cs0601012/v3.tex
1: \documentclass[11pt,onecolumn]{IEEEtran}
2: \usepackage{subfigure}
3: \usepackage{epsfig,graphicx,psfrag}
4: \usepackage{amsfonts,amsmath,amssymb}
5: %amsthm}
6: %\usepackage{amsfonts,amsmath,amssymb,amsthm}
7: \newtheorem{lemma}{Lemma}
8: \newtheorem{theorem}{Theorem}
9: %\theoremstyle{definition}
10: \newtheorem{cor}{Corollary}
11: \newtheorem{assumption}{Assumption}
12: \newtheorem{conjecture}{Conjecture}
13: \newtheorem{definition}{Definition}
14: \newtheorem{example}{Example}
15: \newcommand{\tends}{{\rightarrow}}
16: \newcommand{\hGe}{\hat{G}^\epsilon}
17: \newcommand{\hEe}{\hat{E}^\epsilon}
18: \newcommand{\lf}{\left}
19: \newcommand{\rf}{\right}
20: \newtheorem{corollary}{Corollary}
21: \newtheorem{condition}{Condition}
22: \newcommand{\bone}{{\mathbf 1}}
23: 
24: 
25: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
26: %delimiters
27: 
28: \let\bl\bigl
29: \let\bbl\Bigl
30: \let\bbbl\biggl
31: \let\bbbbl\Biggl
32: \let\br\bigr
33: \let\bbr\Bigr
34: \let\bbbr\biggr
35: \let\bbbbr\Biggr
36: \let\eps\epsilon
37: \let\eps\epsilon
38: \newcommand{\I}{\mathcal{I}}
39: \newcommand{\E}{\mathbb{E}}
40: \newcommand{\R}{\mathbb{R}}
41: \newcommand{\C}{\mathbb{C}}
42: \newcommand{\cC}{\mathcal{C}}
43: \newcommand{\cA}{\mathcal{A}}
44: 
45: \newcommand{\V}{\mathbb{V}}
46: \renewcommand{\l}{\lambda}
47: \newcommand{\N}{\mathbb{N}}
48: \newcommand{\Z}{\mathbb{Z}}
49: \newcommand{\F}{\mathbb{F}}
50: \renewcommand{\P}{\mathbb{P}}
51: \newcommand{\cP}{\mathcal{P}}
52: \newcommand{\HH}{\mathbb{H}}
53: \renewcommand{\S}{\mathbb{S}}
54: \newcommand{\cS}{\mathcal{S}}
55: 
56: %\pagestyle{empty}
57: \newcommand{\ie}{\emph{i.e.}}
58: \newcommand{\eg}{\emph{e.g.}}
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60: \begin{document}
61: \title{\huge Product Multicommodity Flow in Wireless Networks}
62: \author{Ritesh Madan, Devavrat Shah, and Olivier Leveque
63: \thanks{R. Madan is at QUALCOMM-Flarion Technologies, D. Shah is with Departments of EECS and ESD at MIT, and Olivier
64: Leveque is at EPFL. Emails:~{ \tt rkmadan@stanfordalumni.org, devavrat@mit.edu,
65: olivier.leveque@epfl.ch.}
66:  }}
67: %\date{}
68: \maketitle
69: 
70: %\markboth{Submitted to the IEEE Trans. on Information
71: %Theory}{Submitted to the IEEE Trans. on Information Theory }
72: 
73: \begin{abstract}
74: 
75: We provide a tight approximate characterization of the
76: $n$-dimensional product multicommodity flow (PMF)
77: region for a wireless network of $n$ nodes. Separate
78: characterizations in terms of the spectral properties of
79: appropriate network graphs are obtained in both an information theoretic
80: sense and for a combinatorial interference model (\eg, Protocol
81: model). These provide an inner approximation to the $n^2$ dimensional
82: capacity region. These results answer the following
83: questions which arise naturally from previous work: (a) What is the
84: significance of $1/\sqrt{n}$ in the scaling laws for the Protocol
85: interference model obtained by Gupta and Kumar (2000)? (b) Can we
86: obtain a tight approximation to the ``maximum supportable flow" for
87: node distributions more general than the geometric random
88: distribution, traffic models other than randomly chosen
89: source-destination pairs, and under very general assumptions on the
90: channel fading model?
91: 
92: We first establish that the random source-destination model is
93: essentially a one-dimensional approximation to the capacity region,
94: and a special case of product multi-commodity flow. For
95: a wireline network (graph), a series of results starting from the result
96: of Leighton and Rao (1988) relate the product multicommodity flow to
97: the spectral (or cut) property of the graph. Building on these results,
98: for a combinatorial interference model given by a network
99:     and a conflict
100: graph, we relate the product multicommodity flow to the spectral
101: properties of the underlying graphs resulting in computational upper
102: and lower bounds. These results show that the $1/\sqrt{n}$ scaling
103: law obtained by Gupta and Kumar for a geometric random network can
104: be explained in terms of the combinatorial properties of a
105: geometric random network and the scaling law of the \emph{conductance}
106: of a grid graph. For the more interesting random fading
107: model with additive white Gaussian noise (AWGN), we show that the
108: scaling laws for PMF can again be tightly characterized by the
109: spectral properties of appropriately defined graphs. As an
110: implication, we obtain computationally efficient upper and lower
111: bounds on the PMF for any wireless network with a guaranteed
112: approximation factor.
113: 
114: 
115: \end{abstract}
116: 
117: \begin{keywords}
118: Product multicommodity flow, wireless network, scaling law, capacity
119: region.
120: \end{keywords}
121: % Note that keywords are not normally used for peer review papers.
122: 
123: % For peer review papers, you can put extra information on the cover
124: % page as needed:
125: % \begin{center} \bfseries EDICS Category: 3-BBND \end{center}
126: %
127: % For peerreview papers, inserts a page break and creates the second title.
128: % Will be ignored for other modes.
129: %\IEEEpeerreviewmaketitle
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131: \section{Introduction}
132: 
133: \subsection{Prior Work}\label{s:intro} \vspace{.1in}
134: 
135: An important open question in network information theory is that of
136: characterizing the capacity region of a wireless network of $n$
137: nodes, i.e., the set of all achievable rates between the $n^2$ pairs of
138: nodes in terms of the joint statistics of the channels between
139: these nodes. This has proved to be a very challenging question; even
140: the capacity of a relay network comprising of three nodes is
141: not known in complete generality.
142: 
143: Instead of trying to characterize the
144: capacity region for a general wireless network, the seminal paper by
145: Gupta and Kumar~\cite{GK} concentrated on obtaining the maximum achievable rate
146: for a particular communication model, geometric random distribution
147: of nodes, and randomly chosen source-destination pairs. They showed
148: that the maximum rate for the protocol interference model scales as
149: $\Theta(1/\sqrt{n})$ for $n$ nodes randomly placed on a sphere of
150: unit area. This precise characterization has been followed by many
151: interesting results for both combinatorial interference models and
152: the random fading information theoretic model for large random
153: networks; these include~\cite{GT, KV, EMPS_journal,F, GV, XXK} for
154: communication theoretic models, and~\cite{LT,XK} for
155: information theoretic results. These results are crucially based on
156: the assumption that a large number of nodes are randomly distributed
157: in a certain region, and the inherent symmetry in the random
158: source-destination pair traffic model.
159: 
160: 
161: Since the relative locations of wireless nodes play an
162: important role in the characterization of the capacity region, the
163: notion of transport capacity was defined in~\cite{GK}. A scaling law
164: for the transport capacity for the protocol interference model was
165: obtained in~\cite{GK}, while that for an information theoretic
166: setting was obtained in~\cite{XXK}. The transport capacity can be used
167: to obtain an upper bound on the achievable rate-region for
168: certain rate-tuples, but is not of much use in determining the feasibility of
169: a certain rate-tuple.
170: More
171: recently, information theoretic  outer bounds to the capacity
172: region of a wireless network with a finite number of nodes were
173: obtained in~\cite{AJV} for any wireless network, using the cut-set
174: bound~\cite[Ch. 14]{Cover_book}. We note that any achievable scheme
175: can be used to obtain a set of lower bounds. While the above is only
176: a discussion of a representative set of results in this area
177: (see~\cite{XK_mono} for a more detailed summary), we note that there
178: is no result which provides upper and lower bounds with a guaranteed
179: approximation factor for a general wireless network with a generic
180: random fading model. In this paper, we take the first steps towards
181: providing such a tight characterization under very general
182: assumptions. In doing so, we make  connections between
183: spectral graph theoretic results and network information theory.
184: This results in efficient methods to compute the above tight upper
185: and lower bounds.
186: 
187: %\vspace{.05in}
188: 
189: \subsection{Contribution and Organization}
190: \vspace{.1in}
191: 
192: In Section~\ref{sec1}, we consider the product multicommodity flow (PMF)
193: as an $n$ dimensional approximation of the $n^2$ dimensional
194: capacity region. We show that the random souce-destination pair
195: traffic model is a special case of PMF and it is essentially a
196: one-dimensional approximation of the capacity region.
197: 
198: In Section~\ref{sec2}, we study the PMF for an arbitrary topology and a general
199: combinatorial interference model, of which the protocol model is a
200: special case. We show that the normalized cut capacity
201: (equivalently conductance) of a capacitated network graph induced by
202: the node placement and the interference model characterizes the PMF
203: (within $\log n$ factor). For this model, we also obtain a precise
204: scaling law for average delay using very elementary and almost
205: structure independent arguments.
206: We provide, possibly simpler, re-derivation of the (weaker by $\log^{2.5}n$ factor)
207: lower bound on the maximum flow obtained by Gupta and Kumar for randomly chosen
208: permutation flow on a geometric random graph with a protocol interference model.
209: Our derivation illustrates the connections between the combinatorial properties of geometric random
210: graphs and the maximum PMF.
211: While we do not discuss in detail in this
212: paper, the spectral properties of appropriate induced capacitated
213: graphs characterize the scaling laws obtained for mobile networks
214: in~\cite{GT}, \cite{DGT}.
215: 
216: 
217: In Section~\ref{sec3}, we address the question of characterizing the
218: PMF for a wireless network with Gaussian channels and random fading.
219: This is substantially more challenging than for the combinatorial
220: interference model because there is no obvious underlying network
221: graph that specifies the links which should be used for data
222: transmission. We construct a capacitated graph whose cut capacity
223: characterizes (in terms of tight upper and lower bounds) the PMF in
224: the wireless network. This construction allows one to use classical
225: network flow arguments to characterize and compute the PMF. We
226: illustrate the generality of our results by obtaining scaling laws for a
227: geometric random network and for a network where the number of
228: commodities is constant.
229: 
230: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
231: \section{Traffic Flows}\label{sec1}
232: 
233: In this section, we describe a class of traffic flows, namely,
234: the class of product multicommodity flows, that we study in this paper, and its
235: relevance. Consider a wireless network of $n$ nodes and denote the
236: node set as $V=\{1,\hdots,n\}$. A traffic matrix $\lambda =
237: [\lambda_{ij}] \in \R_+^{n\times n}$ is said to be {\em feasible},
238: if for each pair of nodes $(i,j), ~1\leq i, j \leq n$, data can be
239: transmitted from node $i$ to node $j$ at rate $\lambda_{ij}$. Note
240: that whether a traffic matrix $\lambda$ is feasible or not depends
241: on the model for the underlying wireless network, and we shall
242: describe the precise models for wireless networks in the later
243: sections.
244: 
245: We denote the capacity region by  $\Lambda$, i.e., $\Lambda$ is the
246: set of all feasible traffic matrices. Ideally, we would like to
247: characterize $\Lambda$. However, this is a hard problem in most cases. Instead, we
248:  characterize an {\em approximation} of $\Lambda$ under
249: general assumptions on the wireless network. For this, we consider
250: product multicommodity flow (PMF), defined as follows.
251: 
252: \vspace{.1in}
253: 
254: \begin{definition}[Product Multicommodity Flow (PMF)]
255: Let node $i$ be assigned a weight $\pi(i)$, for $1\leq i\leq n$.
256: Then the PMF corresponding to the weights $\pi\in \R_{+}^n$ and a flow rate $f\in \R_{+}$ is given by the function~\cite{LR99}
257: $M: \R^{n+1} \mapsto \R^{n \times n} $:
258: \[ M(f,\pi) = f \left[ \begin{array}{llll}
259: 0 & \pi(1)\pi(2) &\hdots & \pi(1)\pi(n)\\
260: \pi(2)\pi(1) & 0 & \hdots & \pi(2)\pi(n)\\
261: \vdots & \vdots &\vdots &\vdots\\
262: \pi(n) \pi(1) & \pi(n)\pi(2)&\hdots& \ \ 0
263: \end{array}\right].
264: \]
265: \end{definition}
266: \vspace{.1in} The PMF is an $n$-dimensional approximation of the
267: $n^2$ dimensional capacity region $\Lambda$ with {\em product}
268: constraints. An important special case arises when
269: all the weights are 1, i.e., $\pi(i) = 1$ for $i=1,\hdots,n$. We call
270: such a flow \emph{uniform multicommodity flow (UMF)}. \vspace{.1in}
271: 
272: \begin{definition}[Uniform Multicommodity Flow (UMF)]
273: UMF with flow rate $f\in \R_{+}$ is given by $U(f) = f {\bf 1}$, where
274: ${\bf 1}\in \mathbb{R}_+^{n\times n}$ is a matrix with all
275: entries equal to 1.
276: \end{definition}
277: \vspace{0.1in}
278: We denote by $f_\pi^*$ as the supremum over the flow rates for which the PMF
279: corresponding to the weights $\pi$ is feasible, i.e.,
280: \[ f^*_\pi = \sup \{ f\in R_{+}~:\ M(f, \pi) \text{ is feasible}\}.\]
281: We abuse notation and denote the corresponding quantity for UMF as
282: simply $f^*$.
283: 
284: 
285: 
286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
287: 
288: 
289: 
290: 
291: \subsection{Inner Approximation to $\Lambda$}
292: 
293: 
294: 
295: 
296: We first show that the maximum UMF $f^*$ is a one-parameter
297: approximation to the capacity region $\Lambda$. Consider the
298: following parameter defined in terms of the capacity region as
299: follows.
300: 
301: \vspace{.1in}
302: 
303: \begin{definition}[$\rho^*$]
304: For any $\lambda \in \R^{n\times n}_+$, let $ \rho(\lambda)
305: \stackrel{\triangle}{=} \max_{i} \left\{  \sum_{k=1}^n \lambda_{ik},
306: \sum_{k=1}^n \lambda_{ki} \right\}.$ Let \\\mbox{$L(x) =
307: \{\lambda\in\R_+^{n\times n}: \rho(\lambda)\leq x\}$}. Then,  define
308: $\rho^*$ as
309: $$\rho^* = \sup\{x\in \R_+: L(x)\subseteq \Lambda\}.$$
310: \end{definition}
311: \vspace{.1in} Thus the quantity $\rho^*$ is a parametrization of a
312: (regular) polyhedral inner approximation to the capacity region
313: $\Lambda$. It is tight in the sense for any $x>\rho^*$, there is an
314: infeasible traffic matrix in the set $L(x)$.
315: %Since the set
316: %$\{x\in \R_+: L(x)\subseteq \Lambda\}$ is closed and bounded, any $\lambda$
317: %with $\rho(\lambda)\leq \rho^*$ is feasible.
318: 
319: Roughly speaking, the following result shows that UMF $f^*$ and
320: $\rho^*$ are equally good approximations to the capacity region
321: $\Lambda$.
322: 
323: \vspace{.1in}
324: \begin{lemma}
325: \label{lem0} If $U(f)$ is feasible, then any $\lambda\in
326: \R_{+}^{n\times n}$ such that $\rho(\lambda)\leq nf/2$ is feasible.
327: \end{lemma}
328: \begin{proof}
329: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330: Consider any $\lambda$ such that $\rho(\lambda) \leq nf/2$. Suppose
331: that $U(f)$ is feasible. Then there exists a transmission scheme such
332: supports $U(f)$.
333: We now consider the two stage routing scheme of Valiant and Brenber \cite{valiant}
334: which routes $U(\rho(\lambda)/n)$ in each stage. Since $U(f)$ is feasible,
335: any $\lambda$ with $\rho(\lambda) \leq nf/2$ is supportable by time sharing
336: between the two transmission schemes corresponding to the two stages.
337: To complete the
338: proof of the Lemma, we need to describe this two stage routing
339: scheme.
340: 
341: In the first stage, each node $i$ sends data to all the remaining
342: nodes uniformly (ignoring its actual destination). Thus, node $i$
343: sends data to any node $j$ at rate $\sum_{k} \lambda_{ik}/n \leq
344: \rho(\lambda)/n$.  In the second stage, a node, say $j$, on
345: receiving data (from the first stage) from any source $i$ sends it
346: to the appropriate destination. It is easy to see that due to the
347: uniform spreading of data in the first stage, each node say $j$
348: routes data at rate $\sum_k \lambda_{ki}/n \leq \rho(\lambda)/n$ to
349: node $i$ in the second stage. Thus, the traffic matrices routed in
350: both the stages are dominated by $U(\rho(\lambda)/n)$. That is, the
351: sum traffic matrix is dominated by $U(2\rho(\lambda)/n)$. Hence, if
352: $U(f)$ is feasible then $\rho(\lambda)\leq nf/2$ is feasible. This
353: completes the proof of Lemma \ref{lem1}.
354: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
355: \end{proof}
356: 
357: 
358: \vspace{.1in}
359: \begin{theorem}
360: \label{lem1}
361: $f^*$ and $\rho^*$ are related as
362: \[ \frac{nf^*}{2} \leq \rho^*  \leq nf^* \]
363: \end{theorem}
364: \begin{proof} Note that in general, the capacity region $\Lambda$ may not be
365: closed, and so we need a more careful argument\footnote{We present such a formal
366: argument only once; similar arguments are implicit in many results that follow}.
367: We first show that $\frac{nf^*}{2} \leq \rho^*$. By definition of
368: $\sup$ it follows that for any $\epsilon>0$, $U(f^* - n\epsilon/2)$
369: is feasible. Hence, from Lemma~\ref{lem0}, any $\lambda\in
370: \R_{+}^{n\times n}$ such that $\rho(\lambda)\leq nf^*/2 - \epsilon$
371: is feasible. Hence, again using the definition of $\sup$,
372: $\rho^*\geq nf^*/2$.
373: 
374: Now for the other bound, assume that $\rho^*  > nf^*$, and $\epsilon
375: = (\rho^* - nf^*)/2$. Then, by definition of $\sup$ and $\rho$,
376: $U(nf^* + \epsilon/2)$ is feasible, which is a contradiction. Hence,
377: it follows that $\rho^*  \leq nf^*$.
378: 
379: 
380: 
381: 
382: \end{proof}
383: 
384: 
385: 
386: \vspace{.1in}
387: Thus, bounds on $f^*$ give bounds on $\rho^*$ which differ by at most a factor of 2. Subsequently,
388: a  scaling law for $f^*$ as a function of $n$ is the same as a scaling law for $\rho^*$,
389: i.e. $f^* = \Theta(\rho^*)$ as a function of $n$.
390: 
391: 
392: The set of all feasible PMF clearly provides an $n$ dimensional
393: inner approximation to the capacity region, which is, in general,
394: $n^2$ dimensional. Thus the characterization of the set of feasible
395: PMFs provides a much better approximation to the capacity region
396: than that the one-dimensional approximation given by set of feasible
397: UMF. We next establish the equivalence of UMF with a traffic model
398: with a randomly chosen permutation flow.
399: 
400: \vspace{.1in}
401: 
402: \subsubsection{UMF and Random Permutation Flow}
403: 
404: In some previous work, (e.g.,~\cite{GK}), the capacity scaling laws
405: were derived for the case where $n$ distinct source-destination
406: pairs are chosen at random such that each node is a source
407: (destination) for exactly one destination (source) and such a
408: pairing is done uniformly at random over all possible such pairings.
409: Thus the traffic matrix corresponds to a randomly chosen permutation
410: flow which is defined as follows.
411: 
412: \vspace{.1in}
413: 
414: \begin{definition}[Permutation Flow]
415: Let $S_n$ denote the set of permutation matrices in $\mathbb{R}_+^{n\times n}$.
416: Then the permutation flow corresponding to a permutation $\Sigma\in S_n$ and flow rate
417: $f \in \R_+$ is given by  $S(f,\Sigma)= f \Sigma$.
418: \end{definition}
419: \vspace{.1in}
420: Many previous works  study the scaling of $\bar{f}$, where $\bar{f}$
421: is the supremum over the set of $f \in \R_+$  such when a permuation
422: $\Sigma$ is randomly chosen from $S_n$, the permutation flow $S(f,\Sigma)$
423: is feasible with probability at least $1-1/n^2$. We now show that
424: when a permutation flow with flow rate $nf$ and a randomly chosen permutation
425: is feasible with a high enough probability, then the uniform mulicommodity
426: flow $U(f)$ can be ``almost" supported when $n$ is large enough.
427: \vspace{.1in}
428: \begin{lemma}\label{lem2}
429: For $\Sigma \in S_n$ chosen uniformly at random, if $(nf)\Sigma$
430: is feasible with probability at least $1-n^{-1-\alpha}, \alpha > 0$, then
431: there exists a sequence of feasible rate matrices $\Gamma_n$ such that
432: \[  \| {U_n(f) - \Gamma_n \|} = O(f n^{-\alpha}) \to 0 \mbox{~as~$n\to\infty$,}\]
433: where $\| \cdot \|$ denotes the standard 2-norm for matrices\footnote{Given
434: a matrix $M \in \R^{n \times n}$, the 2-norm of $M$ is $\| M \| = \sup\{ \| Mx \| : x \in \R^n , \|x\| = 1 \}$,
435: where $\|x\|$ is the $\ell_2$ norm of vector $x \in \R^n$.}, and
436: $U_n(f)$ is the uniform multicommodity flow for $n$ nodes.
437: \end{lemma}
438: \begin{proof}
439: From the hypothesis of the Lemma, it is clear that for at least
440: $(1-n^{-1-\alpha})$ fraction of all $n!$ permutations in $S_n$, the
441: permutation flow $(nf)\Sigma$ is feasible. By definition and
442: symmetry of permutations, we can write
443: \[ U_n(f) = \frac{1}{n!}\sum_{i=1}^{n!} (nf)\Sigma_i.\]
444: Let us define the following indicator function
445: \[{\bf 1}_i = \left\{\begin{array}{ll}
446: 1&\quad\text{$(nf)\Sigma_i$ is supportable}\\
447: 0&\quad\text{otherwise}
448: \end{array}\right. \]
449: Consider a uniform time sharing scheme between all the $n!$
450: permutation flows. Then the following traffic matrix is supportable.
451: \[ \Gamma_n = \frac{1}{n!}\sum_{i=1}^{n!} {\bf 1}_i (nf)\Sigma_i\]
452: Thus
453: \begin{eqnarray*}
454: \|{U_n(f) - \Gamma_n\|} &=  &\left|\left|\frac{1}{n!} \sum_{i=1}^{n!}(1- {\bf 1}_i)(nf)\Sigma_i\right |\right|\\
455:                            &\stackrel{(a)}{\leq} & \frac{1}{n!}\sum_{i=1}^{n!} \left|\left| (1- {\bf 1}_i)(nf)\Sigma_i\right |\right|\\
456:                            &\stackrel{(b)}{=} & \frac{1}{n!}\sum_{i=1}^{n!}(1- {\bf 1}_i) nf ~
457:                             \leq ~ \frac{nf}{n!}\frac{n!}{n^{1+\alpha}}\\
458:                            &= & \frac{f}{n^\alpha}
459:                             ~\rightarrow 0, ~~\mbox{as $n\to\infty$.}
460: \end{eqnarray*}
461: Step (a) uses  triangle inequality for a norm and step (b) uses
462: $||\Sigma_i||=1$ for any permutation matrix $\Sigma_i$.
463: \end{proof}
464: 
465: 
466: 
467: \vspace{.1in} From Lemma~\ref{lem1}, if $U(f)$ is feasible, then
468: $S(nf/2, \Sigma)$ is feasible for all $\Sigma\in S_n$. Thus, using
469: an argument identical to that in the proof of Theorem~\ref{lem1}, a
470: scaling law for $\bar{f}$ is equivalent to a scaling law for $f^*$,
471: i.e.
472: $$ f^* = \Theta(n\bar{f}).$$
473: 
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475: 
476: 
477: 
478: \subsection{Wireline Networks: PMF Over a Graph}
479: 
480: We briefly review the key results known for PMF on graphs with
481: fixed edge capacities. These results will be useful in our analysis
482: for PMF for wireless networks.
483: 
484: Consider a directed graph $G=(V,E)$, where an edge $(i,j)\in E$ has
485: a capacity $C(i,j)$. Also, for $(i,j)\notin E$, we take $C(i,j)=0$.
486:  Then for a given $\pi$,
487: $f_\pi^*$ for graph $G=(V,E)$ is given by the solution of the
488: following linear program (LP).
489: %which can be solved in polynomial time~\cite{karmarkar_1984}.
490: \[\begin{aligned}
491:  \text{maximize}  \qquad & \qquad \qquad f,\\
492: \text{subject to} \qquad &
493: \sum_{k: (i,k)\in E}\bbl( x_{ij}(i,k) - x_{ij}(k,i) \bbr)  = f\pi(i)\pi(j), \ 1\leq i,j \leq n,\\
494: & \sum_{m: (k,m)\in E}\bbl( x_{ij}(k,m) - x_{ij}(m,k) \bbr)  = 0, \ \forall k\neq i,j,\ 1\leq i,j\leq n,\\
495: & \sum_{i=1}^n\sum_{j=1}^n x_{ij}(k,m)  \leq C(k,m), \ \forall (k,m)\in E,
496: \end{aligned}\]
497: where the variables are $f$ and $\{x_{ij}(k,m): (k,m)\in E, i,j,k,m = 1,\hdots, n\}$. The first two
498: are flow conservation constraints and the third one is the capacity constraint. The total number of variables is
499: less than $2n^4$ and the total number of constraints is less than $(n^3 + 2n^2)$. Hence, the above LP can
500: be solved in polynomial time~\cite{karmarkar_1984}.
501: 
502: The well-known max-flow min-cut characterization for a single
503: commodity flow naturally gives rise to the following question.
504: Though the maximum PMF $f^*_\pi$ for a given weight
505: vector can be computed in polynomial time, is there a corresponding
506: result that relates $f^*_\pi$ and the properties of the graph.
507: In their seminal paper, Leighton
508: and Rao \cite{LR88} obtained a characterization of $f^*_\pi$ in
509: terms of the weighted min-cut of graph. We summarize their
510: main result below. Let $p_\pi = | \{ i \in V : \pi(i) > 0 \}| $,  denote
511: the number of nodes for which the corresponding element of $\pi$
512: is non-zero. Then, without loss of generality we assume that
513: $\sum_{i=1}^n \pi(i) = p_\pi$.
514: 
515: \vspace{.1in}
516: 
517: 
518: \begin{definition}
519: For the graph $G$ and weight vector $\pi$, define the min-cut by
520: \[ \Upsilon(G, \pi) = \min_{U\subseteq V} \frac{\sum_{(i,j):i\in U, j\in U^C} C(i,j)}{\pi(U) \pi(U^c)},\]
521: with notation that $\pi(S) = \sum_{i\in S} \pi(i)$ for any set $S$.
522: \end{definition}
523: \begin{theorem}[Theorem 17, \cite{LR99}]
524: \label{thm:PMF_LR}
525: In any directed graph $G$, the maximum PMF for weight $\pi$ is related to $\Upsilon(G, \pi)$ as follows:
526: \[ \Omega\left( \frac{ \Upsilon(G, \pi)} {\log p_\pi}\right ) \leq f^*_\pi \leq  \Upsilon(G, \pi), \]
527: where the constants for the lower bound do not depend on the graph.
528: %where $n_\pi = \{ i \in V : \pi(i) > 0 \}$.
529: \end{theorem}
530: \vspace{.1in}
531: Note that the upper bound follows easily because for a given PMF $f_\pi$, the total flow
532: from $U$ to $U^C$ is $\pi(U)\pi(U^C)f_\pi$, which has to be less than the total capacity of the
533: links from $U$ to $U^C$.
534: The above characterization was crucial to the design of subsequent
535: approximation algorithms for many NP-hard problems; a summary of these algorithms can be found in~\cite{LR99}. An important case
536: of the above result is when $\pi(i)=1$ for all $i=1,\hdots, n$, i.e., the special case of uniform mulitcommodity flow.
537: In this case, we have
538: \[ \Upsilon(G) = \Upsilon(G, {\bf 1}) = \min_{U\subseteq V} \frac{\sum_{(i,j):i\in U, j\in U^C} C(i,j)}{|U||U^C|},\]
539: and
540: \[ \Omega\left( \frac{ \Upsilon(G)} {\log n}\right ) \leq f^* \leq  \Upsilon(G).\]
541: 
542: \vspace{0.1in}
543: %We note that  identical results can also be derived for directed graphs; the details can again be found in~\cite{LR99}.
544: 
545: 
546: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
547: \section{Combinatorial Interference Model}\label{sec2}
548: 
549: 
550: A combinatorial interference model defines constraints such that simultaneous data transmissions
551: over only certain sets of links (or edges) can be successful. This is a simplified abstraction of
552: a wireless network because in reality whether or not multiple simultaneous data transmissions
553: are successful depends on the rate of data transmission and the interference power at the various receivers.
554: We next describe the combinatorial interference model formally and illustrate it with example scenarios where
555: this abstraction is a reasonable one.
556: 
557: \subsection{Model}
558: \label{subsec:model}
559: A combinatorial interference model for a given set of wireless nodes $V=\{1,\hdots,n\}$
560: defines the following two objects:
561: \begin{itemize}
562: 
563: \item[(a)] A directed graph $G = (V, E)$ where $E$ is the set of directed links (edges)
564: over which data can be transmitted.
565: %Let $\delta$ denote the maximum vertex degree of this graph.
566: 
567: \item[(b)] For each directed edge $e \in E$, let  $\I(e) = \{ \hat{e} \in E \}$ be the set
568: of edges (directed links) that interfere with a transmission on link
569: $e$. Data can be successfully transmitted on link $e$ at rate $W(e)$
570: if and only if no transmission on any link in $\I(e)$ takes place
571: simultaneously. In general, the rate $W(e)$ for a given power
572: constraint can be different for different edges.
573: % Further, it can be a function of the set of links that transmit simultaneously.
574: The
575: proof methods and results of the paper will not change
576: (qualitatively) in this scenario. However, for ease of exposition
577: we will assume $W(e)=1$\footnote{As long
578: as $W(e)$ is bounded below and above by a constant, scaling laws
579: do not change even though the bounds for a given number of nodes $n$ will change.} for all $e \in E$.
580: \end{itemize}
581: 
582: \vspace{0.1in}
583: We assume that for every edge $(i,j)\in E$, edge $(j,i)\in E$,  i.e., the graph $G$ is essentially
584: an undirected graph without the interference constraints given by the sets $\I(e)$'s. This is a reasonable
585: assumption in many time division and frequency division systems where the channels are reciprocal~\cite{molisch_book}.
586: The interference sets $\I((i,j))$ and $\I((j,i))$ may not be identical because the transmissions which interfere
587: with a signal received at node $i$ may not be the same as transmissions which interfere with a signal
588: received at node $j$.
589: 
590: 
591: The above definitions can be used to induce a \emph{dual conflict
592: graph} as follows.
593: 
594: \vspace{0.1in}
595: \begin{definition}
596: The dual conflict graph  is an undirected graph $G^D = (E,E^D)$ with
597: vertex set $E$ and edge set $E^D$, where an edge $e^D\in E^D$
598: exists between $e_1$ and $e_2$ if $e_1$ and $e_2$ cannot transmit
599: simultaneously due to interference constraints.
600: Thus, each link $e\in E$ is connected to all links in
601: $\I(e)$.
602: \end{definition}
603: 
604: \vspace{0.1in}
605: 
606: 
607: For the rest of the section, we will suppress the explicit
608: dependence of all quantities on the combinatorial interference model
609: parameterized by the graphs $G$ and $G^D$; this helps to simplify
610: notation. Let us denote the node degree  and the chromatic number\footnote{The chromatic
611: number of a graph is the minimum number of colors needed to color the vertices of the graph
612: such that no two nodes  of the graph which are connected by an edge share the same color.}
613: dual conflict graph $G^D$ by $ \Delta = \max_{e \in E} |\I(e)|$ and
614: $\kappa$ respectively. Note that $\kappa\leq
615: (1+\Delta)$. %~(see, for example,~\cite{color}).
616: Let $\{E_k\}, E_k
617: \subseteq E$, be the set of all possible link sets that can be active
618: simultaneously, i.e., simultaneous transmissions
619: on all the links in $E_k$ at rate $W(e)=1$ are feasible for
620: the given interference model. Each $E_k$
621: corresponds to a vector $C_k\in \R^{|E|}$, where $C_k(e) =
622: {\mathbf 1}_{\{e \in E_k\}} = {\mathbf 1}_{\{e \in E_k\}} $.
623: Let $\cC$
624: be the convex-hull of all such vectors $\{C_k\}$. Thus $\cC$ is the set of all
625: vectors $C$ such that link capacities $C(e)$ (for link $e$) can be obtained
626: by time-sharing between the $C_k$'s for the given interference model. We then define the capacity
627: region to be the set of traffic matrices which can be routed over the graph $G = (V,E)$, where each edge
628: $e$ has capacity $C(e)$ for $C \in \cC$. The formal definition is as follows.
629: 
630: \vspace{0.1in}
631: \begin{definition}[Capacity Region ($\Lambda$)]
632: \label{def:cap_region}
633: The capacity region is the set of traffic matrices $\lambda \in \R^{n\times n}$ such that the following set
634: of conditions are feasible for some $C\in \cC$:
635: \begin{equation}
636: \label{eqn:cap_reg}
637: \begin{aligned}
638: &\sum_{k: (i,k)\in E}\bbl( x_{ij}(i,k) - x_{ij}(k,i) \bbr) = \lambda_{ij}, \ 1\leq i,j \leq n,\\
639: &\sum_{m: (k,m)\in E}\bbl( x_{ij}(k,m) - x_{ij}(m,k) \bbr) = 0, \ \forall k\neq i,j,\ 1\leq i,j \leq n,\\
640: &\sum_{i=1}^n\sum_{j=1}^n x_{ij}(k,m) \leq C(k,m), \ \forall (k,m)\in E,
641: \end{aligned}\end{equation}
642: where $C(k,m) = C(e)$ with $e=(k,m)\in E$; variables are $\{x_{ij}(k,m): (k,m)\in E, 1\leq i,j,k,m \leq n\}$.
643: \end{definition}
644: \vspace{0.1in} Thus, the capacity region consists of all traffic
645: matrices which are feasible using routing and link scheduling
646: (time-sharing between the sets $\{E_k\}$).  We now illustrate this
647: capacity region by a couple of special cases corresponding to widely
648: used models for wireless networks.
649: 
650: \subsubsection{Protocol Model}
651: The protocol model parameterized by the maximum radius of
652: transmission, $r$, and the amount of acceptable interference,
653: $\eta$, is defined in~\cite{GK} as follows.
654: \begin{itemize}
655:  \item[(a)] A node $i$ can transmit to any
656: node $j$ if the distance between $i$ and $j$, $r_{ij}$ is less than
657: the transmission radius $r$.
658: \item[(b)] For transmission from node $i$ to $j$ to be successful, no other node
659: $k$ within distance $(1+\eta) r_{ij}$ ($\eta > 0$ a constant) of
660: node $j$ should transmit simultaneously.
661: \end{itemize}
662: The corresponding definitions of $E$ and $E_D$ follow. A directed
663: link from node $i$ to node $j$ is in $E$ if $r_{ij}\leq r$. For a
664: link, $e\in E$, let $e^+$ denote the transmitter and let $e^-$
665: denote the receiver. Then
666: %\[{\cal I}(e)=\left\{\hat{e}\in E: \min\bl(d(e^+,\hat{e}^-),d(\hat{e}^+,e^-)\br)\leq (1+\eta)d(e^+,e^-)\right\}.\]
667: \[{\cal I}(e)=\left\{\hat{e}\in E:  r_{\hat{e}^+e^-} \leq (1+\eta)r_{e^+e^-}\right\}.\]
668: Thus the protocol model is a special case of the combinatorial interference model.
669: \vspace{.1in}
670: 
671: \subsubsection{SINR Threshold Model}
672: Assume that all transmissions occur at power $P$, and
673: the channel gain from the transmitter
674: of node $j$ to the receiver of node $i$ is given by $h_{ij}$, i.e.,
675: if node $j$ transmits at power $P$, the received signal power at
676: node $i$ will be $h_{ij}P$. A signal to interference and noise ratio (SINR) threshold model is parametrized by
677: a threshold $\alpha$ such that a transmission from node $i$ to node $j$ is successful if and only if the SINR is above
678: $\alpha$, i.e.,
679: \[ \frac{P h_{ij}}{ \sum_{k\neq i}Ph_{kj}+ N_0B} \geq \alpha \]
680: For example, if we assume that each link transmits Gaussian signals and that the Shannon capacity on each link is achievable,
681: then the threshold is given by $N_0B(2^W-1)$ (assuming $W(e)=W$ for all $e\in E$ as before).
682: 
683: We can define a corresponding combinatorial interference model such that the feasible
684: simultaneous transmissions defined by the combinatorial interference model are a subset of that
685: described by the SINR threshold model.
686: Consider the set of directed links
687: $E_\gamma$ such that a link, $(i,j)$, from node $i$ to node $j$, is
688: in $E_\gamma$ if and only if $h_{ji}\geq \gamma$. Also, define
689: ${\cal I}(e) = \{\hat{e}\in E_\gamma: h_{e^-\hat{e}^+ }\geq \beta\}
690: $.
691: Then link $e$
692: can transmit at rate $W$ if no other links in ${\cal I}(e)$ transmit simultaneously,
693: if and only if $\gamma$ and $\beta$ are such that
694: \begin{equation}
695: \label{eqn:interference}   \frac{P h_{e^-e^+}}
696: {\sum_{\hat{e}\in E_\gamma ,\ \hat{e}\notin {\cal
697: I}(e)}Ph_{e^-\hat{e}^+} + N_0B} \geq \alpha, \qquad \forall e\in
698: E_\gamma
699: \end{equation}
700: It is easy to see that the above condition is satisfied
701: if the following condition holds.
702: \[ \beta \leq \frac{1}{nP}\left(\frac{P\gamma}{2^W-1}-N_0B\right)\]
703: 
704: 
705: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
706: 
707: \subsection{Results}
708: 
709: We now derive results for the combinatorial interference model
710: which relate the maximum PMF $f_\pi^*$ to spectral properties of the
711: underlying graphs induced by the interference model. Most of the results
712: in this section use ideas from known results. While important
713: in their own right, these results and their proofs motivate the
714: results for an information theoretic setting for wireless networks
715: with Gaussian channels. Also, they provide alternate derivations
716: for known capacity scaling laws in random networks. Towards
717: the end of this section, we obtain simple bounds on the delay.
718: 
719: \vspace{.2in}
720: 
721: \subsubsection{Bounds on PMF}
722: For any $C\in \cC$, we denote the maximum PMF
723: on graph $G$ where each edge $e$ has capacity $C(e)$, by $f_\pi(C)$ , and the corresponding
724: min-cut by
725: $$ \Psi_\pi(C) = \min_{S \subset V} \frac{\sum_{(i,j): i \in S, j\in S^c} C(i,j)}{\pi(S)\pi(S^c)}. $$
726: We denote the corresponding quantities for the special case of UMF by $f(C)$ and $\Psi(C)$, respectively.
727: Then we have the following lemmas.
728: 
729: \begin{lemma}
730: \label{lem:cut_continuity}
731: $\Psi_\pi: \cC \mapsto \R$ is a continuous function for any $\pi \geq 0$.
732: \end{lemma}
733: \begin{proof}
734: Consider a cut $S$ such that $\pi(S)\pi(S^C) >0$. Then, the following is a continuous function of $C$:
735: $$\frac{\sum_{(i,j): i \in S, j\in S^c} C(i,j)}{\pi(S)\pi(S^c)}.$$
736: The lemma then follows since the minimum of a finite number of continuous functions is continuous.
737: \end{proof}
738: 
739: 
740: \vspace{0.1in}
741: 
742: \begin{lemma}
743: \label{lem:pmf_continuity}
744: $f_\pi: \cC \mapsto \R$ is a continuous function for any $\pi \geq 0$.
745: \end{lemma}
746: \begin{proof}
747: \iffalse
748: We first show that for any $C \in \cC$, we have the following:
749: \[f_\pi(C) =  \frac{f_\pi(\alpha C)}{\alpha},\]
750: where $\alpha >0$ is chosen such that $\alpha C\in \cC$.
751: By scaling all the variables in~(\ref{eqn:cap_reg}), it follows that
752: the flow $U(\alpha f)$ is feasible for $\alpha C$, and so
753: \[f_\pi(\alpha C) \geq \alpha f_\pi(C).\]
754: For the other direction, we argue by contradiction. Assume that
755: \[f_\pi(C) <  \frac{f_\pi(\alpha C)}{\alpha}.\]
756: Then for the LP in~(\ref{eqn:cap_reg}), for the vector $\alpha C$, we can scale
757: down all variables by $\alpha$, which means that a flow of greater than
758: $f_\pi(C)$ is feasible, which is incorrect.
759: \fi
760: For $\hat{C}\in \cC$ and any $\epsilon>0$ define the set
761: \[ {\cal B}_\delta = \{C \in \cC : ||C - \hat{C}||< \delta\}.\]
762: To prove the lemma, we have to show that for any $\epsilon >0$, there exists a
763: $\delta >0$ such that for all $C\in {\cal B}_\delta$,
764: $|f_\pi(C) - f_\pi(\hat{C})|<\epsilon$.
765: 
766: For  $\hat{C}\in \cC$ consider
767: \[ \delta_1 = \min_{(k,m)\in E} \left\{ \alpha \hat{C}(k,m) : \hat{C}(k,m) >0\right \}, \qquad 0<\alpha < 1,\]
768: and
769: \[ \delta = \min \left\{ \delta_1, \min_{i,j: \pi(i)\pi(j)>0}\frac{\epsilon}{2\pi(i)\pi(j)}\right \}.\]
770: Then for any $C \in {\cal B}_\delta $, it follows from~\ref{eqn:cap_reg} that $f_\pi(C) \leq f_\pi(\hat{C}) + \epsilon$.
771: It only remains to show that $f_\pi(C) \geq  f_\pi(\hat{C}) - \epsilon$.
772: For this note that $C \succ \underline{C}$  for all $C\in {\cal B}_\delta $, where $\underline{C}$ is as follows:
773: \[ \underline{C}(k,m) = \left \{
774: \begin{array}{ll}
775: 0 & \text{C(k,m) = 0}\\
776: \hat{C}(k,m) - \delta_1 & \text{otherwise}
777: \end{array}
778:   \right..\]
779: Now by scaling all the variables by $(1-\alpha)$ in the LP~(\ref{eqn:cap_reg}) for $\hat{C}$ and
780: using the monotonicity of $f_\pi(C)$ in $C$, we can see that
781: $f_\pi(C) \geq (1-\alpha) f_\pi(\hat{C})$ for all $C\in {\cal B}_\delta $. If $f_\pi(\hat{C})=0$,
782: we are done. If not, choose $\alpha = \min(\epsilon/f_\pi(\hat{C}), 0.5)$, which gives
783: $f_\pi(C) \geq  f_\pi(\hat{C}) - \epsilon$, and so we are done.
784: 
785: \end{proof}
786: \vspace{0.1in}
787: 
788: 
789: We now define a quantity for the combinatorial interference model
790: corresponding to the min-cut of a graph.
791: 
792: \vspace{0.1in}
793: \begin{definition}
794: The min-cut for the combinatorial interference model is defined as
795: $$ \Psi_\pi^* = \max_{C \in \cC} \min_{S \subset V} \frac{\sum_{(i,j): i \in S, j\in S^c} C(i,j)}{\pi(S)\pi(S^c)}. $$
796: \end{definition}
797: 
798: \vspace{0.1in}
799: 
800: Note that $\Psi_\pi^*$ is well defined since $\Psi_\pi(C)$ is a continuous function of $C$, and $\cC$ is closed and
801: bounded because it is the convex hull of a finite number of points.
802: The above definition can be interpreted as the min-cut of the graph $G$, where each edge has capacity $C(e)$, and
803: the vector $C$ is chosen from the set $\cC$ such that it maximizes the min-cut of this graph $G$.
804: The following result is an extension of Theorem~\ref{thm:PMF_LR} to combinatorial
805: interference models.
806: \vspace{0.1in}
807: \begin{theorem}
808: \label{thm:UMF_mincut}
809: $f^*_\pi$ is bounded as
810: \begin{equation}
811: \Omega\left(\frac{\Psi_\pi^*}{\log p_\pi} \right) \leq f_\pi^* \leq \Psi_\pi^*.
812: \end{equation}
813: %where we recall that $n_\pi = \{ i \in V : \pi(i) > 0\}$.
814: \end{theorem}
815: \vspace{0.1in}
816: \begin{proof}
817: Since $\cC$ is closed and bounded, it follows from Lemma~\ref{lem:pmf_continuity} that
818: there exists $C^*\in \cC$ such that $f^*_\pi = f_\pi(C^*)$. Then, using
819: Theorem~\ref{thm:PMF_LR}, it follows that
820: \[
821: f^*_\pi \leq   \Psi(C^*) \leq \Psi_\pi^*.
822: \]
823: %where $n_\pi = \{ i \in V : \pi(i) > 0 \}$.
824: 
825: Now, from Lemma~\ref{lem:cut_continuity}, there is $\hat{C}\in \cC$ such that
826: $ \Psi_\pi^* = \Psi_\pi(\hat{C})$. Using Theorem~\ref{thm:PMF_LR}, it follows that
827: \[f_\pi^* \geq  f_\pi(\hat{C}) = \Omega\left(\frac{\Psi_\pi^*}{\log p_\pi } \right). \]
828: This completes the proof of Theorem~\ref{thm:UMF_mincut}.
829: 
830: 
831: \end{proof}
832: 
833: 
834: 
835: Note that unlike the case for wireline networks (or equivalently graphs), $f^*$ is a hard
836: quantity to compute. Also, note that $\Psi$ is a function of both $G$ and the dual graph $G^D$.
837: We next relate the maximum UMF $f^*$ to spectral properties of graphs $G$ and $G^D$.
838: 
839: \vspace{0.1in}
840: 
841: \begin{definition}
842: The conductance of a graph $G$ is defined as follows.
843: \[\Phi(G) = \min_{U\subseteq V, |U|\leq n/2}\frac{\sum_{i\in U, j\in U^C} {\bf 1}_{[(i,j)\in E] }}{|U|},\]
844: where ${\bf 1}_{[.]}$ is the indicator function.
845: \end{definition}
846: %Then we have the following result.
847: \vspace{0.1in}
848: 
849: \begin{corollary}
850: \label{thm:UMF_cond}
851: Recall that $\kappa$ is the chromatic number of the dual graph $G^D$.
852: Then, $f^*$ is related to $\Phi(G)$ as follows.
853: \[\Omega\left(\frac{\Phi(G)}{\kappa n\log n} \right) \leq f^* \leq \frac{\Phi(G)}{n}.\]
854: \end{corollary}
855: \begin{proof}
856: Consider vertex coloring for the dual graph $G^D = (E,E^D)$.
857: The chromatic number of $G^D$ is defined to be $\kappa$ and hence we need
858: $\kappa$ colors for vertex coloring of $G^D$.  Thus we have partitioned
859: the set $E$ into subsets, say, $E_1,\hdots, E_{\kappa}$ such that the links
860: in each subset can transmit simultaneously at rate 1.
861: Now let $C_k(e) = {\bf 1}_{\{e\in E_k\}}$. Then, $C$ corresponding to
862: uniform time-sharing between the $\kappa$ edge sets $E_1,\dots, E_k$
863: is given  by
864: \[ C = \frac{1}{\kappa}(C_1 + \hdots + C_{\kappa}), \]
865: which is a convex combination of $C_1,\hdots,C_{\kappa}\in \cC$.
866: Hence, $C(i,j) = 1/\kappa$ for all $i\neq j$, and $C\in \cC$. Then, using Theorem~\ref{thm:PMF_LR} and the definition
867: of conductance above,
868: \[ f^* \geq f^*(C) \geq \Omega\left( \frac{\Psi}{\log n} \right) = \Omega\left(\frac{\Phi(G)}{\kappa n\log n} \right).\]
869: \end{proof}
870: 
871: For the upper bound, note that for any $C\in \cC$, $C \preceq {\bf 1}$, i.e., $C$ is lexicographically
872: less than ${\bf 1}$, and $f(C_1) \geq f(C_2)$ if $C_1\geq C_2$. Hence, $f^* = \max_{C\in \cC} f(C) \leq f({\bf 1})$. Then, the upper bound follows again
873: by a straightforward use of Theorem~\ref{thm:PMF_LR} and the definition
874: of conductance.
875: \vspace{0.2in}
876: 
877: \subsubsection{Average Delay}
878: 
879: We now provide bounds on the average delay for a class of traffic matrices.
880: We measure delay in number of hops. We assume that the packet size is small
881: enough so that the packet delay is essentially equal to the number of hops
882: taken by the packet. This is similar to the assumptions
883: in, for example, \cite{GK},\cite{GT}, and \cite{EMPS_journal}.
884: 
885: We restrict ourselves to periodic link scheduling schemes (similar
886: arguments extend to any ergodic scheduling scheme as well). For
887: fixed networks, $\cC$ is the convex hull of the set, $\{C_k\}$,
888: which has a finite cardinality. Hence, any vector in $\cC$ can be
889: written as a linear combination of the $C_k$'s. Thus to maximize the
890: supportable uniform multicommodity flow it is sufficient to optimize
891: over transmission schemes with periodic scheduling of links where
892: the periodic schedule corresponds to time division between the
893: $C_k$'s.
894: 
895: We obtain the following general scaling of delay.
896: \vspace{.05in}
897: \begin{theorem}\label{thm4}
898: Let $S(n)$ be the total number of transmissions by the $n$ wireless nodes
899: on average per unit time\footnote{The quantity $S(n)$ is well defined
900: since we consider periodic scheduling of links.}. When data is transmitted according to rate matrix
901: $\lambda\in \Lambda$,  the average delay, $D(n)$, over all packets  scales as
902: $$ D(n) = \Theta\left(\frac{S(n)}{\bar{\lambda}}\right),
903: \mbox{~~where $\bar{\lambda} = \sum_{i,j} \lambda_{ij}$.}$$
904: \end{theorem}
905: \begin{proof}
906: Let $\Gamma$ denote the set of all  possible
907: paths (without cycles) in the network. The amount of flow generated at node $i$
908: to be transmitted to node $j$ is $\lambda_{ij}$. Let us consider an arbitrary but fixed\footnote{Here,
909: we consider a fixed deterministic scheme. However, it is easy to see that the result
910: extends for any randomized scheme as well.}
911: routing scheme where a fraction $\alpha_{ij}^\gamma$ of
912: the flow from node $i$ to node $j$ is routed over path $\gamma \in\Gamma$. We assume that
913: the traffic matrix $\lambda$ is feasible.  Hence, there exists a link scheduling
914: and routing scheme to support it. The total number of transmissions per unit time at node $l$
915: is $\sum_{\gamma \ni l}\sum_{i,j} \alpha_{ij}^\gamma\lambda_{ij}$. Hence, the average
916: number of transmissions per unit time in the entire network, denoted by $S(n)$, is
917: \[ \begin{aligned}
918: S(n) \ =\ \sum_{l=1}^n\sum_{\gamma \ni l}\sum_{i,j}
919: \alpha_{ij}^\gamma\lambda_{ij}
920:      \ =\  \sum_{\gamma} H^\gamma \sum_{i,j} \alpha_{ij}^\gamma \lambda_{ij}\\
921: \end{aligned}
922: \]
923: where $H^\gamma$ is the number of hops on path $\gamma$.
924: The total flow over a path $\gamma$ is $\sum_{i,j} \alpha_{ij}^\gamma\lambda_{ij}$, i.e.,
925: the fraction of total flow over path $\gamma$ is $\sum_{i,j} \alpha_{ij}^\gamma\lambda_{ij}/\bar{\lambda}$.
926: Hence, the average number of hops traversed by all
927: packets is given by
928: \[D(n) = \frac{1}{\bar{\lambda}}\sum_\gamma H^\gamma \sum_{i,j} \alpha_{ij}^\gamma\lambda_{ij} =\frac{S(n)}{\bar{\lambda}}. \]
929: This completes the proof of Theorem \ref{thm4}.
930: \end{proof}
931: 
932: 
933: \vspace{.2in}
934: 
935: We note that the above result uses very little information about the specific underlying
936: transmission scheme.
937: Next, we present an immediate corollary of the above result and Theorem~\ref{thm:UMF_cond} that
938: bounds the delay for a scheme that maximizes the UMF in the network.
939: 
940: \vspace{.05in}
941: 
942: \begin{cor}
943: \label{cor2} Since $f^* = \Omega\left(\frac{\Phi(G) }{\kappa n\log n} \right)$, the
944: corresponding delay scales as $D(n) = O\left( \frac{|E|\log n}{n\Phi(G)} \right)$.
945: \end{cor}
946: \begin{proof}
947: Consider the link scheduling scheme in the proof of
948: Corollary~\ref{thm:UMF_cond}, where we partition the set of links $E$ into
949: subsets $E_1,\hdots,E_\kappa$ such that all the links in each subset
950: $E_i$ can transmit simultaneously. Note that this scheme can support UMF
951: $f = \Omega\left(\frac{\Phi(G)}{\kappa n\log n}\right)$. For this transmission
952: scheme, every link transmits
953: at rate 1 for at most $1/\kappa$ fraction of the time. Hence, we have $ S(n) \leq
954: \frac{|E|}{\kappa}$. Thus, it follows from Theorem~\ref{thm4} that
955: $D(n) = O\left( \frac{|E|\log n}{n\Phi(G)} \right)$.
956: 
957: 
958: \end{proof}
959: 
960: \vspace{.1in}
961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
962: 
963: \subsection{Computational Methods}
964: 
965: We now describe computational methods to obtain bounds on $f^*$ (the extensions
966: to PMF are straightforward). As
967: noted earlier, for wire-line networks, the computation of $f^*$ is
968: equivalent to solving an LP. However, in a wireless network, the
969: link capacity is a function of the link schedule. Since, the number
970: of link schedules is combinatorial, it is a hard problem.
971: Specifically, the question of checking feasibility of a rate vector
972: $\lambda$ was proved to be NP-hard by Arikan \cite{Ar84}, that is
973: there exists an interference model and graph under which checking
974: feasibility of $\lambda$ is NP-hard.
975: % citation: E. Arikan, Some Complexity Results about Packet Radio Networks,
976: % IEEE Transaction on Information Thoery, 30:4, pp: 681-685.
977: Motivated by this, here we address the question of providing a
978: simple computational method to bound $f^*$.
979: 
980: We use ideas of node coloring to induce a link schedule in a way
981: similar to, for example,~\cite{kodialam_2003}. In particular,
982: we can obtain an upper bound $f_1^*$ and a lower
983: bound $f_2^*$ for maximum UMF $f^*$ in polynomial time such that
984: \[ f_1^* \leq \kappa f_2^* \]
985: The upper bound can be computed by solving the LP in~(\ref{eqn:cap_reg})
986: with $C(e)=1$ for all $e\in E$. For the lower bound, since the dual
987: graph $G^D$ has chromatic number $\kappa$, we can color the nodes of
988: $G^D$ (which are given by the set $E$ of wireless links) such no
989: two nodes which share an edge share the same color. This in turn
990: induces a link scheduling scheme, where each link in $E$ is scheduled
991: for at least a fraction $1/\kappa$ of time, and the resutling $C$ is
992: such that $C(e)\geq 1/\kappa$ for all $e\in E$. Again, the lower bound
993: can be computed by solving the LP in~(\ref{eqn:cap_reg}) with
994: $C(e)=1/\kappa$ for all $e\in E$. It is easy to see that
995: $f_1^* \leq \kappa f_2^*$.
996: 
997: Now from Theorem~\ref{thm:UMF_mincut}, we know that
998: $\Omega\left(\frac{\Psi}{\log n} \right) \leq f^* \leq \Psi$.
999: Thus, we can now also bound $\Psi$ as
1000: \[ f_2^*     \leq   \Psi \leq O\bl( f_1^* \log n\br ) \]
1001: Thus, the upper and lower bounds differ by at most a factor
1002: of $\kappa \log n$. In addition, using the algorithm
1003: in~\cite{LR88}, we can find a vector $C(e)$ and the
1004: corresponding cut $(U,U^C)$ such that the capacity of
1005: this cut, $\min_{S \subset V} \frac{\sum_{(i,j): i \in S, j\in S^c} C(i,j)}{|S||S^c|}$,
1006: is within a factor $\kappa (\log n)^2$ of $\Psi$.
1007: 
1008: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1009: 
1010: \subsection{Application}
1011: We now illustrate our results for the combinatorial interference model through an application
1012: to \emph{geometric random graphs}.
1013: The geometric random graph has been widely used to model the topology of wireless networks after the work
1014: of Gupta and Kumar~\cite{GK}. However, it has been a combinatorial object of interest for more than 60 years.
1015: We derive scaling laws for a combinatorial interference model which is more restrictive than the protocol model. Note that
1016: the lower bound hence obtained is also a lower bound for the protocol model. Specifically, the lower bound is weaker by $\log^{2.5}n$
1017: compared to the lower bound obtained in~\cite{GK}. Specifically, we show that the scaling of the lower bound is closely tied
1018: to the known combinatorial properties of geometric random graphs.
1019: 
1020: 
1021: We first define the \emph{restricted protocol interference model}.
1022: It is also parameterized by the maximum radius of
1023: transmission, $r$, and the amount of acceptable interference,
1024: $\eta$.
1025: \begin{itemize}
1026:  \item[(a)] A node $i$ can transmit to any
1027: node $j$ if the distance between $i$ and $j$, $r_{ij}$ is less than
1028: the transmission radius $r$.
1029: \item[(b)] For transmission from node $i$ to $j$ to be successful, no other node
1030: $k$ within distance $(1+\eta) r$ (where $\eta > 0$  is a constant) of
1031: node $j$ should transmit simultaneously.
1032: \end{itemize}
1033: 
1034: \vspace{0.2in}
1035: 
1036: We now state a version of the well-known Chernoff's bound
1037: for binomial random variables that we use multiple times in this paper.
1038: 
1039: 
1040: 
1041: \begin{lemma}\label{lem:chernoff}
1042: Let $X_1,\dots, X_N$ be i.i.d. binary random variables with $\Pr(X_1 = 1) = p$. Let $S_n = \sum_{k=1}^n X_k$ for
1043: $n=1,\hdots, N$.
1044: Then, for
1045: any $\delta \in (0,1)$
1046: $$ \Pr\left( | S_n - np | \geq \delta n p \right) \leq 2 \exp\left(-\frac{\delta^2np}{2}\right).$$
1047: Specifically, for $\delta = \sqrt{\frac{2L \log n}{np}}$, we have
1048: $$ \Pr\left( |S_n - np| \geq \sqrt{2L np \log n} \right) \leq \frac{2}{n^L}.$$
1049: \end{lemma}
1050: 
1051: 
1052: \vspace{.2in}
1053: 
1054: 
1055: Consider $n$ wireless nodes  distributed uniformly at random in a unit
1056: square, and the interference model given by the protocol model with transmission
1057: radius $r$. We denote such a wireless network by $G(n,r)$.
1058: It is well-known that for $G(n,r)$ to be connected with high probability,
1059: it is necessary to have $r = \Omega(\sqrt{\log n/n})$.
1060: We take $r = \Theta(\log^{3/4} n/\sqrt{n})$ and prove the following bounds on the maximum UMF, $f^*$,
1061: for the restrictive protocol model;
1062: the lower bound is only $\log^{2.5} n$ weaker than the result of Gupta and Kumar for the protocol model
1063: with $r = \Theta(\sqrt{\log n/n})$.
1064: \begin{lemma}\label{lem:gnrcomb}
1065: For $G(n,r)$, with $r = \Theta(\log^{3/4} n/\sqrt{n})$, maximum UMF is bounded as
1066: $$ \Omega\left(\frac{1}{n^{3/2} \log^{5/2} n}\right) ~\leq~ f^* ~\leq~O\left(\frac{1}{n^{3/2} \log^{3/4} n}\right).$$
1067: \end{lemma}
1068: \begin{proof}
1069: To prove the above bounds, we obtain appropriate upper and lower bounds on the
1070: quantity $\Psi$. These bounds along with  Theorem \ref{thm:UMF_mincut}
1071: imply Lemma \ref{lem:gnrcomb}. To obtain an upper bound on $\Psi$,
1072: we evaluate the cut-capacity for a specific cut-set. For
1073: the lower bound, we first, establish that a grid
1074: graph on $n$ nodes is a sub-graph of $G(n,r)$ and then use the known conductance
1075: of the grid graph.
1076: 
1077: First, consider the upper bound on $\Psi$. Specifically, consider the square, say $\cS$, of area
1078: $1/9$ (of side $1/3$) that is in the center of the unit square. Let
1079: $S$ be the set of nodes that fall inside this square.
1080:  By definition, we have
1081:  $$ \Psi ~\leq~ \Psi(S) ~=~ \sup_{C\in \cC} \frac{\sum_{i\in S, j\in S^c} C(i,j)}{|S||S^c|}.$$
1082: Therefore, it is sufficient to required obtain bound on $\Psi(S)$.
1083: 
1084: 
1085: Corresponding to
1086: node $i$, define a random variable $X_i \in \{0, 1\}$  which is $1$ is $i$
1087: is in $S$, and 0 otherwise.
1088: Since nodes
1089: are placed uniformly and independently at random in the unit area square,
1090: $X_i$ are i.i.d. binary random variable with $\Pr(X_i = 1) = 1/9$. Now,
1091: $\sum_{i=1}^n X_i$ is the number of nodes in $S$. Using Lemma~\ref{lem:chernoff} with
1092: $\delta = 0.009$, it follows that for large enough $n$,
1093: $|S| \in (0.1n,0.2n)$ (and so $|S^c| \in [0.8n, 0.9n]$) with probability at least
1094: $1-n^{-4}$. Now, consider squares $\cS^0, \cS^1$ of sides $1/3 + 2r$
1095: and $1/3 -2r$ respectively with their centers being the same as
1096: that of $\cS$. That is, $\cS^1 \subset \cS \subset \cS^0$. Let
1097: $\cA^0 = \cS^0 - \cS$ and $\cA^1 = \cS - \cS^i$.
1098: Thus, $\cA^1$ is a strip of width $r$ surrounding $\cS$ of
1099: and $\cA^0$ is a strip of width $r$ on the boundary and inside $\cS$.
1100: Since, $r = \Theta(\log^{3/4} n/\sqrt{n})$, it can be easily shown that
1101: $\cA^0$ and $\cA^1$ is $\Theta(r)$.
1102: 
1103: \iffalse It is useful
1104: to visualize that $\cA^o$ is a "strip" surrounding $\cS$ of
1105: width $r$ while $\cA^i$ is a "strip" on the boundary of $\cS$
1106: of width $r$. By Protocol model, two nodes can transmit
1107: to each other only if they are within distance $r$ of each
1108: other. Based on this, it is not hard to show that area of
1109: $\cA^o$ and $\cA^i$ is $\Theta(r)$ for large enough $n$.
1110: \fi
1111: 
1112: 
1113: Now, nodes that are
1114: in $S$ (i.e. physically inside $\cS$) can only be connected
1115: to those nodes in $S^c$ that lie in $\cA^1$. Similarly, nodes
1116: in $S^c$ that are connected to nodes $S$ must lie in $\cA^1$.
1117: Thus, nodes that can communicate across the cut $(S, S^c)$ must
1118: lie within a region of area $\Theta(r)$. For the protocol model,
1119: if a node transmits, nodes within distance $r(1+\eta)$ of the receiver
1120: must not
1121: transmit. That is, each transmission effectively silences nodes
1122: within an area of $\Theta(r^2)$.
1123: Thus, at any given time, the maximum number
1124: of simultaneous transmissions between $S$ and $S^c$ is $\Theta(1/r)$.
1125: This along with $|S|, |S^c| =\Theta(n)$ implies that
1126: $$ \Psi ~\leq~ \Psi(S) ~= ~ \frac{O(1/r)}{\Theta(n^2)} = O\left(\frac{1}{n^2 r}\right)~=~O\left(\frac{1}{n^{3/2} \log^{3/4} n}\right).$$
1127: 
1128: 
1129: 
1130: For the lower bound, we identify
1131: a grid subgraph of $G(n,r)$
1132: with $r = \Theta(\log^{3/4} n/\sqrt{n})$.
1133: Consider a grid graph $G_n$ of $\sqrt{n}\times\sqrt{n}$ nodes
1134: with each node connected to one of its four neighbors (with
1135: suitable modifications at the boundaries). The nodes of $G_n$ are placed in
1136: a uniform manner in a unit square; each node is at a distance
1137: $1/\sqrt{n}$ from its neighbors.
1138: Now consider a minimax matching between
1139: nodes of $G_n$ and $n$ randomly placed nodes in the unit square,
1140: where a minimax matching is a perfect matching between the $n$
1141: nodes of $G_n$ and the nodes of $G(n,r)$ with maximum
1142: length minimized. Leighton and Shor \cite{LS86} established
1143: that the maximum edge length in a minimax matching, say $r^*$, is
1144: $\Theta(\log^{3/4} n/\sqrt{n})$ with probability at least $1-1/n^4$.
1145: Now we identify the subgraph $G^\prime_n$ (with grid graph
1146: structure) of $G(n,r)$ as follows. $G^\prime_n$ has all $n$ nodes.
1147: Consider the minimax matching between $G_n$ and $G(n,r)$. If a node
1148: of $G(n,r)$ is connected to node number $m$ of $G_n$, then renumber
1149: it as $m$ to obtain nodes of $G^\prime_n$. Now by setting
1150: $r \geq r^* + 2/\sqrt{n}$, clearly a node $m$ and $m'$ are connected
1151: in $G^\prime_n$ if they are connected in $G_n$. Thus, we have established
1152: that $G_n \subset G^\prime_n$. Now, we will focus only on the edges
1153: of $G(n,r)$ that belong to $G^\prime_n$ and provide them with positive capacity by an appropriate
1154: communication scheme that is feasible for the restricted protocol model.
1155: For this, note that in $G(n,r)$
1156: each node is connected to at most $O(\log^{3/2}n)$ nodes with probability
1157: at least $1-1/n^4$ (using Chernoff's bound and Union bound) for large enough
1158: $n$. Hence, using a simple TDMA scheme based on vertex coloring of $G(n,r)$,
1159: each node gets to transmit once in every $\Theta(1/\log^{3/2} n)$ time slots.
1160: This transmission can be along any outgoing edge.  Since, we are interested
1161: in providing positive capacity to only at most $4$ outgoing edges,
1162: we have established that there is a simple TDMA scheme which provides
1163: $\Theta(1/\log^{3/2} n)$ capacity to each edge of a grid subgraph of
1164: $G(n,r)$. To complete the proof, we recall that
1165: the conductance of a grid graph is $\Theta(1/\sqrt{n})$ \cite{info_gossip}.
1166: That is,
1167: $$ \Phi(G_n) = \min_{S} \frac{\sum_{i\in S, j\in S^c} {\mathbf 1}_{\{(i,j) \in E\}}}{|S||S^c|} = \Theta\left(\frac{1}{n^{3/2}}\right).$$
1168: Now, putting all the above discussion together we have the following.
1169: \begin{eqnarray}
1170: \Psi & = & \sup_{C\in\cC} \min_{S\subset V} \frac{\sum_{i \in S, j \in S^c} C(i,j)}{|S||S^c|} \nonumber \\
1171:      & \geq & \Phi(G_n) \Theta\left(\frac{1}{\log^{3/2} n}\right) \nonumber \\
1172:      & = & \Omega\left(\frac{1}{n^{3/2} \log^{3/2} n}\right).
1173: \end{eqnarray}
1174: In summary, upper and lower bound on $\Psi$ along with Theorem \ref{thm:UMF_mincut}
1175: implies the Lemma \ref{lem:gnrcomb}.
1176: \end{proof}
1177: \vspace{.1in} Now, we discuss briefly delay. In~\cite{EMPS_journal}, delay
1178: was defined as the average number of hops per packet, and the
1179: packet size was assumed to scale to an arbitrarily small value. For any
1180: communication scheme feasible for the protocol model with maximum transmission radius
1181: $r = \Theta(\log^{3/4} n/\sqrt{n})$, the maximum number of transmissions per
1182: unit time is upper bounded as $O(n/\log^{3/2} n)$. Using this and
1183: Theorem \ref{thm4} we obtain the following result immediately.
1184: 
1185: \begin{lemma}\label{lem:delay}
1186: The delay $D(n)$ for any scheme achieving
1187: $f^* = \Omega\left(\frac{1}{n^{3/2} \log^{5/2} n}\right)$ is bounded
1188: above as
1189: $$D(n)  =  O\left(\sqrt{n}\log n\right).$$
1190: \end{lemma}
1191: 
1192: \vspace{.2in}
1193: 
1194: 
1195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1196: 
1197: \section{Gaussian Fading Channel Model}\label{sec3}
1198: 
1199: In the previous section, we assumed that the wireless
1200: network was defined by two graphs $G$ and $G^D$.
1201: We extended the results of Leighton and Rao
1202: to wireless networks modeled by a combinatorial interference model;
1203: this mainly exploited the fact that all possible transmission schemes could
1204: be described in terms of routing over a set of capacitated graphs, where
1205: the set of edge capacity vectors belonged to the convex hull of a finite
1206: number of vectors. Thus, in this sense, the inherent \emph{discrete}
1207: nature of the model worked to our advantage.
1208: 
1209: While the combinatorial interference model
1210: can allow for arbitrary scheduling and routing
1211: schemes, it does not model all the degrees of freedom
1212: in a wireless network. Specifically, the results are
1213: not {\em information theoretic}.
1214: In this section, we provide an information
1215: theoretic characterization of PMF in a wireless
1216: network with Gaussian fading channels.
1217: The techniques for the combinatorial model can be easily extended to obtain a feasible scheme and
1218: a lower bound on the maximum PMF $f^*$. However, for information theoretic upper
1219: bounds we have to work harder, especially to obtain a bound that relates to the lower bound
1220: and allows us to quantify the gap.
1221: 
1222: Our key contribution is in quantifying the suboptimality of the UMF/PMF for
1223: a \emph{simple feasible} scheme and an upper bound on the UMF/PMF
1224: for an arbitrary network topology, in terms of a simple
1225: graph property. The bound is general when channel side information (CSI) is assumed to be
1226: available only at the receiver.
1227: For AWGN channels, we quantify only for UMF, and when the SNR is low enough.   To the best
1228: of our knowledge, this is the first such result which guarantees that a feasible scheme
1229: achieves rates within a certain factor of an outer bound for an arbitrary graph. We also
1230: illustrate these results through applications.
1231: The results hence obtained are interesting in their own right.
1232: 
1233: 
1234: 
1235: 
1236: Our main approach is as follows. We construct
1237: two directed capacitated graphs $G^U$ and $G^L$ for
1238: the given wireless network. The graph $G^U$ is such that
1239: the capacity (defined appropriately later) of each
1240: cut in $G^U$ upper bounds the corresponding cut-capacity
1241: in the wireless network. The graph $G^L$ is such that
1242: there exists a communication scheme
1243: that simultaneously achieves the capacity of each edge in
1244: $G^L$, and the ratio of capacity of each cut in $G^U$
1245: and $G^L$ is bounded above by a quantifiable term.
1246: This leads to an approximate characterization of
1247: PMF in an arbitrary wireless network with Gaussian
1248: fading channels. Moreover, the feasible scheme that induces the capacities in $G^L$
1249: supports PMF which is within a quantifiable factor of the optimal.
1250: 
1251: 
1252: 
1253: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1254: 
1255: \subsection{Channel Model}
1256: 
1257:  This is similar to the model in, for example,~\cite{JVK}. We have
1258: $V=\{1,\dots, n\}$ wireless nodes with transceiver capabilities
1259: located arbitrarily in a plane. Node transmissions happen at
1260: discrete times, $t \in \Z_+$. Let $X_i(t)$ be the signal transmitted
1261: by node $i$ at time $t \in \Z_+$. We assume that each node has a
1262: power constraint\footnote{For notational simplicity we assume that
1263: each node has the same power constraint. The general case, where
1264: each node has different maximum average power can be handled using
1265: identical techniques.} such that
1266: \mbox{$\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{t=1}^N
1267: |X_i^2(t)| \leq P$.}  Then $Y_i(t)$, the signal received by node $i$
1268: at time $t$, is given by
1269: \begin{equation} \label{eqn:gauss_ch}
1270: \textstyle Y_i(t) = \sum_{k \neq i} H_{ik} X_k(t) + Z_i(t),
1271: \end{equation}
1272: where $Z_i(t)$ denotes a complex zero mean white Gaussian noise
1273: process with independent real and imaginary parts with variance 1/2
1274: such that $Z_i(t)$ are i.i.d. across all $i$. Let $r_{ij}$ denote the
1275: distance between nodes $i$ and $j$. Let $H_{ik}(t)$ be such that
1276: $$ H_{ik}(t) = \sqrt{g(r_{ik})} \hat{H}_{ik}(t), $$
1277: where $\hat{H}_{ik}(t)$ is a stationary and ergodic zero mean
1278: complex Gaussian process with independent real and imaginary parts
1279: (with variance 1/2). It models channel fluctuations due to frequency
1280: flat fading. Also, $g(\cdot)$ is a monotonically decreasing function
1281: that models path loss with $g(x) \leq 1$ for all $ x \geq 0$. We
1282: assume also that the $\hat{H}_{ik}(t)$'s are independent.
1283: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1284: 
1285: \subsection{Graph Definitions}
1286: 
1287: Consider the following two graphs induced by a wireless network of
1288: $n$ nodes:
1289: \begin{itemize}
1290: \item[(1)]~$K_n$ is the fully
1291: connected graph with node set $V$;
1292: \item[(2)]
1293: $G_r$ is the graph where
1294: each node $i \in V$ is connected to all nodes that are within a
1295: distance $r$ of $i$. Let $E_r$ denote the edge set of $G_r$.
1296: Let $\Delta(r)$ be the maximum vertex degree of $G_r$. Finally, define
1297: $$ r^* = \min\{r : ~G_r ~\text{~is connected}\}. $$
1298: \end{itemize}
1299: 
1300: \vspace{0.1in}
1301: 
1302: 
1303: 
1304: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1305: \subsection{Preliminaries}
1306: In the analysis in this section, we utilize the following two simple lemmas.
1307: 
1308: \begin{lemma}\label{lem:ineq1}
1309: Given $x_i \in (0,1), 1\leq i\leq N$,
1310: $$ \textstyle \sum_{i=1}^N \log (1+\sqrt{x_i}) \leq \sqrt{2N} \sqrt{\sum_{i=1}^N \log(1+x_i)}.$$
1311: \end{lemma}
1312: \begin{proof}
1313: For any $x \in (0,1)$, $x/2 \leq \log (1+x) \leq x$, so
1314: \begin{eqnarray}
1315: \textstyle \sum_{i=1}^N \log (1+\sqrt{x_i}) & \leq & \textstyle \sum_{i=1}^N \sqrt{x_i} \leq  \sqrt{N} \sqrt{\sum_{i=1}^N x_i}  \label{eq:leq2x}\\
1316: & \leq & \textstyle \sqrt{2N} \sqrt{\sum_{i=1}^N \log(1+x_i)},\nonumber
1317: \end{eqnarray}
1318: where (\ref{eq:leq2x}) follows from Cauchy-Schwarz inequality.
1319: \end{proof}
1320: 
1321: \vspace{0.2in}
1322: 
1323: \begin{lemma}
1324: \label{lem:ineq2}
1325: For any $x \geq 0$, $\alpha \in (0,1)$,
1326: $\frac{1}{\alpha} \log (1 + \alpha x) \geq \log (1+x)$.
1327: \end{lemma}
1328: \begin{proof}
1329:     Define $f(x) = \frac{1}{\alpha} \log (1+ \alpha x) - \log (1+x)$. Note that $f'(x) \geq 0$ for $x \geq 0$ and $f(0) = 0$.
1330: \end{proof}
1331: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1332: \subsection{Results}
1333: We obtain bounds on the maximum PMF for three different cases:
1334: \begin{itemize}
1335: \item[(1)] fading channel with AWGN, and channel side information (CSI) available only
1336: at the receiver,
1337: \item[(2)] deterministic (no fading) additive white Gaussian noise (AWGN) channel,
1338:  and
1339: \item[(3)] fading channel with AWGN, and CSI available at both the transmitter and the receiver.
1340: \end{itemize}
1341: The exact bounds for the above cases are different, but the analysis and bounding techniques are similar.
1342: 
1343: \vspace{0.3in}
1344: 
1345: \subsubsection{Random Fading with Rx-only CSI}
1346: We first obtain bounds on the PMF for Gaussian channels with random fading under the assumption
1347: that CSI is available at the receiver, but not the transmitter. We then relate the bounds for PMF, and show
1348: that the gap can be quantified well, and under very general assumptions. We note that this is the case for which we can obtain the
1349: strongest results.
1350: \vspace{0.1in}
1351: \begin{theorem}
1352: \label{thm:coh_bounds}
1353: With channel state information (CSI) only at receivers, $f^*_\pi$ is
1354: bounded as follows:
1355: \[\begin{aligned}
1356: &f^*_\pi \leq  \min_{S\subset V} \frac{\sum_{i\in S, j\in S^c}
1357: \E ( \log(1+ P |H_{ji}|^2 )) }{\pi(S)\pi(S^c)}, \\
1358: &f^*_\pi = \Omega\left(\sup_{r\geq r^*, \ \eta\geq
1359: 0}\left[\frac{1}{1+\Delta(r)\Delta(r(1+\eta))}\right]  \times
1360: \left[\min_{S\subset V}\frac{\sum_{i\in S, j\in S^c} {\bf 1}_{(i,j)\in
1361: E_r}\E\log\left( 1 + \frac{P|H_{ji}|^2}{1+nPg(r(1+\eta))}\right ) }{\log
1362: p_\pi \pi(S)\pi(S^c)} \right] \right).
1363: \end{aligned}
1364: \]
1365: \end{theorem}
1366: \vspace{.1in}
1367: 
1368: Theorem \ref{thm:coh_bounds} provides bounds on $f^*$ which relates to the ``cut capacity" of appropriate capacitated graphs. Specifically, \emph{we can compute the information theoretic upper bound (for any PMF) in polynomial time using flow arguments, and by solving an LP as detailed in Sec.~\ref{subsec:info_comp}}.
1369: However, it is not clear how {\em tight} these bounds are. We now quantify the \emph{gap} between
1370: the upper and lower bounds.
1371: 
1372: 
1373: \begin{corollary}
1374: \label{cor:fading}
1375: For any $r\geq r^*$, denote $\delta(r) = \max_{i}\sum_{j: r_{ij} \geq r} Pg(r_{ij})$. Then,
1376: $$
1377: \Omega\left(\frac{\Upsilon }{(1+\Delta^2(r)) (1+\delta(r)) \log p_\pi} \right) \leq f_\pi^* \leq (1+\gamma(r))\Upsilon , $$
1378: where
1379: $$\Upsilon = \min_{S\subset V}\frac{\sum_{i\in S, j\in S^c:  r_{ij} \leq r}
1380: \E \log(1+P |H_{ij}|^2)}{\pi(S)\pi(S^C)}, $$
1381: and
1382: $$ \gamma(r) = \max_{S\subset V: \pi(S), \pi(S^C)>0} \frac{\sum_{i\in S, j\in S^c:  r_{ij} > r} \E \log(1+P |H_{ij}|^2)}
1383: {\sum_{i\in S, j\in S^c:  r_{ij} \leq r} \E \log(1+P |H_{ij}|^2)} .$$
1384: \end{corollary}
1385: 
1386: 
1387: 
1388: \vspace{0.2in}
1389: 
1390: Note that both $\gamma(r)$ and $\delta(r)$ are decreasing functions of $r$, while $\Delta(r)$ is an increasing function
1391: of $r$. Also, since power typically decays as $1/r^a$ for $2\leq a\leq 6$, while  for uniformly distributed networks
1392: $\Delta(r)$ grows only linearly with $r$, the decay of $\gamma(r)$ and $\delta(r)$ is much faster than the growth of $\Delta(r)$.
1393: Hence, for $r$ large enough the \emph{gap} is dominated by the term $\log p_\pi(1+\Delta(r)^2)$.
1394: Specifically, assume that there exists an $\epsilon > 0$ such that the graph $\hGe =
1395: (V, \hEe)$ is connected, where $\hEe = \{(i,j) : \E \lf[\log \lf(1+
1396: P |H_{ij}|^2\rf)\rf] \geq n^{-\epsilon/2}\}$. Then the above bound for UMF reduces to~\cite{ISIT_paper}
1397: $$
1398: \Omega\left(\frac{\mbox{\sf min-cut}_R}{\Delta^2(r_\epsilon) \log n} \right) = f^* = O(\mbox{\sf min-cut}_R),
1399: $$
1400: where $\mbox{\sf min-cut}_R = \min_{S\subset V}\frac{\sum_{i\in S, j\in S^c:  r_{ij} \leq r_\epsilon}
1401: \E \log(1+P |H_{ij}|^2)}{|S||S^C|}$, and $r_\epsilon$ is such that $\delta(r_\epsilon)\leq \frac{1}{n^{1+\epsilon}}$.
1402: 
1403: 
1404: \vspace{0.2in}
1405: 
1406: \begin{proof}[Theorem \ref{thm:coh_bounds}]
1407: We first prove the upper bound.
1408: Following the steps in the proof of Theorem~2.1 in~\cite{JVK} and using $\left(1+\sum_{i=1}^n \alpha_i\right) \leq \prod_{i=1}^n (1+\alpha_i)$ for $\alpha_i>0$, we obtain that for $\lambda \in \Lambda$,
1409: \begin{eqnarray}
1410: \textstyle \sum_{i \in S, j \in S^c} \lambda_{ij} & \leq & \max_{Q_S\succeq 0, (Q_S)_{ii}\leq P} \E[\log \det (I + H_S Q_S
1411: H_S^*)] \nonumber \\
1412: & \leq & \textstyle \sum_{i\in S, j\in S^c} \E \left( \log(1+ P
1413: |H_{ji}|^2) \right). \label{ezx1}
1414: \end{eqnarray}
1415: Now, for any PMF $M = M(f_\pi, \pi)$,  it must be that $\sum_{i\in S, j\in S^C}M_{ij} = f_\pi \pi(S)\pi(S^C)$.
1416: Hence, for any such PMF $M(f,\pi)\in \Lambda$, the upper bound in the Theorem holds.
1417: \vspace{.03in}
1418: 
1419: To establish the lower bound, we construct a
1420: transmission scheme for which the PMF is greater
1421: than or equal to that in the lower bound. For $r\geq r^*$, consider
1422: the graph $G_r = (V,E_r)$ on the $n$ nodes defined above. We use
1423: $\Delta(r(1+\eta))$ to denote the maximum vertex degree of the graph
1424: $G_{r(1+\eta)}$. Now, consider the following transmission scheme. A
1425: node $i$ can transmit to a node $j$ only if $r_{ij}\leq r$. Also,
1426: when a node $i$ transmits, no node within a distance $r(1+\eta)$ of
1427: the receiver can transmit. Thus, when a link $(i,j)\in E_r$ is
1428: active, at most $\Delta(r(1+\eta))$ nodes are constrained to remain
1429: silent, i.e., at most $\alpha=\Delta(r(1+\eta))\Delta(r)$ links are
1430: constrained to remain inactive. Hence, the chromatic number of the
1431: dual graph is at most $(1+\Delta(r(1+\eta))\Delta(r)$. In addition,
1432: we assume that the signal transmitted by each node has a Gaussian
1433: distribution.
1434: %For the lower bound, we will use the
1435: %time-division scheme in the proof of Theorem~\ref{thm:awgn} which
1436: %schedules every link in the network for a fraction $\alpha
1437: %=\frac{1}{1+\Delta(r)\Delta(r(1+\eta))}$ of the time.
1438: For any given
1439: link that transmits data at a particular time, we treat all other
1440: simultaneous transmissions in the network as interference. Now
1441: focus on any one link, say link $(1,2)$ between node $1$ and $2$,
1442: without loss of generality. We claim the following.
1443: \begin{lemma}\label{lemma:cx1}
1444: For the above scheme, the following rate on link $(1,2)$ is achievable:
1445: \[ \textstyle \lambda_{12} =\  \alpha^{-1} \E \log \left( 1 + \frac{P |H_{21}|^2}{1 + n Pg(r(1+\eta))} \right).\]
1446: \end{lemma}
1447: We prove Lemma \ref{lemma:cx1} later. First we explain how it implies the proof
1448: of Theorem \ref{thm:coh_bounds}.  A similar analysis holds for
1449: other links that $(1,2)$ in $E_r$. Thus, for graph $G_r$
1450: the following rate are achievable on link $(i,j)\in E_r$:
1451: $$ \alpha^{-1} \E \log \left( 1 + \frac{P |H_{ji}|^2}{1 + n Pg(r(1+\eta))} \right),$$
1452: Now given the capacitated graph $G_r$, we can
1453: use classical wireline network based routing algorithms for obtaining
1454: a product multicommodity flow that is lower bounded by the following
1455: quantity:
1456: $$ f_{LB}(r, \eta) = \Omega\left(\left[\frac{1}{1+\Delta(r)\Delta(r(1+\eta))}\right]  \times
1457: \left[\min_{S\subset V}\frac{\sum_{i\in S, j\in S^c} {\bf 1}_{(i,j)\in
1458: E_r}\E\log\left( 1 + \frac{P|H_{ji}|^2}{1+nPg(r(1+\eta))}\right ) }{\log
1459: p_\pi \pi(S)\pi(S^c)} \right] \right).$$
1460: This implies the following lower bound on $f^*_\pi$:
1461: $$ f^*_\pi \geq \sup_{r \geq r^*, \eta \geq 0} f_{LB}(r, \eta).$$
1462: This is precisely the claimed lower bound in the statement of Theorem \ref{thm:coh_bounds}
1463: and thus completing the proof.
1464: \end{proof}
1465: 
1466: \begin{proof}[Lemma \ref{lemma:cx1}]
1467: We will use the following result, that follows directly from Theorem~1 in~\cite{kashyap}.
1468: \begin{theorem}
1469: \label{thm:jamming} Consider a complex scalar channel where the
1470: output $Y$ when $X$ is transmitted is given by
1471: \[ \textstyle Y = hX + Z + S,\]
1472: where $Z$ is a complex circularly symmetric Gaussian random variable
1473: with unit variance, and $S$ satisfies $\E[S^*S]\leq \hat{P} $. Also,
1474: $h$ is zero mean and i.i.d over channel uses. If $X$ is a complex
1475: zero mean circularly symmetric Gaussian random variable with
1476: $\E[X^*X]= P $, then $ \textstyle I(X;(Y,h)) \geq  \E \log\left ( 1
1477: + \frac{P|h|^2}{1 + \hat{P}} \right )$.
1478: \end{theorem}
1479: 
1480: \vspace{0.1in}
1481: 
1482: We consider a transmission scheme where the signal transmitted over
1483: each link, when active, is a complex zero mean white circularly
1484: symmetric Gaussian with variance $P$.
1485:  Moreover, we assume that the transmissions on
1486: all links are mutually independent. Let $t_1,t_2,\hdots$ denote
1487: times at which link $(1,2)$ is scheduled. Hence, at any such time
1488: $t\in\{t_1,t_2,\hdots\}$, the received signal at node 2 is given by
1489: \[ \textstyle Y_2(t) = H_{21}(t)X_1(t) + \sum_{k\neq 1,2} H_{2k}(t) X_k(t) + Z_2(t).\]
1490: Using the mutual independence of transmissions and zero mean
1491: property along with the construction of the scheduling scheme,
1492: \[ \textstyle \E \left| \sum_{k\neq 1,2} H_{2k}(t) X_k(t) + Z_2(t)  \right |^2 \leq 1 + nPg(r(1+\eta)).\]
1493: From Theorem~\ref{thm:jamming},
1494: \begin{equation} \label{zza}
1495: \textstyle I( X_1(t); (Y_2(t), H_{21}(t))) \geq \E \log \left ( 1 + \frac{P|H_{21}|^2}{(1 +n Pg(r(1+\eta) )} \right).
1496: \end{equation}
1497: Since the channel is assumed to be i.i.d. over channel uses, a random
1498: coding argument can be used to achieve this rate with a probability
1499: of error that goes to zero as the block length goes to infinity.
1500: 
1501: Combining this with the time-sharing between different sets of links
1502: described above, since each link gets to transmit at least once in $\alpha$
1503: times slots, or at least $1/\alpha$ fraction of the time, it follows that
1504: \[ \lambda_{12}  \geq \alpha^{-1} \E \log \left( 1 + \frac{P |H_{21}|^2}{1 + n
1505: Pg(r(1+\eta))} \right).\]
1506: \end{proof}
1507: 
1508: 
1509: 
1510: \begin{proof}[Corollary \ref{cor:fading}]
1511: Consider any $S$ such
1512: that $\pi(S), \pi(S^C)>0$. Then,
1513: \begin{eqnarray}
1514: & &  \textstyle \mbox{\sf Cut}(S,S^c) = \sum_{i\in S, j\in S^c} \E \left( \log(1+ P
1515: |H_{ji}|^2) \right). ~~~~~~ \nonumber \\
1516: & & \textstyle  ~\leq (1+\gamma(r)) \sum_{i\in S, j\in S^c: r_{ij} \leq r} \E \left( \log(1+ P
1517: |H_{ji}|^2) \right), ~~~~~\label{eq:cut_bound_rxcsi}
1518: \end{eqnarray}
1519: where the second line follows from the concavity of the $\log$ function, Jensen's inequality,  $\log(1+x)\leq x$
1520: for $x>0$ and definition of $\gamma(r)$.
1521: Thus,
1522: \begin{equation}
1523: \frac{\textstyle \mbox{\sf Cut}(S,S^c)}{\pi(S)\pi(S^C)}\leq \Upsilon (1+\gamma(r)).
1524: \end{equation}
1525: The upper
1526: bound then follows from the upper bound in Theorem~\ref{thm:coh_bounds}.
1527: 
1528: Next, we consider the transmission scheme that led to the lower bound
1529: in~(\ref{zza}) with $\eta = 0$.
1530: Note that in (\ref{zza}), we used the term $n Pg(r(1+\eta))$ as a bound on the interference power. However, here we consider
1531:  the actual interference $I_{ij} = \sum_{k \in V: r_{jk} \geq r}  Pg(r_{jk})$ for a transmission from $i$ to $j$. Note that
1532:  $I_{ij}\leq \delta(r)$.
1533: Now, by Lemma \ref{lem:ineq2}, we have
1534: \begin{equation} \label{x2aa}
1535: \textstyle \E \lf[\log\lf( 1+
1536: \frac{P|H_{ji}|^2}{1 + I }\rf)\rf] \geq
1537: \frac{1}{1+\delta(r)}\textstyle \E\lf(\log(1+ {P|H_{ji}|^2})
1538: \rf).
1539: \end{equation}
1540: Using (\ref{eq:cut_bound_rxcsi}) and (\ref{x2aa}) along with the
1541: lower bound obtained via time-division scheme that led to (\ref{zza}),
1542: the lower bound in Theorem~\ref{thm:coh_bounds} gives us
1543: \begin{eqnarray}
1544: f^* & = & \Omega\lf( \min_{S\subset V} \frac{\sum_{i\in S, j\in S^c: r_{ij} \leq r} \E \left( \log(1+ P
1545: |H_{ji}|^2) \right)}{\pi(S)\pi(S^C)(1+ \Delta(r)^2) (1+\delta(r))\log p_\pi} \rf) \nonumber \\
1546: & = & \Omega\lf( \frac{\Upsilon}{(1+\Delta(r)^2) (1+\delta(r))\log p_\pi}  \rf). \label{fx0}
1547: \end{eqnarray}
1548: \end{proof}
1549: 
1550: \vspace{.2in}
1551: \subsubsection{Deterministic AWGN Channels}
1552: We now consider an AWGN channel without fading, i.e., we have $\hat{H}_{kj}=1$ w.p.
1553: $1$, $\forall k,j=1,\hdots,n$.
1554: We first obtain the following set of bounds on maximum PMF using standard arguments.
1555: \begin{theorem}
1556: \label{thm:awgn} The
1557: maximum PMF $f^*_\pi$ is bounded as
1558: follows.
1559: %\footnote{For the lower bound, we use the $\Omega$ notation in
1560: %order not to have to write out constants explicitly; we note that the
1561: %constants are independent of the graph structure and $n$~\cite{LR88}.}.
1562: \[
1563: f^*_\pi \leq \min_{S\subset V} \frac{2\sum_{i\in S, j\in S^c} \log(1+
1564: \sqrt{P g(r_{ij}}))}{\pi(S)\pi(S^c)},
1565: \]
1566: \[\begin{aligned} &f^*_\pi = \Omega\left(\sup_{r\geq r^*, \ \eta\geq 0}\left[\frac{1}{1+\Delta(r)\Delta(r(1+\eta))}\right]
1567: \times \left[\min_{S\subset V}\frac{\sum_{i\in S, j\in S^c:  r_{ij} \leq r}
1568: \log\left( 1 + \frac{Pg(r_{ij})
1569: %\Delta(r(1+\eta))
1570: }{1+nPg(r(1+\eta))}\right) }{\log p_\pi \pi(S)\pi(S^c)} \right]\right).
1571: \end{aligned}\]
1572: \end{theorem}
1573: 
1574: \vspace{0.2in}
1575: 
1576: 
1577: 
1578: 
1579: 
1580: Next, we present a Corollary of Theorem \ref{thm:awgn} which characterizes the tightness of the above bound for UMF for
1581: low signal to noise ratio (SNR).
1582: 
1583: \begin{corollary}\label{cor:lbub}
1584: Define $I(r) = \min\{ I > 0: \sum_{j: r_{ij} \geq r} Pg(r_{ij}) \leq I, ~\mbox{for all $i$}\}$;
1585: $r(\delta) = \min\{ r > 0: I(r) \leq \delta\}$ for $\delta > 0$. Then,
1586: $$
1587: \Omega\left(\frac{n}{(1+\delta)\Delta(r(\delta))(1+\Delta(r(\delta))^2)\log n }  ~{\Upsilon}^2   \right) \leq f^* \leq  2\Upsilon + O\left(\frac{\delta}{n}\right), $$
1588: where
1589: $$\Upsilon = \min_{U\subset V} \frac{\sum_{i\in U, j\in U^c:  r_{ij} \leq r(\delta)} \log(1+ \sqrt{P g(r_{ij}}))}{|U||U^C|}.$$
1590: \end{corollary}
1591: 
1592: \vspace{.1in}
1593: 
1594: 
1595: We now present the proofs of Theorem~\ref{thm:awgn} and Corollary~\ref{cor:lbub}.
1596: The main idea in the proof of Theorem~\ref{thm:awgn}
1597: is to neglect interference to upper bound achievable rates on links, and
1598: to construct a transmission scheme to induce achievable rates on the links. In particular the scheme
1599: that we construct consists of time sharing between multiple transmission schemes, each of which
1600: enables direct transmissions between nodes that are separated by at most distance $r$. Then the lower
1601: bound on $f^*$ is obtained by routing over graph $G_r$, where each edge has a capacity given by this
1602: time division scheme.
1603: 
1604: \begin{proof}[Theorem \ref{thm:awgn}]
1605: We first prove the upper bound. In order to bound the sum-rate across
1606: each given cut, we refer to the proof of the max-flow min-cut lemma
1607: in \cite{XK}, which yields for any $S \subset V$ and $\lambda \in \Lambda$,
1608: $$
1609: \textstyle \sum_{i \in S, j \in S^c} \lambda_{ij} \leq \sum_{j \in S^c} \log (1 + \E(|\tilde{X}_j|^2)),
1610: $$
1611: where $\tilde{X}_j = \sum_{i \in S} \sqrt{g(r_{ji})} \, X_i$.
1612:  We therefore deduce that
1613: \begin{eqnarray*}
1614: \lefteqn{\textstyle \sum_{i \in S, j \in S^c} \lambda_{ij}}\\
1615: &\leq & \textstyle \sum_{j \in S^c} \log
1616: [1 + \sum_{i,k \in S} \sqrt{g(r_{ji}) \, g(r_{jk})} \,
1617: |\E(X_i \overline{X_k})| ]\\
1618: & \leq & \textstyle \sum_{j \in S^c} \log [ 1 + P (\sum_{i \in S}
1619: \sqrt{g(r_{ji})})^2],
1620: \end{eqnarray*}
1621: since $|\E(X_i \overline{X_k})| \leq \sqrt{P_i P_k} \leq P$. Finally,
1622: we obtain
1623: \[\begin{aligned}
1624: \textstyle \sum_{i \in S, j \in S^c} \lambda_{ij} &\textstyle \leq \sum_{j \in S^c} 2 \log (1 + \sqrt{P} \sum_{i \in S} \sqrt{g(r_{ji})})\\
1625: & \textstyle \leq \sum_{i \in S, j \in S^c} 2 \log (1 + \sqrt{P g(r_{ji})}).
1626: \end{aligned}\]
1627: Now, for any PMF $M = M(f_\pi, \pi)$,  it must be that $\sum_{i\in S, j\in S^C}M_{ij} = f_\pi \pi(S)\pi(S^C)$.
1628: Hence, for any such PMF $M(f,\pi)\in \Lambda$, the upper bound in the Theorem holds.
1629: 
1630: To establish the lower bound, we construct a
1631: transmission scheme for which the PMF is greater
1632: than or equal to that in the lower bound. For $r\geq r^*$, consider
1633: the graph $G_r = (V,E_r)$ on the $n$ nodes defined above. We use
1634: $\Delta(r(1+\eta))$ to denote the maximum vertex degree of the graph
1635: $G_{r(1+\eta)}$. Now, consider the following transmission scheme. A
1636: node $i$ can transmit to a node $j$ only if $r_{ij}\leq r$. Also,
1637: when a node $i$ transmits, no node within a distance $r(1+\eta)$ of
1638: the receiver can transmit. Thus, when a link $(i,j)\in E_r$ is
1639: active, at most $\Delta(r(1+\eta))$ nodes are constrained to remain
1640: silent, i.e., at most $\Delta(r(1+\eta))\Delta(r)$ links are
1641: constrained to remain inactive. Hence, the chromatic number of the
1642: dual graph is at most $(1+\Delta(r(1+\eta))\Delta(r)$. In addition,
1643: we assume that the signal transmitted by each node has a Gaussian
1644: distribution. Then, subject to the maximum average power constraint,
1645: for any node pair $i, j$, such that $r_{ij} \leq r$,
1646: the following rate is achievable from $i \to j$:
1647: \begin{equation}
1648: \label{cap_lb} \lambda_{ij} \geq  \frac{\log\left( 1 +
1649: \frac{Pg(r_{ij})}{ 1 + nPg(r(1+\eta))}\right )}{1
1650: + \Delta(r)\Delta(r(1+\eta))}.
1651: \end{equation}
1652: Note that the interference is due to at most $n$ nodes and all the interfering nodes are at least a distance $r(1+\eta)$ away from the receiver. We now consider routing over the graph $G_r$, where each edge $(i,j)$ has capacity $\lambda_{ij}$. The lower bound
1653: then follows from the lower bound in Theorem~\ref{thm:PMF_LR}.
1654: \end{proof}
1655: 
1656: 
1657: 
1658: 
1659: 
1660: \vspace{0.2in}
1661: 
1662: \begin{proof}[Corollary \ref{cor:lbub}]
1663: Consider any cut defined by $(S,S^c)$. Due to the symmetry of the upper bound in Theorem~\ref{thm:awgn}, without loss of generality, assume $|S| \leq n/2$.
1664: Consider any $\delta$ such that $r(\delta)\geq r^*$. Then,
1665: \begin{eqnarray}
1666: & &  \textstyle \mbox{\sf Cut}(S,S^c) =  \sum_{i\in S, j\in S^c} \log (1+\sqrt{Pg(r_{ij})})~~~~~~ \nonumber \\
1667: & & \textstyle  ~=  \sum_{i\in S, j\in S^c :  r_{ij} \leq r(\delta)} \log \lf(1+\sqrt{Pg(r_{ij})}\rf) +
1668: \sum_{i\in S, j\in S^c :  r_{ij} > r(\delta)} \log \lf(1+\sqrt{Pg(r_{ij})}\rf)\nonumber \\
1669: & & \textstyle ~\leq  \sum_{i\in S, j\in S^c :  r_{ij} \leq r(\delta)} \log \lf(1+\sqrt{Pg(r_{ij})}\rf) + |S|\delta,~~~~~\label{eq:leq3}
1670: \end{eqnarray}
1671: where the last step follows from the definition of $r(\delta)$, and $\log(1+\sqrt{x})\leq x$ for
1672: $x\leq 1$.
1673: Hence, the upper bound in the Corollary follows from the upper bound in Theorem~\ref{thm:awgn}.
1674: Since we assume $Pg(r_{ij})\leq 1$ for all $i$ and $j$, from Lemma~\ref{lem:ineq1}, we have
1675: {\small
1676: \begin{eqnarray}
1677: & & \sum_{i\in S, j\in S^c :  r_{ij} \leq r(\delta)} \log \left[ 1+\sqrt{Pg(r_{ij})} \right ]
1678:   ~~\leq  \sqrt{2 \Delta(r(\delta)) |S| \sum_{i\in S, j\in S^c :  r_{ij} \leq r(\delta)} \log \lf(1+Pg(r_{ij})\rf)}. \label{eq:leq4}
1679: \end{eqnarray}}
1680: 
1681: 
1682: For the lower bound, consider the choice of $r = r(\delta)$ and $\eta = 0$ for the
1683: scheme described in the proof of Theorem~\ref{thm:awgn}. Then, the interference during data transmission
1684: from $i$ to $j$, $I_{ij} = \sum_{k \in V: r_{jk} \geq r(\delta)}  Pg(r_{jk})\leq \delta$.
1685: Now, Lemma~\ref{lem:ineq2}, implies that
1686: \begin{equation}
1687: \label{eqn:xyz}
1688: \log \left(1 + \frac{Pg(r_{ij})}{1+I_{ij}}\right)
1689: \geq \frac{1}{1+\delta}\log( 1 +  Pg(r_{ij})).
1690: \end{equation}
1691: 
1692: Using an appropriately modified lower bound in
1693: Theorem \ref{thm:awgn} for the choice of $r = r(\delta)$, $\eta=0$, it follows that
1694: {\small \begin{eqnarray}
1695: f^* & = &  \Omega\left[ \min_{S\subset V} \frac{\sum_{i\in S,
1696: j\in S^c :  r_{ij} \leq r(\delta)} \log (1+Pg(r_{ij}))}
1697: {(1+\delta)(1+\Delta(r(\delta))^2) \log n|S||S^c|}\right] \nonumber \\
1698: & = & \Omega\left[
1699: \frac{\Upsilon^2 n}{(1+\delta)\Delta(r(\delta))(1+\Delta(r(\delta))^2)\log n }\right], \label{x2}
1700: \end{eqnarray}}
1701: where the second step follows from~(\ref{eq:leq4}). The lower bound
1702: in Theorem~\ref{thm:awgn} then implies the lower bound in the Corollary.
1703: This completes the proof.
1704: \end{proof}
1705: 
1706: %
1707: 
1708: 
1709: \vspace{0.3in}
1710: 
1711: \subsubsection{Random Fading with CSI at both Tx and Rx}
1712: We now obtain bounds on the PMF for a Gaussian channel with random fading when CSI is available at
1713: both the transmitter and the receiver.
1714: Qualitatively, these bounds are very similar to the case of deterministic
1715: AWGN channels. The main result is as follows.
1716: \vspace{.1in}
1717: \begin{theorem}
1718: \label{thm:coh_bounds1}
1719: With CSI at both transmitters and receivers, $f^*_\pi$ is bounded as follows.
1720: \[
1721: f^*_\pi \leq \min_{S\subset V}\frac{\sum_{i\in S, j\in S^c} 2
1722: \, \E ( \log(1+ \sqrt{P} \, |H_{ji}|)}{\pi(S)\pi(S^c)}.
1723: \]
1724: The lower bound for the receiver only CSI case is a (weak) lower
1725: bound for this case as well.
1726: \end{theorem}
1727: \vspace{.2in}
1728: \begin{proof}
1729: The upper bound follows again from the proof of Theorem 2.1 in \cite{JVK}, from which we deduce that for
1730: any $\lambda \in \Lambda$,
1731: \begin{eqnarray*}
1732: \lefteqn{\textstyle \hspace{-0.2in}\sum_{i \in S, j \in S^c} \lambda_{ij} \leq \E [\max_{Q\succeq 0, Q_{ii}\leq P} \log \det (I + H_S Q_S H_S^*)]}\\
1733: & \leq & \textstyle \sum_{j \in S^c} \E [\max_{Q\succeq 0, Q_{ii}\leq P} \log (1 + h_j Q_S h_j^*)],
1734: \end{eqnarray*}
1735: where $h_j$ is the $j^{th}$ row of $H$. Since $h_j Q_S h_j^*$
1736: is maximum when $(Q_S)_{ik} \equiv P$ for all $i,k \in S$,
1737: we obtain, following the steps of the proof of Theorem~\ref{thm:awgn},
1738: \begin{eqnarray*}
1739: \textstyle \sum_{i \in S, j \in S^c} \lambda_{ij}
1740: & \leq & \textstyle \sum_{j \in S^c} \E \log (1 + P (
1741: \sum_{i \in S} |H_{ij}|)^2 ]\\
1742: & \leq & \textstyle \sum_{j \in S^c} 2 \, \E \log (1 + \sqrt{P}
1743: \sum_{i \in S} |H_{ij}| )]\\
1744: & \leq & \textstyle \sum_{i  \in S, j \in S^c}
1745: 2 \, \E (\log (1 + \sqrt{P} \, |H_{ij}|),
1746: \end{eqnarray*}
1747: so the upper bound on $f^*_\pi$ follows from Theorem \ref{thm:PMF_LR}.
1748: \end{proof}
1749: 
1750: 
1751: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1752: \subsection{Computational Methods}
1753: \label{subsec:info_comp}
1754: We discuss the implications of the bounds
1755: for the case of CSI availability at the receiver only as stated in Corollary \ref{cor:fading}.
1756: Similar implications follow for the case where CSI is available to both transmitters and
1757: receivers as well.
1758: 
1759: Corollary~\ref{cor:fading} shows that an upper bound
1760: on $f^*_\pi$ can be obtained via the maximum PMF on graph $K_n$,
1761: where each edge $(i,j)$ has a capacity  $\log(1+P |H_{ij}|^2)$, and there is no interference;
1762: specifically, $\log n$ times the PMF thus computed for $K_n$ is an upper bound on $f^*$.
1763: The lower bound is obtained via routing on $G_{r}$ with edge $(i,j)$
1764: having capacity $\frac{\log(1+P |H_{ij}|^2)}{(1+\Delta^2(r))(1+\delta)}$. Hence, the PMF
1765: on $G_{r(\delta)}$ is a lower bound on $f_\pi^*$.
1766: Both the above computations can be done by solving an LP in polynomial time. Moreover, the ratio
1767: of the bounds is quantified in~\ref{cor:fading}. We note that
1768: such an approximation ratio could be obtained easily for the
1769: combinatorial interference model using node coloring arguments.
1770: The arguments  here are more complicated, as detailed
1771: in the proof of  Corollary~\ref{cor:fading}.
1772: 
1773: 
1774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1775: 
1776: 
1777: 
1778: 
1779: \subsection{Application}
1780: 
1781: We now apply the information
1782: theoretic characterization of PMF in the previous subsection to
1783: obtain a scaling law for average UMF in a geometric random network
1784: with a fading channel, and when CSI is available at the receivers. The scaling law
1785: we obtain is along similar lines to those that exist in the literature.
1786: Similar bounds can be obtained when CSI is available both
1787: at the transmitter and the receiver or when the channels are AWGN channels.
1788: 
1789: We consider a geometric random graph model
1790: with a constant node density:~$n$ nodes are placed uniformly
1791: at random in a torus of area $n$ (and not unit area). Thus the
1792: distance between two nodes is a random variable taking values
1793: in $(0,\Theta(\sqrt{n}))$. We assume that all nodes have the
1794: same transmission power equal to $1$, i.e. $P_i = 1$ for all
1795: $1\leq i\leq n$. We state the following result characterizing
1796:  $f^*$.
1797: \begin{lemma}
1798: \label{lem:coh_expl}
1799: Consider the Gaussian channel model with random fading and CSI available only
1800: at the receivers. Let $g(r) = (1+r)^{-\alpha}$, $\alpha > 3$ and $P = 1$. Then
1801: for a geometric random graph with constant node density (described above), the average (over
1802: random position of nodes) $f^*$ is bounded as
1803: $$ \Omega\left(\frac{1}{n^{3/2} \log^{1+\alpha} n}\right) \leq \E[f^*] \leq O\left(\frac{1}{n^{3/2}}\right),$$
1804: if $\Pr(|\hat{H}_{ij}|^2 \geq \beta) \geq \gamma$ for some strictily positive constants
1805: $\beta, \gamma$ (independent of $n$) for all $1\leq i, j \leq n$. (Note that the condition is for
1806: the normalized channel gains $\hat{H}_{ij}$s and not the actual gains $H_{ij}$s.)
1807: \end{lemma}
1808: \begin{proof}
1809: We use Theorem \ref{thm:coh_bounds} to evaluate the bounds. First we
1810: obtain an upper bound by evaluating the bound of Theorem \ref{thm:coh_bounds}
1811: for a specific cut $(U, U^c)$. Then, we  evaluate lower bound by
1812: relating it to an appropriate grid-graph as in Lemma \ref{lem:gnrcomb}.
1813: 
1814: 
1815: Now, we consider the upper bound. Consider a horizontal line dividing
1816: the square of area $n$ into equal halves. Let $U$ be set of nodes
1817: that lie in bottom half, and so $U^c$ is the set of nodes that lie in the
1818: top half. From Theorem \ref{thm:coh_bounds}, we have
1819: \begin{eqnarray}
1820: f^*|U||U^c| & \leq & \sum_{i \in U, j \in U^c} \E\log (1+P|H_{ij}|^2) \nonumber \\
1821:   & \leq & \sum_{i \in U, j \in U^c} \log (1+P\E[|H_{ij}|^2]) \nonumber \\
1822:   & =  & \sum_{i \in U, j \in U^c} \log (1+Pg(r_{ij})) \nonumber \\
1823:   & \leq & \sum_{i \in U, j\in U^c} Pg(r_{ij}) \nonumber \\
1824:   & = & \sum_{i\neq j} (1+r_{ij})^{-\alpha} \bone_{\{i \in U\}}\bone_{\{j\in U^c\}}, \label{eq:xx1}
1825: \end{eqnarray}
1826: where we have used Jensen's inequality, $\log (1+x) \leq x$ for all $x \geq 0$,
1827: and the hypothesis of Lemma. Since, the nodes are thrown uniformly at random,
1828: the expectation of each term in (\ref{eq:xx1}) for a pair $(i,j)$ is the same.
1829: Using linearity of expectation, we obtain that
1830: \begin{eqnarray}
1831: \E[\sum_{i\neq j}(1+ r_{ij})^{-\alpha} \bone_{\{i \in U\}}\bone_{\{j\in U^c\}}]
1832: & = & n(n-1) \E[(1+r_{12})^{-\alpha}\bone_{\{1 \in U\}}\bone_{\{2\in U^c\}} ] \nonumber \\
1833: & \leq & O\left(n^2 \int_{1}^{\sqrt{n}}  \int_{r}^{\sqrt{n}} s^{-\alpha} \frac{s ds}{n}
1834: \frac{\sqrt{n} dr}{n}\right) \nonumber \\
1835: & = & O\left(\sqrt{n} \int_1^{\sqrt{n}} \int_{r}^{\sqrt{n}} s^{-\alpha+1} ds dr \right) \nonumber \\
1836: & = & O\left(\sqrt{n}\int_1^{\sqrt{n}} r^{-\alpha+2} dr\right) \nonumber \\
1837: & = & O(\sqrt{n}),\label{lx1}
1838: \end{eqnarray}
1839: where we used the fact that for $\alpha > 3$ the last integral is bounded
1840: above by a constant. The above evaluation can be justified as follows. First
1841: note that $\Pr(1 \in U, 2 \in U^c) = 1/4$. Given $\{1\in U, 2 \in U^c\}$,
1842: node $1$ in the bottom rectangle and node $2$ in the top rectangle
1843: are uniformly distributed. Now, consider a thin horizontal strip of
1844: width $dr$ and length $\sqrt{n}$ at distance $r$ below the
1845: horizontal line dividing the square (and inducing $U, U^c$). The node
1846: $1 \in U$ belongs to this strip with probability $2\sqrt{n} dr/n$. Now,
1847: the node $2$ is at distance at least $r$ from node $1$. Consider a ring
1848: of width $ds$, centered at node $1$'s position and of radius $s \geq r$.
1849: The area of this ring is $2\pi s ds$. The probability of node $2$ being
1850: in this ring is bounded above by $4\pi s ds/n$. When the above described
1851: condition is true, the nodes $1$ and $2$ are at distance $s$. Integrating
1852: over the appropriate ranges justifies the final outcome (\ref{lx1}).
1853: 
1854: 
1855: 
1856: 
1857: 
1858: Now, it is easy to see that under
1859: any configuration of nodes, $f^* = O(n^2)$ since $g(r) \leq 1$ for any $r \geq 0$,
1860: $P = 1$ and elementary arguments. Let
1861: event $A = \{ |U||U^c| = \Theta(n^2)\}$. Using Chernoff's bound, it is
1862: easy to see that (with appropriate selection of constants in definition of $A$)
1863: for large enough $n$, we have
1864: $$ \Pr(A) = 1-1/n^6. $$
1865: Using this estimate and bound $f^* = O(n^2)$ we obtain that
1866: \begin{eqnarray}
1867: \E[f^*] & = & \E[f^* \bone_A] + \E[f^* \bone_{A^c}] \nonumber \\
1868:         & \leq & \E[f^* \bone_A] + O\left(\frac{1}{n^4}\right) \nonumber \\
1869:         & = & \Theta\left(\frac{\E[f^* |U||U^c|\bone_{A}]}{n^2}\right) + O\left(\frac{1}{n^4}\right) \nonumber \\
1870:         & \leq & \Theta\left(\frac{\E[f^* |U||U^c|]}{n^2}\right) + O\left(\frac{1}{n^4}\right) \nonumber \\
1871:         & = & O\left(\frac{1}{n^{3/2}}\right).
1872: \end{eqnarray}
1873: 
1874: Next, we prove the lower bound. For this we construct a graph
1875: with achievable link capacities for which the average $f^*$ is
1876: lower bounded as claimed in the Lemma.
1877: Consider $r = \Theta(\log n)$. Then
1878: the corresponding $G_r$, which is the geometric random graph $G(n,r)$, is
1879: connected with high probability (at least $1-1/n^4$ by
1880: appropriate choice of constants in selection of $r$).
1881: For this choice of $r$, using the Chernoff and Union bounds it follows
1882: that with probability at least $1-1/n^4$,
1883: $$ \Delta(2r) = \Theta\left(\log^{2} n\right).$$
1884: Again, we can identify
1885: a grid graph structure as a subgraph structure of
1886: $G_r$ based on the argument used in Lemma \ref{lem:gnrcomb}.
1887: Denote the edges of this grid sub-graph structure as $\hat{E}$.
1888: We note that $\Theta(1)$ edges are incident on each
1889: of the $n$ nodes that belong to $\hat{E}$ (which is a property of
1890: the grid-graph structure). Next, we design a feasible transmission scheme
1891: for which each edge in $\hat{E}$ can support a transmission rate of
1892: $\Omega(\log^{-\alpha} n)$.
1893: 
1894: Specifically, we consider a TDMA schedule for the
1895: graph $G_r$ similar to that described in the proof of the lower bound
1896: for Theorem \ref{thm:coh_bounds}.
1897: It is easy to
1898: see that $G_{2r}$ can be vertex colored using
1899: $\Theta(\Delta(2r))$ colors.
1900: We use a randomized
1901: scheme to do  TDMA scheduling
1902: as follows: in each time-slot, each node becomes
1903: tentatively active with probability $1/\Delta(2r)$ and
1904: remains inactive otherwise. If a node becomes tentatively active
1905: and none of its neighbors in $G_{2r}$ is tentatively-active, then
1906: it will become active. Else, it becomes inactive. All
1907: active nodes transmit in the time-slot simultaneously.
1908: It is easy to see that each node transmits for
1909: $\Theta(1/\Delta(2r))$ fraction of the time on an average. The
1910: randomization here is used to facilitate the  computation
1911: of a simple bound on the average interference experienced
1912: by a node due to transmissions by  nodes that are
1913: not its neighbor.
1914: 
1915: Now under the above vertex coloring, each node
1916: gets to transmit once in $\Theta(\Delta(2r))$ time-slots on average
1917: at power $\Theta(P \Delta(2r)) = \Theta(\Delta(2r))$.
1918: We wish to concentrate on transmissions for
1919: edges that belong to $\hat{E}$, which are a subset of
1920: edges of $G_r$. For any such transmission,
1921: say from $u \to v$ with $(u,v) \in \hat{E}$, according
1922: to the above coloring of $G_{2r}$ no other node within distance $r$ of $v$ is transmits
1923: simultaneously. Also, any node
1924: that is at least distance $r$ away from $v$ can be active
1925: with probability at most $1/\Delta(2r)$. Hence, the average power
1926: corresponding to the interference received
1927: by node $v$, say $I_v$, can be bounded above as follows:
1928: \begin{eqnarray}
1929:  I_v & = & O\left(\sum_{j: r_{vj} > r} \frac{P\Delta(2r)\E[|H_{ij}|^2]}{\Delta(2r)}
1930:  \right) = O\left(\sum_{j: r_{vj} > r} g(r_{vj})\right). \label{eq:xx4}
1931: \end{eqnarray}
1932: where we used the fact that each node transmits at
1933: power $P\Delta(2r) = \Delta(2r)$ for $1/\Delta(2r)$ fraction
1934: of the time and $\E[|H_{ij}|^2] = g(r_{ij})$.
1935: By another application of Chernoff's bound and union bound, it
1936: can be shown that the number of nodes in an annulus around node $v$ with unit
1937: width and radius $R$ for $R \in \N, r \leq R \leq \sqrt{n}$, is
1938: $\Theta(R)$ with probability at least $1-1/n^5$. Then it follows that
1939: $$ \sum_{j: r_{vj} > r} g(r_{vj}) = O\left(\sum_{R=\lfloor r \rfloor}^{\sqrt{n}} g(R) R\right) = O\left(r^{-\alpha+2}\right),$$
1940: where we have used fact that $\alpha > 3$. Using this in
1941: (\ref{eq:xx4}), we obtain that with probability at least $1-1/n^{5/2}$ we
1942: have
1943: \begin{eqnarray}
1944:  I_v & = & O\left(r^{-\alpha+2}\right) = O\left(\log^{2-\alpha} n\right).
1945:  \end{eqnarray}
1946: That is, $I_v \to 0$ as $n\to\infty$ for $\alpha > 3$. Thus, by selection
1947: of large enough $n$, $I_v$ can be made as small as possible.
1948: That is, when transmission from $u\to v$ happens, the average noise
1949: received by node $v$ due to other simultaneous transmission is very
1950: small, say less than $\delta$ for some small enough $\delta >0$.
1951: 
1952: Given this, the arguments used in Lemma \ref{lemma:cx1} imply that
1953: when $u$ transmits to $v$ at power $P\Delta(2r)$ once
1954: in $\Theta(\Delta(2r))$ time-slot, considering other transmissions
1955: as noise, we obtain that the effective rate between $u \to v$ is
1956: lower bounded as
1957: $$ \lambda_{u\to v} = \Omega\left(\Delta(2r)^{-1} \E\left[\log \left( 1+ \frac{P \Delta(2r) g(r_{uv}) |\hat{H}_{uv}|^2}{1+I_v}\right ) \right]\right).$$
1958: Now, $\hat{H}_{uv}$ is independent of everything else and
1959: $$\Pr(|\hat{H}_{uv}|^2 \geq \beta) \geq \gamma,  $$
1960: for some positive constants $\beta, \gamma$ as per our hypothesis.
1961: Therefore, use of Lemma \ref{lem:ineq2} implies
1962: $$ \lambda_{u\to v} \geq \Omega\left(\Delta(2r)^{-1} \log \left (1+ \frac{P \Delta(2r) g(r_{uv})}{1+I_v}\right) \right).$$
1963: Further, $I_v \leq \delta$ for small enough $\delta$. Therefore,
1964: another use of Lemma \ref{lem:ineq2} implies that
1965:  \[\lambda_{u\to v} \geq \Omega\left(\Delta(2r)^{-1} \log \left( 1+ P \Delta(2r) g(r_{uv})\right) \right).\]
1966: Now
1967: $$ P g(r_{uv})\Delta(2r) = \Omega\left( \log^{2-\alpha} n\right),$$
1968: where we have used the fact that $r_{uv} \leq r$ and $g(\cdot)$ is
1969: monotonically decreasing.  For $x \in (0,0.5)$, $\log (1+x) \geq x/2$.
1970: Therefore, for $\alpha > 3$
1971: $$ \lambda_{u\to v} = \Omega\left(\Delta(2r)^{-1} \log^{2-\alpha} n\right).$$
1972: Since $\Delta(2r) = \Theta(\log^2 n)$,  we have established that
1973: the effective capacity of transmissions for each edge under
1974: the above described TDMA scheme is $\Omega(\log^{-\alpha} n)$.
1975: That is, each edge of $\hat{E}$ gets capacity at least
1976: $\Omega(\log^{-\alpha} n)$. Now recall that a grid graph
1977: with unit capacity has $f^*$ lower bounded
1978: as $ \Omega\left(\frac{1}{n^{3/2} \log n}\right)$.
1979: Hence, using this routing of UMF along edges of $\hat{E}$
1980: with capacity $\Omega(\log^{-\alpha} n)$ we obtain that
1981: $$ f^* = \Omega\left(\frac{1}{n^{3/2} \log^{1+\alpha} n}\right).$$
1982: By careful accounting of probability of relevant events above
1983: and Union bound of events will imply that the above stated lower
1984: bound on $f^*$ holds with probability at least $1-1/n^2$. Since
1985: $f^* \geq 0$ with probability $1$, it immediately implies the
1986: desired lower bound of Lemma:
1987: $$ \E[f^*] = \Omega\left(\frac{1}{n^{3/2} \log^{1+\alpha} n}\right).$$
1988: This completes the proof of Lemma \ref{lem:coh_expl}.
1989: 
1990: 
1991: 
1992: \end{proof}
1993: 
1994: \vspace{.1in}
1995: 
1996: 
1997: 
1998: 
1999: \vspace{0.3in}
2000: 
2001: %\subsubsection{Computational Example}
2002: 
2003: \section*{Acknowledgment}
2004: The authors would like to thank A. Jovicic and S. Tavildar for a helpful discussion on the information
2005: theoretic min-cut bound for networks with Gaussian channels.
2006: We would like to thank Ashish Goel for sharing his proof about evaluating
2007: mixing time of $G(n,r)$ -- it has influenced the proof of Lemma \ref{lem:gnrcomb}.
2008: 
2009: 
2010: \bibliographystyle{IEEEtran}
2011: \bibliography{cap}
2012: 
2013: 
2014: \end{document}
2015: 
2016: 
2017: