single commodity problem,
1778: solvable using standard arguments based on flows and cuts.
1779:
1780: As mentioned in the Introduction, this work grows out of our desire
1781: to cast what we deem to be the most useful insights in~\cite{GuptaK:00}
1782: (basically, that the constriction in capacity results from the need
1783: to share constant capacity links by a growing number of nodes), in
1784: a form that makes more intuitive sense to us.\footnote{However, we
1785: certainly would have not been able to obtain our simpler and more
1786: general proofs without the insights provided by cultivating an
1787: appreciation for the line of reasoning employed by Gupta and Kumar
1788: in~\cite{GuptaK:00}.} And we feel we have accomplished that:
1789: \begin{itemize}
1790: \item By reducing the problem to counting the average number of
1791: edges that cross a cut and then proving a sharp concentration
1792: result around this mean, the computational task becomes very
1793: simple, involving only elementary tools from combinatorics
1794: and discrete probability. In~\cite{GuptaK:00}, similar results
1795: had been obtained based essentially on generalizations of the
1796: Glivenko-Cantelli theorem (that add uniformity to convergence
1797: in the law of large numbers), due to Vapnik and
1798: Chervonenkis~\cite[Ch.\ 2]{Pollard:84}, \cite{VapnikC:71}.
1799: \item In our formulation, it is straightforward to see that
1800: capacity limitations arise essentially from the geometry of
1801: the problem: edges have a constant capacity, and only about
1802: $\sqrt{n}$ of them are available at a minimum cut to transport
1803: the traffic generated by $n$ sources.
1804: \end{itemize}
1805:
1806: %\subsection{Towards a Pure Information-Theoretic Treatment of the
1807: % Capacity Problem for Dense Networks}
1808: %
1809: %A brief moment of thought makes it clear that ``the protocol and
1810: %physical layer models inspired by current technology'' described
1811: %in~\cite{GuptaK:00}, essentially boil down to the definition of
1812: %a network of point-to-point {\em independent} channels. But these
1813: %networks have received a fair amount of attention in the literature:
1814: %\begin{itemize}
1815: %\item Communication over networks of point-to-point links is the
1816: % main problem studied in {\em network coding}~\cite{AhlswedeCLY:00}.
1817: %\item Under a different data and traffic model (transmission of
1818: % correlated sources to a single sink), such networks have been
1819: % studied extensively from a pure information-theoretic perspective
1820: % in~\cite{BarrosS:03e}.
1821: %\end{itemize}
1822: %The two references above (\cite{AhlswedeCLY:00} and~\cite{BarrosS:03e})
1823: %identify specific scenarios in which Shannon information has all the
1824: %exact same properties of a network flow: multicast sessions in the
1825: %case of~\cite{AhlswedeCLY:00}, incast of correlated information in the
1826: %case of~\cite{BarrosS:03e} (even though in the first case network coding
1827: %is required to achieve the flow bounds, but not in the second). In this
1828: %paper, we have studied a network of point-to-point links with a traffic
1829: %pattern that does not fit either one of these previous patterns, but
1830: %
1831: % In particular, the problem of multiple unicast sessions considered
1832: % in this work, from a pure network coding perspective, has been
1833: % considered recently in~\cite{LiL:04b}.
1834:
1835: \subsection{Future Work}
1836:
1837: In terms of future work, there are a number of interesting
1838: questions opened up by this work. One deals with the generalization
1839: of these results to nodes distributed on arbitrary manifolds (instead
1840: of the square $[0,1]\times[0,1]\subset\mathbb{R}^2$). Another
1841: deals with exploring other combinatorial structures (such as
1842: hypergraphs~\cite{Berge:87}), to develop better collision models,
1843: especially in the omnidirectional case. Of particular interest
1844: to us however is the development of a purely information-theoretic
1845: formulation for the results in this paper, by exploiting the
1846: connections between Shannon information and network flow theory
1847: discovered in~\cite{BarrosS:03e}.
1848:
1849: \section*{Acknowledgements}
1850:
1851: Will appear in the final version.
1852: % The authors would like to thank T.\ L.\ Fine, for pointing them to
1853: % the text of Pollard~\cite{Pollard:84}, a reference that proved to be
1854: % extremely useful in the process of developing some basic understanding
1855: % on what VC theory is about~\cite{VapnikC:71}. They would also like
1856: % to thank two of the anonymous reviewers of their ACM MobiHoc 2003
1857: % paper~\cite{PerakiS:03}, for their unusually useful feedback.
1858: %%%% Also: \'Eva Tardos, Zhi-Hern Loh.
1859:
1860:
1861: %\bibliographystyle{plain}
1862: %\bibliography{library}
1863: \begin{thebibliography}{10}
1864:
1865: \bibitem{AhlswedeCLY:00}
1866: R.~Ahlswede, N.~Cai, S.-Y.~R. Li, and R.~W. Yeung.
1867: \newblock {Network Information Flow}.
1868: \newblock {\em IEEE Trans. Inform. Theory}, 46(4):1204--1216, 2000.
1869:
1870: \bibitem{BarrosS:03e}
1871: J.~Barros and S.~D. Servetto.
1872: \newblock {Network Information Flow with Correlated Sources}.
1873: \newblock Submitted to the IEEE Transactions on Information Theory, November
1874: 2003. Revised, January 2005 (Original title: {\em The Sensor Reachback
1875: Problem}.) Available from from \href{http://cn.ece.cornell.edu/}{{\tt
1876: http://cn.ece.cornell.edu/}}.
1877:
1878: \bibitem{Berge:87}
1879: C.~Berge.
1880: \newblock {\em Hypergraphes (Combinatoires des ensembles finis)}.
1881: \newblock Gauthier-Villars, 1987.
1882:
1883: \bibitem{ChoudhuryYRV:02}
1884: R.~R. Choudhury, X.~Yang, R.~Ramanathan, and N.~Vaidya.
1885: \newblock {Medium Access Control in Ad Hoc Networks Using Directional
1886: Antennas}.
1887: \newblock In {\em Proc. ACM MobiCom}, Atlanta, GA, 2002.
1888:
1889: \bibitem{CormenLRS:01}
1890: T.~H. Cormen, C.~E. Leiserson, R.~L. Rivest, and C.~Stein.
1891: \newblock {\em {Introduction to Algorithms (2nd ed)}}.
1892: \newblock MIT Press, 2001.
1893:
1894: \bibitem{cover-thomas:it-book}
1895: T.~M. Cover and J.~Thomas.
1896: \newblock {\em {Elements of Information Theory}}.
1897: \newblock John Wiley and Sons, Inc., 1991.
1898:
1899: \bibitem{FordF:62}
1900: L.~R. Ford and D.~R. Fulkerson.
1901: \newblock {\em {Flows in Networks}}.
1902: \newblock Princeton University Press, 1962.
1903:
1904: \bibitem{GastparV:05}
1905: M.~Gastpar and M.~Vetterli.
1906: \newblock {On the Capacity of Large Gaussian Relay Networks}.
1907: \newblock {\em IEEE Trans. Inform. Theory}, 51(3):765--779, 2005.
1908:
1909: \bibitem{GhezVS:89}
1910: S.~Ghez, S.~Verd\'u, and S.~Schwartz.
1911: \newblock {Optimal Decentralized Control in the Random Access Multipacket
1912: Channel}.
1913: \newblock {\em IEEE Trans. Autom. Control}, 34(11):1153--1163, 1989.
1914:
1915: \bibitem{GrahamKP:94}
1916: R.~L. Graham, D.~E. Knuth, and O.~Patashnik.
1917: \newblock {\em {Concrete Mathematics---A Foundation for Computer Science}}.
1918: \newblock Addison Wesley, 1994.
1919:
1920: \bibitem{GrossglauserT:02}
1921: M.~Grossglauser and D.~Tse.
1922: \newblock {Mobility Increases the Capacity of AdHoc Wireless Networks}.
1923: \newblock {\em IEEE Trans. Networking}, 10(4):477--486, 2002.
1924:
1925: \bibitem{GroetschelLS:88}
1926: M.~Gr{\"o}tschel, L.~Lov\'{a}sz, and A.~Schrijver.
1927: \newblock {\em {Geometric Algorithms and Combinatorial Optimization}}.
1928: \newblock Springer Verlag, 1988.
1929:
1930: \bibitem{GuptaK:98}
1931: P.~Gupta and P.~R. Kumar.
1932: \newblock {Critical Power for Asymptotic Connectivity in Wireless Networks}.
1933: \newblock In W.~M. McEneany, G.~Yin, and Q.~Zhang, editors, {\em Stochastic
1934: Analysis, Control, Optimization and Applications: A Volume in Honor of W. H.
1935: Fleming}. Birkhauser, 1998.
1936:
1937: \bibitem{GuptaK:00}
1938: P.~Gupta and P.~R. Kumar.
1939: \newblock {The Capacity of Wireless Networks}.
1940: \newblock {\em IEEE Trans. Inform. Theory}, 46(2):388--404, 2000.
1941:
1942: \bibitem{GuptaK:01}
1943: P.~Gupta and P.~R. Kumar.
1944: \newblock {Towards an Information Theory of Large Networks: An Achievable Rate
1945: Region}.
1946: \newblock {\em IEEE Trans. Inform. Theory}, 49(8):1877--1894, 2003.
1947:
1948: \bibitem{KargerP:95}
1949: D.~Karger and S.~Plotkin.
1950: \newblock {Adding Multiple Cost Constraints to Combinatorial Optimization
1951: Problems, with Applications to Multicommodity Flows}.
1952: \newblock In {\em Proc. ACM Symp. Theory of Computing (STOC)}, 1995.
1953:
1954: \bibitem{KoetterM:03}
1955: R.~Koetter and M.~M\'edard.
1956: \newblock {An Algebraic Approach to Network Coding}.
1957: \newblock {\em IEEE/ACM Trans. Networking}, 11(5):782--795, 2003.
1958:
1959: \bibitem{KulkarniV:04}
1960: S.~R. Kulkarni and P.~Viswanath.
1961: \newblock {A Deterministic Approach to Throughput Scaling in Wireless
1962: Networks}.
1963: \newblock IEEE Trans. Inform. Theory, to appear. Available from
1964: \href{http://www.ifp.uiuc.edu/~pramodv/} {{\tt
1965: http://www.ifp.uiuc.edu/\~{}pramodv/}}.
1966:
1967: \bibitem{LevequeT:05}
1968: O.~L\'ev\^eque and I.~E. Telatar.
1969: \newblock {Information-Theoretic Upper Bounds on the Capacity of Large Extended
1970: {\em Ad Hoc} Wireless Networks}.
1971: \newblock {\em IEEE Trans. Inform. Theory}, 51(3):858--865, 2005.
1972:
1973: \bibitem{LiL:04b}
1974: Z.~Li and B.~Li.
1975: \newblock {Network Coding: the Case of Multiple Unicast Sessions}.
1976: \newblock In {\em Proc. 42nd Allerton Conf. on Communication, Control and
1977: Computing}, Urbana, IL, 2004.
1978:
1979: \bibitem{MergenT:02}
1980: G.~Mergen and L.~Tong.
1981: \newblock {On the Capacity of Regular Wireless Networks with Transceiver
1982: Multipacket Communication}.
1983: \newblock In {\em Proc. IEEE Int. Symp. Inform. Theory (ISIT)}, Lausanne,
1984: Switzerland, 2002.
1985:
1986: \bibitem{MotwaniR:95}
1987: R.~Motwani and P.~Raghavan.
1988: \newblock {\em {Randomized Algorithms}}.
1989: \newblock Cambridge University Press, 1995.
1990:
1991: \bibitem{PerakiS:03}
1992: C.~Peraki and S.~D. Servetto.
1993: \newblock {On the Maximum Stable Throughput Problem in Random Networks with
1994: Directional Antennas}.
1995: \newblock In {\em Proc. ACM MobiHoc}, Annapolis, MD, 2003.
1996:
1997: \bibitem{PerakiS:04}
1998: C.~Peraki and S.~D. Servetto.
1999: \newblock {Capacity, Stability and Flows in Large-Scale Random Networks}.
2000: \newblock In {\em Proc. IEEE Inform. Theory Workshop (ITW)}, San Antonio, TX,
2001: 2004.
2002:
2003: \bibitem{Pollard:84}
2004: D.~Pollard.
2005: \newblock {\em {Convergence of Stochastic Processes}}.
2006: \newblock Springer-Verlag, 1984.
2007:
2008: \bibitem{Robacker:56}
2009: J.~T. Robacker.
2010: \newblock {Concerning Multicommodity Networks}.
2011: \newblock Research Memo RM-1799, Rand Corp., Santa Monica, CA, 1956.
2012:
2013: \bibitem{TongZM:01}
2014: L.~Tong, Q.~Zhao, and G.~Mergen.
2015: \newblock {Multipacket Reception in Random Access Wireless Networks: From
2016: Signal Processing to Optimal Medium Access Control}.
2017: \newblock {\em IEEE Communications Magazine}, 39(11):108--112, 2001.
2018:
2019: \bibitem{Toumpis:PhD}
2020: S.~Toumpis.
2021: \newblock {\em {Capacity and Cross-Layer Design of Wireless Ad Hoc Networks}}.
2022: \newblock PhD thesis, Stanford University, 2003.
2023:
2024: \bibitem{ToumpisG:02}
2025: S.~Toumpis and A.~J. Goldsmith.
2026: \newblock {Capacity Regions for Wireless Adhoc Networks}.
2027: \newblock {\em IEEE Trans. Wireless Comm.}, 2(4):736--748, 2003.
2028:
2029: \bibitem{TsybakovM:79}
2030: B.~S. Tsybakov and V.~A. Mikhailov.
2031: \newblock {Ergodicity of a Slotted ALOHA System}.
2032: \newblock {\em Problemy Peredachi Informatsii}, 15(4):301--312, 1979.
2033:
2034: \bibitem{VapnikC:71}
2035: V.~N. Vapnik and A.~{Ya}. Chervonenkis.
2036: \newblock {On the Uniform Convergence of Relative Frequencies of Events to
2037: their Probabilities}.
2038: \newblock {\em Theory of Probability and its Applications}, XVI(2):264--280,
2039: 1971.
2040:
2041: \bibitem{WieselthierNE:02}
2042: J.~E. Wieselthier, G.~D. Nguyen, and A.~Ephremides.
2043: \newblock {Energy-Efficient Broadcast and Multicast Trees in Wireless
2044: Networks}.
2045: \newblock {\em ACM/Kluwer Mobile Networks and Applications}, 7(6), 2002.
2046:
2047: \bibitem{XieK:02}
2048: L.-L. Xie and P.~R. Kumar.
2049: \newblock {A Network Information Theory for Wireless Communication: Scaling
2050: Laws and Optimal Operation}.
2051: \newblock Submitted to the IEEE Trans. Inform. Theory, April 2002. Available
2052: from \href{http://decision.csl.uiuc.edu/~prkumar/} {{\tt
2053: http://decision.csl.uiuc.edu/}}
2054: \href{http://decision.csl.uiuc.edu/~prkumar/} {{\tt \~{}prkumar/}}.
2055:
2056: \bibitem{XueK:02}
2057: F.~Xue and P.~R. Kumar.
2058: \newblock {The Number of Neighbors Needed for Connectivity of Wireless
2059: Networks}.
2060: \newblock Submitted to Wireless Networks, April 2002. Available from
2061: \href{http://decision.csl.uiuc.edu/~prkumar/} {{\tt
2062: http://decision.csl.uiuc.edu/\~{}prkumar/}}.
2063:
2064: \end{thebibliography}
2065:
2066:
2067: \end{document}
2068:
2069: