1: \documentclass[12pt,final]{article}
2: \usepackage{amssymb,amsmath,amsthm}
3: \usepackage{epsfig}
4: \usepackage{graphics}
5:
6: \oddsidemargin 0.0in \textwidth 6.4in \topmargin 0.0in \headheight 0.0in \headsep 0.0in \textheight 9.0in
7:
8: \newtheorem{theorem}{Theorem}[section]
9: \newtheorem{lemma}[theorem]{Lemma}
10: \newtheorem{conjecture}[theorem]{Conjecture}
11: \newtheorem{claim}[theorem]{Claim}
12: \newtheorem{definition}[theorem]{Definition}
13: \newtheorem{proposition}[theorem]{Proposition}
14: \newtheorem{corollary}[theorem]{Corollary}
15: \newtheorem{example}[theorem]{Example}
16:
17: \newcommand{\SNR}{\sf SNR}
18: \newcommand{\SIR}{\sf SIR}
19: \newcommand{\SINR}{\sf SINR}
20:
21: \begin{document}
22: \title{Spectrum Sharing between Wireless Networks}
23: \author{Leonard Grokop \quad David N.C. Tse\\
24: Department of Electrical Engineering and Computer Sciences\\
25: University of California\\
26: Berkeley, CA 94720, USA\\
27: \{lgrokop,dtse\}@eecs.berkeley.edu} \maketitle
28:
29: \begin{abstract}
30: We consider the problem of two wireless networks operating on the same (presumably unlicensed) frequency band. Pairs
31: within a given network cooperate to schedule transmissions, but between networks there is competition for spectrum. To
32: make the problem tractable, we assume transmissions are scheduled according to a random access protocol where each
33: network chooses an access probability for its users. A game between the two networks is defined. We characterize the
34: Nash Equilibrium behavior of the system. Three regimes are identified; one in which both networks simultaneously
35: schedule all transmissions; one in which the denser network schedules all transmissions and the sparser only schedules
36: a fraction; and one in which both networks schedule only a fraction of their transmissions. The regime of operation
37: depends on the pathloss exponent $\alpha$, the latter regime being desirable, but attainable only for $\alpha>4$. This
38: suggests that in certain environments, rival wireless networks may end up naturally cooperating. To substantiate our
39: analytical results, we simulate a system where networks iteratively optimize their access probabilities in a greedy
40: manner. We also discuss a distributed scheduling protocol that employs carrier sensing, and demonstrate via
41: simulations, that again a near cooperative equilibrium exists for sufficiently large $\alpha$.
42: \end{abstract}
43:
44: \section{Introduction}
45: The recent proliferation of networks operating on unlicensed bands, most notably 802.11 and Bluetooth, has stimulated
46: research into the study of how different systems competing for the same spectrum interact. Communication on unlicensed
47: spectrum is desirable essentially because it is free, but users are subject to random interference generated by the
48: transmissions of other users. Most research to date has assumed devices have no natural incentive to cooperate with one
49: another. For instance, a wireless router in one apartment is not concerned about the interference it generates in a
50: neighboring apartment. Following from this assumption, various game-theoretic formulations have been used to model the
51: interplay between neighboring systems \cite{spectrum1}, \cite{spectrum2}, \cite{spectrum3}, \cite{spectrum4},
52: \cite{spectrum5}. An important conclusion stemming from this body of work is that for single-stage games the Nash
53: Equilibria (N.E.) are typically unfavorable, resulting in inefficient allocations of resources to users. A
54: quintessential example is the following. Consider a system where a pair of competing links is subjected to white-noise
55: and all cross-gains are frequency-flat. Suppose the transmitters wish to select a one-time power allocation across
56: frequency subject to a constraint on the total power expended (this problem is studied in \cite{gg1}, \cite{gg2},
57: \cite{gg3} where it is referred to as the {\it Gaussian Interference Game}). It is straightforward to reason (via a
58: waterfilling argument) that the selection by both users of frequency-flat power allocations, each occupying the entire
59: band, constitutes a N.E.. This {\it full spread} power allocation can be extremely inefficient. Consider a symmetric
60: system where the cross-gains and direct-gains are equal. At high $\SNR$ each link achieves a throughput of only 1
61: b/s/Hz, instead of $\frac{1}{2}\log_2(1+\SNR)$ b/s/Hz, which would be obtained if the links cooperated by occupying
62: orthogonal halves of the spectrum. At an $\SNR$ of 30 dB, the throughput ratio between cooperative behavior and this
63: full spread N.E. behavior, referred to as the {\it price of anarchy}, is about 5. This example highlights an important
64: point in relation to single-stage games between competing wireless links: users typically have an incentive to occupy
65: all of the available resources.
66:
67: In this work a different approach is taken. Rather than assuming total anarchy, that is, competition between {\it all}
68: wireless links, we instead assume competition only between wireless links belonging to {\it different networks}.
69: Wireless links belonging to the same network are assumed to {\it cooperate}. In short, we assume competition on the
70: network level, not on the link level. In a practical setting this may represent the fact that neighboring wireless
71: systems are produced by the same manufacturer, or are administered by the same network operator. Alternatively one may
72: view the competition as being between coalitions of users \cite{coal1}.
73:
74: To make the problem analytically tractable but still retain its underlying mechanics, we assume each network operates
75: under a random-access protocol, where users from a given network access the channel independently but with the same
76: probability. Analysis of random access protocols provides intuition for the behavior of systems operating under more
77: complex protocols, as the access probability can broadly be interpreted as the average degrees of freedom each user
78: occupies. For the case of competition on the link level, game-theoretic research of random-access protocols such as
79: ALOHA have been conducted in \cite{ra1} and \cite{ra2}. In our model each network has a different density of nodes and
80: chooses its access probability to maximize average throughput per user. Note this access protocol is essentially
81: identical to one in which users select a random fraction of the spectrum on which to communicate. Thus an access
82: probability of one corresponds to a full spread power allocation.
83:
84: We first assume all links in the system have the same transmission range and afterwards show that the results are only
85: trivially modified if each link is assumed to have an i.i.d. random transmission range. We characterize the N.E. of
86: this system for a fixed-rate model, where all users transmit at the same data rate. We show that unlike the case of
87: competing links, a N.E. always exists and is {\it unique}. Furthermore for a large range of typical parameter values,
88: the N.E. is not full spread ---nodes in at least one network occupy only a fraction of the bandwidth. We also identify
89: two modes, delineated by the pathloss exponent $\alpha$. For $\alpha>4$, the N.E. behavior is distinctly different than
90: for $\alpha<4$ and possesses pseudo-cooperative properties. Following this we show that the picture for the
91: variable-rate model, in which users individually tailor their transmission rates to match the instantaneous channel
92: capacity, remains unchanged. Before concluding we present simulation results for the behavior of the system when the
93: networks employ a greedy algorithm to optimize their throughput, operating under both a random access protocol, and a
94: carrier sensing protocol.
95:
96: In section II we formulate the system model explicitly. In section III.A we introduce the random access protocol and
97: analyze its N.E. behavior in the fixed-rate model. In section III.B we analyze the variable-rate model. In section IV
98: we extend our results to cover the case of variable transmission ranges. Section V presents simulation results and the
99: carrier sensing protocol. Section VI summarizes and suggests extensions. Section VII contains proofs of the main
100: theorems presented.
101:
102: \section{Problem Setup}
103: Consider two wireless networks consisting of $n\lambda_1$ and $n\lambda_2$ tx-rx pairs, respectively. Without loss of
104: generality we will assume $\lambda_1 \le \lambda_2$. The transmitting nodes are uniformly distributed at random in an
105: area of size $n$. To avoid boundary effects suppose this area is the surface of a sphere. Thus $\lambda_i$ is the
106: density of transmitters (or receivers) in network $i$. For each transmitter, the corresponding receiver is initially
107: assumed to be located at a fixed range of $d$ meters with uniform random bearing. Time is slotted and all users are
108: assumed to be time synchronized.
109:
110: Both networks operate on the same band of (presumably unlicensed) spectrum and at each time slot a subset of tx-rx
111: pairs are simultaneously scheduled in each network. When scheduled a tx-rx pair uses all of the spectrum. It is
112: generally desirable to schedule neighboring tx-rx pairs in different time slots. This scheduling model is a form of
113: TDM, but is more or less analogous to an FDM model where each tx-rx pair is allocated a subset of the spectrum
114: (typically overlapping in some way with other tx-rx pairs in the network).
115:
116: Transmitting nodes are full buffer in that they always have data to send. Transmissions are assumed to use Gaussian
117: codebooks and interference from other nodes is treated as noise. Initially we analyse the model where all transmissions
118: in network $i$ occur at a common rate of $\log(1+\beta_i)$. We refer to $\beta$ as the {\it target} $\SINR$. Thus a
119: transmission in network $i$ is successful iff ${\SINR} > \beta_i$. Later we explore the model where transmission rates
120: are individually tailored to match the instantaneous capacities of the channels. The signal power attenuates according
121: to a power law with {\it pathloss exponent} $\alpha > 2$. We assume a high-$\SNR$ or {\it interference limited}
122: scenario where the thermal noise is insignificant relative to the received power of interfering nodes and thus refer to
123: the $\SIR$ as the $\SIR$. For a given realization of the node locations the time-averaged throughput achieved by the
124: $j$th tx-rx pair in network $i$ is then
125: \begin{equation*}
126: {\overline R}_j = f_j\mathbb P({\SIR_j(t)}>\beta_i)\log(1+\beta_i)
127: \end{equation*}
128:
129: \noindent per complex d.o.f., where $f_j$ is the fraction of time the $j$th tx-rx pair is scheduled. The average
130: (represented by the bar above the $R$) is essentially taken over the distribution of the interference as at different
131: times different subsets of transmitters are scheduled.
132:
133: As for $\alpha > 2$ the bulk of the interference is generated by the strongest interferer, to make the problem
134: tractable, we compute the $\SIR$ as the receive power of the desired signal divided by the receive power of the {\it
135: nearest} interferers signal. We refer to this as the {\it Dominant Interferer} assumption. Denote the range of the
136: nearest interferer to the $j$th receiver at time $t$ by $r_j(t)$. Then
137: \begin{equation*}
138: {\SIR}_j(t) = \frac{d^{-\alpha}}{r_j^{-\alpha}(t)}
139: \end{equation*}
140:
141: \noindent The metric of interest to each network is its expected time-averaged rate per user,
142: \begin{equation*}
143: {\mathbb E}{\overline R} = {\mathbb E}_g \left[ f_j\mathbb P({\SIR_j(t)}>\beta_i)\log(1+\beta_i) \right]
144: \end{equation*}
145:
146: \noindent The subscript $g$ indicates this expectation is taken over the geographic distribution of the nodes. As the
147: setup is statistically symmetric, this metric is equivalent to the expected {\it sum rate} of the system, divided by
148: $n\lambda_i$, in the limit $n\rightarrow \infty$.
149:
150: \section{Random Access protocol}
151: \subsection{Fixed-Rate model}
152: Suppose each network uses the following random access protocol. At each time slot each link is scheduled i.i.d. with
153: probability $p_i$. The packet size is $\log(1+\beta_i)$ for all communications in network $i$. The variables $\beta_i$
154: are optimized over.
155:
156: Let us first compute the optimal access probability for the case of a single network operating in isolation on a
157: licensed band, as a function of the node density and the transmission range. This problem has recently been studied
158: independently in \cite{jindal1}-\cite{jindal4} with equivalent results derived. In \cite{baccelli} similar results are
159: derived for the case where the $\SIR$ is computed based on all interferers, not just the nearest.
160:
161: Let the r.v. $N_j(x)$ denote the number of interfering transmitters within range $x$ of the $j$th receiver.
162: \begin{align*}
163: {\mathbb E}{\overline R} &= {\mathbb E}_g \left[ f_j\mathbb P(r_j(t)>\beta_i^{1/\alpha}d)\log(1+\beta_I) \right] \\
164: &= p_i\sum_{k=1}^{n\lambda_i-1} {\mathbb P}(N_j(\beta^{1/\alpha}d) = k)(1-p_i)^k\log(1+\beta_i) \\
165: &= p_i\sum_{k=1}^{n\lambda_i-1} \left(\frac{\pi\beta_i^{2/\alpha}d^2}{n}\right)^k
166: \left(1-\frac{\pi\beta_i^{2/\alpha}d^2}{n}\right)^{n\lambda_i-k-1} \\
167: &\quad\quad\quad\quad\quad \times \binom{n\lambda_i-1}{k}(1-p_i)^k\log(1+\beta_i)\\
168: &= p_i\left(1-p_i\frac{\pi\beta_i^{2/\alpha}d^2}{n}\right)^{n\lambda_i-1}\log(1+\beta_i) \\
169: &\rightarrow p_i \log(1+\beta_i) e^{-\pi\lambda_id^2p_i\beta_i^{2/\alpha}}
170: \end{align*}
171:
172: \noindent in the limit $n\rightarrow \infty$.
173:
174: In order to obtain better insight into the problem at hand, a change of variables is required. We refer to the set of
175: all points within the transmission range as the {\it transmission disc}. The quantity $\pi\lambda_i d^2$ is the {\it
176: average number of nodes (tx or rx) per transmission disc}. We often refer to it simply as the {\it number of nodes per
177: disc} and represent it by the symbol
178: \begin{equation*}
179: N_i \triangleq \pi\lambda_i d^2.
180: \end{equation*}
181:
182: Assume $N_i$ is larger than a certain threshold (we make this precise later). Maximizing over the access probability
183: yields
184: \begin{equation}\label{eqn:one_network_thpt}
185: {\mathbb E}{\overline R} \rightarrow \frac{\log(1+\beta_i)}{N_i \beta_i^{2/\alpha}} e^{-1}
186: \end{equation}
187:
188: \noindent with the optimal access probability being
189: \begin{equation}\label{eqn:p_star} p_i^* =
190: \frac{1}{N_i\beta_i^{2/\alpha}}.
191: \end{equation}
192:
193: \noindent One can further optimize over the target $\SIR$ so that $\beta_i$ is replaced by $\beta_i^*$ in the above two
194: equations. Inspection of equation (\ref{eqn:one_network_thpt}) reveals that the optimal target $\SIR$ is a function of
195: $\alpha$ alone. So if we define the quantity
196: \begin{equation*}
197: \Lambda_i \triangleq N_i p_i,
198: \end{equation*}
199:
200: \noindent $\Lambda_i^*$ will be a constant, independent of $N_i$. The quantity $\Lambda_i$ represents the {\it average
201: number of (simultaneous) transmissions per transmission disc}. We sometimes refer to it simply as the {\it transmit
202: density}. Whereas the domain of $p_i$ is $[0,1]$, the domain of $\Lambda_i$ is $[0,N_i]$. Thus we see that for $N_i$
203: sufficiently large, the access probability should be set such that an optimal number of transmissions per disc is
204: achieved. What is this optimal number? What is the optimal target $\SIR$?
205:
206: For the purposes of optimizing equation (\ref{eqn:one_network_thpt}), define the function $\Lambda^*(\alpha)$ as the
207: unique solution of the following equation
208: \begin{equation}\label{eqn:Lambda_star}
209: \frac{\alpha}{2} = \left( 1+ {\Lambda^*}^{\alpha/2} \right) \log\left(1+\frac{1}{{\Lambda^*}^{\alpha/2}}\right).
210: \end{equation}
211:
212: \noindent A plot of $\Lambda^*(\alpha)$ is given in figure \ref{fig:Lambda_star}. So as to avoid confusion, note that
213: the symbol $\Lambda^*$ represents a pre-defined function, not necessarily the same as the symbol $\Lambda_i^*$, which
214: is a variable. As equation (\ref{eqn:one_network_thpt}) is smooth and continuous with a unique maxima, by setting it's
215: derivative to zero we find that the optimal target $\SIR$ is $\beta^* = {\Lambda^*}^{-\alpha/2}$ and the optimal number
216: of transmissions per disc is $\Lambda_i^*=\Lambda^*$, when $N_i$ is larger than a certain threshold.
217:
218: \begin{figure}
219: \centering
220: \includegraphics[width=420pt]{Lambda_star.eps}
221: \caption{Optimal average number of transmissions per transmission disc as a function of the pathloss
222: exponent.}\label{fig:Lambda_star}
223: \end{figure}
224:
225: When $N_i$ is smaller than this threshold, there aren't enough tx-rx pairs to reach the optimal number of transmissions
226: per disc, even when all of them are simultaneously scheduled. In this case the solution lies on the boundary with
227: $p_i^*=1$. This corresponds to the scenario where the transmission range is short relative to the node density such
228: that tx-rx pairs function as if in isolation. It is intuitive that in this case all transmissions in the network will
229: be simultaneously scheduled. Our discussion is summarized in the following theorem.
230:
231: \begin{theorem}\label{thm:op_access_prob} (Optimal Access Probability)
232: For a single network operating in isolation under the random access protocol, when $N_i > \Lambda^*$ the optimal access
233: probability is $p_i^* = \frac{\Lambda^*}{N_i}$, where $\Lambda^*$ is given by the unique solution of equation
234: (\ref{eqn:Lambda_star}). The optimal target $\SIR$ is $\beta_i^* = {\Lambda^*}^{-\alpha/2}$.
235:
236: When $N_i \le \Lambda^*$, $p_i^*=1$ and $\beta_i^*$ is given by the unique solution to
237: \begin{equation*}
238: \frac{\alpha}{2N_i {\beta_i^*}^{2/\alpha}} = \left( 1+ \frac{1}{\beta_i^*} \right) \log(1+\beta_i^*).
239: \end{equation*}
240: \end{theorem}
241:
242: \noindent The region satisfying $N_i > \Lambda^*$ is referred to as the {\it partial} reuse regime. The complement
243: region is referred to as the {\it full reuse} regime. Note the optimal access probability of the above theorem is
244: equivalent to the results of section IV.B in \cite{jindal1}, and those discussed under the title {\it ``Maximum
245: Achievable Spatial Throughput and TC''} on page 4137 of \cite{jindal3}.
246:
247: Now we perform the same computation for the case where both networks operate on the same unlicensed band. In this case
248: there is both {\it intra-network} and {\it inter-network} interference. It is straightforward to extend the above
249: analysis to show that for network $i$
250: \begin{equation*}
251: {\mathbb E}{\overline R}_i \rightarrow \frac{\Lambda_i}{N_i} \log(1+\beta_i)
252: e^{-(\Lambda_1+\Lambda_2)\beta_i^{2/\alpha}}
253: \end{equation*}
254:
255: \noindent in the limit $n \rightarrow \infty$. For a given $\Lambda_2$ the first network can optimize $\Lambda_1$ and
256: $\beta_1$, and vice-versa. That is each network can iteratively adjust its access probability and target $\SIR$ in
257: response to the other networks. In this sense a game can be defined between the two networks. A strategy for network
258: $i$ is a choice of $\Lambda_i\in [0,N_i]$ and $\beta_i>0$. Its payoff function (also referred to as {\it utility
259: function}) is the limiting form of ${\mathbb E}{\overline R}_i$ times $N_i$,
260: \begin{align}
261: \label{eqn:utility_fns1}
262: U_1\left((\Lambda_1,\beta_1),(\Lambda_2,\beta_2)\right) &= \Lambda_1 \log(1+\beta_1) e^{-(\Lambda_1+\Lambda_2)\beta_1^{2/\alpha}} \\
263: \label{eqn:utility_fns2} U_2\left((\Lambda_1,\beta_1),(\Lambda_2,\beta_2)\right) &= \Lambda_2 \log(1+\beta_2)
264: e^{-(\Lambda_1+\Lambda_2)\beta_2^{2/\alpha}}.
265: \end{align}
266:
267: \noindent Here we have scaled the throughput by $N_i$ to emphasize the simple form of the payoff functions. At first
268: glance this setup seems desirable but there is a redundancy in the way the strategy space has been defined. The problem
269: is that the variable $\beta_i$ only appears in $U_i$ and thus should be optimized over separately rather than being
270: included as part of the strategy. This leads to the following game setup.
271:
272: \begin{definition}\label{def:NE} (Random Access Game)
273: A strategy for network $i$ in the Random Access Game is a choice of $\Lambda_i \in [0,N_i]$. The payoff functions are
274: \begin{align*}
275: U_1(\Lambda_1,\Lambda_2) &= \max_{\beta_1 > 0} \Lambda_1 \log(1+\beta_1) e^{-(\Lambda_1+\Lambda_2)\beta_1^{2/\alpha}} \\
276: U_2(\Lambda_1,\Lambda_2) &= \max_{\beta_2 > 0} \Lambda_2 \log(1+\beta_2) e^{-(\Lambda_1+\Lambda_2)\beta_2^{2/\alpha}}.
277: \end{align*}
278: \end{definition}
279:
280: The above formulation is intuitively appealing as a networks choice of access probability constitutes its entire
281: strategy. If we could explicitly solve the maximization problems, the variables $\beta_i$ would be removed altogether.
282: When $\Lambda_1$ and $\Lambda_2$ are large this can be done and
283: \begin{align}
284: \label{eqn:appr_uf1}
285: U_1(\Lambda_1,\Lambda_2) &\approx \frac{\Lambda_1}{(\Lambda_1+\Lambda_2)^{\alpha/2}} \cdot (\alpha/2)^{\alpha/2}e^{-\alpha/2} \\
286: \label{eqn:appr_uf2} U_2(\Lambda_1,\Lambda_2) &\approx \frac{\Lambda_2}{(\Lambda_1+\Lambda_2)^{\alpha/2}} \cdot
287: (\alpha/2)^{\alpha/2}e^{-\alpha/2},
288: \end{align}
289:
290: \noindent but in general it is not possible. Instead, since we are only interested in analyzing the Nash equilibrium
291: (N.E.) or equilibria of this game, we do the following.
292:
293: Observe that the objective function within the maximization is smooth and continuous. This enables the order of
294: maximization to be swapped. That is, for a given $\Lambda_2$, we first maximize over $\Lambda_1$ in equation
295: (\ref{eqn:utility_fns1}) and then over the $\beta_1$. Likewise for equation (\ref{eqn:utility_fns2}). The benefit of
296: this approach is that the maximizing $\Lambda_i$ can be explicitly expressed as a function of $\beta_i$. This was
297: demonstrated earlier for the single network scenario. The resulting expressions are
298: \begin{align*}
299: &U_1(\Lambda_1,\Lambda_2) \\
300: &= \left\{
301: \begin{array}{ll}
302: \Lambda_1\log \left(1+\Lambda_1^{-\alpha/2} \right) e^{-\Lambda_2/\Lambda_1-1}, & \hbox{$\Lambda_1 < N_1$;} \\
303: \max_{\beta_1 > 0} \log(1+\beta_1) e^{-(N_1+\Lambda_2)\beta_1^{2/\alpha}}, & \hbox{$\Lambda_1 = N_1$.}
304: \end{array}
305: \right. \\ \\
306: &U_2(\Lambda_1,\Lambda_2) \\
307: &= \left\{
308: \begin{array}{ll}
309: \Lambda_2\log \left(1+\Lambda_2^{-\alpha/2} \right) e^{-\Lambda_1/\Lambda_2-1}, & \hbox{$\Lambda_2 < N_2$;} \\
310: \max_{\beta_2 > 0} \log(1+\beta_2) e^{-(N_2+\Lambda_1)\beta_2^{2/\alpha}}, & \hbox{$\Lambda_2 = N_2$.}
311: \end{array}
312: \right.
313: \end{align*}
314:
315: The set of N.E. of the above game and their corresponding values of $U_1$ and $U_2$ are identical to those of the
316: Random Access Game. Inspection of the above equations reveals a further simplification of the problem is at hand
317: ---the set of N.E. of the above game are identical to the set of N.E. of the following game (though the values of $U_1$ and $U_2$ at the equilibria may be
318: different).
319: \begin{definition}\label{def:NE} (Transformed Random Access Game)
320: A strategy for network $i$ in the Transformed Random Access Game is a choice of $\Lambda_i \in [0,N_i]$. The payoff
321: functions are
322: \begin{align*}
323: U_1(\Lambda_1,\Lambda_2) &= \Lambda_1\log \left(1+\Lambda_1^{-\alpha/2} \right) e^{-\frac{\Lambda_2}{\Lambda_1}-1} \\
324: U_2(\Lambda_1,\Lambda_2) &= \Lambda_2\log \left(1+\Lambda_2^{-\alpha/2} \right) e^{-\frac{\Lambda_1}{\Lambda_2}-1},
325: \end{align*}
326: \end{definition}
327:
328: We now analyze the N.E. of the Random Access Game by analyzing the N.E. of the Transformed Random Access Game. The
329: first question of interest is whether or not there exists a N.E.? It turns out a unique N.E. always exists but its
330: nature depends crucially on the pathloss exponent. There are two {\it modes}, $2<\alpha<4$ and $\alpha>4$. We start
331: with the first.
332:
333: \begin{theorem}\label{thm:opt_p2} (Random Access N.E. for $2 < \alpha <4$)
334: For $2 < \alpha < 4$ the unique N.E. occurs at $\Lambda_1^* = N_1$, and $\Lambda_2^*$ defined by either the solution of
335: \begin{equation} \label{eqn:opt_p2}
336: N_1 = \Lambda_2\left(\frac{\alpha}{2\left(1+{\Lambda_2}^{\alpha/2}\right)\log \left(1+{\Lambda_2}^{-\alpha/2}\right)} -
337: 1\right)
338: \end{equation}
339: \noindent or $N_1$, whichever is smaller.
340: \end{theorem}
341:
342: \noindent The N.E. described in theorem \ref{thm:opt_p2} occurs on the boundary of the strategy space. This is because
343: for $2<\alpha<4$ each network tries to set its number of transmissions per disc higher than the other (see the proof of
344: the theorem). The equilibrium is then only attained when at least one network has maxed out and scheduled all of its
345: transmissions simultaneously.
346:
347: The N.E. can be better understood when $N_1 \gg 1$ corresponding to the case in which transmissions span several
348: intermediate nodes.
349: \begin{theorem}\label{thm:asymp_soln} (Random Access N.E. for $2 < \alpha <4$ and $N_1 \gg 1$)
350: In the limit $N_1 \rightarrow \infty$ the N.E. occurs at
351: \begin{align*}
352: &(\Lambda_1^*,\Lambda_2^*) \\
353: &\quad\quad = \left\{
354: \begin{array}{ll}
355: (N_1,N_2), & \hbox{$N_1\le N_2\le \dfrac{2}{\alpha-2}N_1$} \\
356: \left(N_1,\dfrac{2}{\alpha-2}N_1\right), & \hbox{$\dfrac{2}{\alpha-2}N_1\le N_2$.}
357: \end{array}
358: \right.
359: \end{align*}
360: \end{theorem}
361:
362: \noindent This result stems from using the limiting form $\log(1+x)\rightarrow x$ as $x\rightarrow 0$ in the utility
363: functions $U_1$ and $U_2$, as was done in equations (\ref{eqn:appr_uf1}) and (\ref{eqn:appr_uf2}). From it we see that
364: if the denser network has more than $\approx 2/(\alpha-2)$ times as many nodes as its rival, the N.E. will correspond
365: to partial reuse, i.e. the denser network will only occupy a fraction of the total available bandwidth. This is in
366: stark contrast to the case of competing {\it individual transmissions} where the N.E. typically corresponds to a full
367: spread, i.e. both competing links spread their power evenly across the entire bandwidth. The limit result of theorem
368: \ref{thm:apx_soln} is plotted in figure \ref{fig:Lambda2_star} as a dashed line.
369:
370: \begin{figure}
371: \centering
372: \includegraphics[width=420pt]{ne.eps}
373: \caption{The solid line represents the solution to equation (\ref{eqn:opt_p2}) for $\alpha = 3$. The N.E. value
374: $\Lambda_2^*$ is equal to the minimum of this line and $N_2$. The limiting solution used in theorem \ref{thm:apx_soln}
375: is plotted as a dashed line.} \label{fig:Lambda2_star}
376: \end{figure}
377:
378: We now investigate the average throughput at equilibrium for the mode $2<\alpha<4$. We define the metric
379: \begin{equation*}
380: U_e = U_1(\Lambda_1^*,\Lambda_2^*)+U_2(\Lambda_1^*,\Lambda_2^*)
381: \end{equation*}
382:
383: \noindent This quantity has a natural interpretation. Recall $\mathbb E {\overline R}_i$ is the average throughput per
384: tx-rx pair and $N_i$ is the average number of tx-rx pairs per transmission disc in network $i$. Thus $U_i = N_i\mathbb
385: E {\overline R}_i$ is the {\it average throughput per transmission disc} in network $i$. This is the average number of
386: bits successfully received in network $i$ within an area of size $\pi d^2$ per time slot, per d.o.f.. The quantity
387: $U_e$ is then the average throughput per transmission disc in the system, that is, the average number of bits
388: successfully received {\it in both networks} within an area of size $\pi d^2$ per time slot, per d.o.f..
389:
390: \begin{theorem}\label{thm:asymp_util}
391: In the limit $N_1\rightarrow \infty$
392: \begin{equation*}
393: U_e \rightarrow \left\{
394: \begin{array}{ll}
395: \dfrac{c_1(\alpha)}{N_1^{\alpha/2-1}} & \hbox{$N_1\le N_2\le \frac{2}{\alpha-2}N_1$} \\
396: \dfrac{c_2(\alpha)}{(N_1+N_2)^{\alpha/2-1}}, & \hbox{$\frac{2}{\alpha-2}N_1\le N_2$.}
397: \end{array}
398: \right.
399: \end{equation*}
400: \noindent where $c_1(\alpha) = \left(\alpha/2-1\right)^{\alpha/2-1}(\alpha/2)e^{-\alpha/2}$ and $c_2(\alpha) =
401: (\alpha/2)^{\alpha/2}e^{-\alpha/2}$.
402: \end{theorem}
403:
404: \noindent The important property of this result is that as the number of nodes per transmission disc increases, $U_e$
405: decreases roughly like $1/(N_1+N_2)^{\alpha/2-1}$. Let us compare this to the average throughput per transmission disc
406: in the cooperative case, that is when the two networks behave as if they were a single network with $N_1+N_2$ nodes per
407: disc. From equation (\ref{eqn:one_network_thpt}) this average throughput per disc is
408: \begin{equation*}
409: U_c = \Lambda^*\log(1+{\Lambda^*}^{-\alpha/2})
410: \end{equation*}
411:
412: \noindent which is independent of the number of nodes per disc. Thus as the number of nodes per disc grows, so does the
413: price of anarchy
414: \begin{equation*}
415: \frac{U_c}{U_e} = O\left(N_1^{\alpha/2-1} \right).
416: \end{equation*}
417:
418: \noindent For $\alpha>4$ the N.E. behavior is different. Whereas for $2<\alpha<4$ the solution always lies on the
419: boundary, for $\alpha>4$ it typically does not.
420:
421: \begin{theorem}\label{thm:a_ge_4} (Random Access N.E. for $\alpha > 4$)
422: For $\alpha > 4$ the unique N.E. occurs at
423: \begin{equation*}
424: (\Lambda_1^*,\Lambda_2^*) = ( \sqrt{\Lambda^*(\alpha/2)}, \sqrt{\Lambda^*(\alpha/2)} )
425: \end{equation*}
426:
427: \noindent if $\sqrt{\Lambda^*(\alpha/2)} < N_1$, otherwise $\Lambda_1^* = N_1$ and $\Lambda_2^*$ is defined by either
428: the solution of equation (\ref{eqn:opt_p2}) or $N_2$, whichever is smaller.
429: \end{theorem}
430:
431: \noindent A plot of $\sqrt{\Lambda^*(\alpha/2)}$ versus $\alpha$ is given in figure \ref{fig:Lambda_prime}. The
432: condition $\sqrt{\Lambda^*(\alpha/2)} < N_1$ corresponds to network $1$ having more than $\sqrt{\Lambda^*(\alpha/2)}$
433: nodes per transmission disc. We refer to this as the {\it partial/partial reuse} regime.
434:
435: %\begin{figure}
436: %\centering
437: %\includegraphics[width=250pt]{sqrtLambda_star.eps}
438: %\caption{The function $\sqrt{\Lambda^*(\alpha/2)}$ compared with $\Lambda^*(\alpha)$}\label{fig:sqrtLambda_star}
439: %\end{figure}
440:
441: The interpretation of theorem \ref{thm:a_ge_4} is that for $\alpha>4$ in the partial/partial reuse regime, the solution
442: lies in the strict interior of the strategy space. This is because on the boundary of the space network $i$ can improve
443: its throughput by undercutting the transmit density of network $j$, i.e. setting $\Lambda_i<\Lambda_j$. The symmetry of
444: the N.E. ($\Lambda_1^*=\Lambda_2^*$) then follows by observing the utility functions are symmetric and the solution is
445: unique.
446:
447: \begin{figure}
448: \centering
449: \includegraphics[width=210pt]{regimes2_5a.eps}
450: \includegraphics[width=210pt]{regimes4_1a.eps}
451: \includegraphics[width=210pt]{regimes3_5a.eps}
452: \includegraphics[width=210pt]{regimes4_5a.eps}
453: \includegraphics[width=210pt]{regimes4a.eps}
454: \includegraphics[width=210pt]{regimes6a.eps}
455: \caption{These plots can be used to determine which regime the N.E. is in. The x-axis and y-axis corresponds to the
456: average number of nodes per transmission disc in network 1 and 2, respectively. Note the lower left vertex of the
457: partial/partial reuse regime always occurs at $(\sqrt{\Lambda^*(\alpha)},\sqrt{\Lambda^*(\alpha)})$ and the
458: intersection of the full/full reuse regime with the axes always occurs at $(\Lambda^*(\alpha),0)$ and
459: $(0,\Lambda^*(\alpha))$.} \label{fig:regimes}
460: \end{figure}
461:
462: There is a cooperative flavor to this equilibrium in that both networks set their transmission densities to the same
463: level, and this level is comparable to the optimal single network density $\Lambda^*(\alpha)$. Moreover the equilibrium
464: level does not grow with the number of nodes per transmission disc, as it does for $2<\alpha<4$. Actual cooperation
465: between networks corresponds to setting the access probability based on equation (\ref{eqn:p_star}), taking into
466: account that the effective node density is $\lambda_1+\lambda_2$. Thus the cooperative solution is
467: \begin{equation*}
468: (\Lambda_1^*,\Lambda_2^*) = \left( \frac{\lambda_1}{\lambda_1+\lambda_2}\Lambda^*,
469: \frac{\lambda_2}{\lambda_1+\lambda_2}\Lambda^* \right).
470: \end{equation*}
471:
472: \noindent Under cooperation the average throughput per transmission disc is (from equation
473: (\ref{eqn:one_network_thpt}))
474: \begin{equation*}
475: U_c = \frac{1}{e}\Lambda^*(\alpha)\log\left(1+\frac{1}{{\Lambda^*(\alpha)}^{\alpha/2}}\right).
476: \end{equation*}
477:
478: \noindent Under cooperation in the partial/partial reuse regime it is
479: \begin{equation*}
480: U_e = \frac{2}{e^2}\sqrt{\Lambda^*(\alpha/2)}\log\left(1+\frac{1}{{\Lambda^*(\alpha/2)}^{\alpha/4}}\right)
481: \end{equation*}
482:
483: \begin{figure}
484: \centering
485: \includegraphics[width=420pt]{poa.eps}
486: \caption{For $\alpha>4$, the price of anarchy depends only on the pathloss exponent in the partial/partial reuse
487: regime. }\label{fig:poa}
488: \end{figure}
489:
490: \noindent The price of anarchy is the ratio of these two quantities ($U_e/U_c$) and is plotted in figure \ref{fig:poa}.
491: Comparing the two modes we see that whereas for $2<\alpha<4$ the price of anarchy grows in an unbounded fashion with
492: the number of nodes per transmission disc, for $\alpha>4$ the price of anarchy in the partial/partial reuse regime is a
493: constant depending only on $\alpha$.
494:
495: We now summarize the equilibria results. There are three regimes.
496: \begin{enumerate}
497: \item {\it Full/Full reuse} \\
498: - $N_1 \le \sqrt{\Lambda^*(\alpha/2)}$ and $N_1 \le N_2\left(\alpha/2(1+{N_2}^{\alpha/2})\log
499: (1+{N_2}^{-\alpha/2}) - 1\right)$ \\
500: - both networks schedule all transmissions
501: \item {\it Full/Partial reuse} \\
502: - $N_1 \le \sqrt{\Lambda^*(\alpha/2)}$ and $N_1 > N_2\left(\alpha/2(1+{N_2}^{\alpha/2})\log
503: (1+{N_2}^{-\alpha/2}) - 1\right)$ \\
504: - denser network schedules all transmissions, sparser schedules only a fraction
505: \item {\it Partial/Partial reuse} \\
506: - $N_1 > \sqrt{\Lambda^*(\alpha/2)}$ \\
507: - both networks schedule only a fraction of their transmissions
508: \end{enumerate}
509:
510: \noindent In the full/full reuse regime $(\Lambda_1^*,\Lambda_2^*) = (N_1, N_2)$. In the full/partial reuse regime the
511: sparser network sets $\Lambda_1^* = N_1$ and the denser network sets $\Lambda_2^*$ as the solution to equation
512: (\ref{eqn:opt_p2}). In the partial/partial reuse regime $(\Lambda_1^*,\Lambda_2^*) = (\sqrt{\Lambda^*(\alpha/2)},
513: \sqrt{\Lambda^*(\alpha/2)})$.
514:
515: The regimes are essentially distinguished by which boundary constraints are active. For $2 < \alpha < 4$ the
516: partial/partial reuse regime is not accessible. Figure \ref{fig:regimes} provides an illustrated means for determining
517: which regime the system is in, for a range of values of the pathloss exponent. In these plots we consider all values of
518: $N_1$ and $N_2$, not just those satisfying $N_1\ge N_2$. Notice that as $\alpha \rightarrow 2$ the entire region
519: corresponds to the full/full reuse regime, for $\alpha = 4$ almost the entire region corresponds to the full/partial
520: reuse regime and for $\alpha \rightarrow \infty$ the entire region corresponds to the partial/partial reuse regime.
521:
522: \subsection{Variable-Rate model}
523: In this section we examine the case where tx-rx pairs tailor their communication rates to suit instantaneous channel
524: conditions, sending at rate $\log(1+{\SIR(t)})$ during the $t$th time slot. Various protocols can be used to enable
525: tx-rx pairs to estimate their ${\SIR}(t)$.
526:
527: Consider first the single, isolated network scenario. The expected time-averaged rate per user is now
528: \begin{equation*}
529: \mathbb E{\overline R}_i = p_i\mathbb E\log(1+{\SIR}).
530: \end{equation*}
531:
532: \noindent As before, the rate is both time-averaged over the interference and averaged over the geographic distribution
533: of the nodes. The $\SIR$ is the instantaneous value observed by a given rx node and is distributed according to
534: \begin{align*}
535: \mathbb P ({\SIR} > x) &= \mathbb P (r > x^{1/\alpha}d) \\
536: &= \left( 1 - \frac{\pi x^{2/\alpha}d^2p_i}{n} \right)^{n\lambda_i}
537: \end{align*}
538:
539: \noindent where the variable $r$ denotes the distance to the nearest interferer. Thus
540: \begin{align*}
541: \mathbb E{\overline R}_i &= p_i\int {\mathbb P} \left( \log(1+{\SINR}) > s\right) ds \\
542: &= p_i\int {\mathbb P}\left( {\SINR} > x \right) \frac{dx}{1+x} \\
543: &= p_i\int_{0}^{\left(\frac{n}{p_i\pi d^2}\right)^{\alpha/2}} \left( 1 - \frac{\pi x^{2/\alpha}d^2p_i}{n}
544: \right)^{n\lambda_i} \frac{dx}{1+x} \\
545: &\rightarrow p_i\int_0^{\infty} e^{-\pi p_i \lambda_i d^2 x^{2/\alpha}} \frac{dx}{1+x}
546: \end{align*}
547:
548: \noindent in the limit $n \rightarrow \infty$. Changing variables and optimizing we have
549: \begin{equation}\label{eqn:p_prime}
550: \mathbb E{\overline R}_i \rightarrow \frac{1}{N_i} \max_{0\le \Lambda_i \le 1} \Lambda_i\int_0^{\infty} e^{-\Lambda_i
551: x^{2/\alpha}} \frac{dx}{1+x}.
552: \end{equation}
553:
554: \noindent Define $\Lambda'(\alpha)$ as the maximizing argument of the unconstrained version of the above optimization
555: problem, or more specifically as the unique solution to
556: \begin{equation*}
557: \int_0^{\infty} \frac{1 - \Lambda'x^{2/\alpha}}{1+x} e^{-\Lambda x^{2/\alpha}} dx = 0.
558: \end{equation*}
559:
560: \noindent The function $\Lambda'(\alpha)$ is plotted in figure \ref{fig:Lambda_prime}. Then
561: \begin{equation*}
562: p_i^* = \min(\Lambda'(\alpha)/N_i,1)
563: \end{equation*}
564:
565: \noindent From this we see that the solution for the variable-rate case is the same as the fixed-rate solution,
566: differing only by substitution of the function $\Lambda'(\alpha)$ for $\Lambda^*(\alpha)$.
567:
568: Now we turn to the case of two competing wireless networks. Using an approach similar to the one above it can be shown
569: that
570: \begin{equation*}
571: \mathbb E{\overline R}_i \rightarrow \frac{1}{N_i} \Lambda_i\int_0^{\infty} e^{-(\Lambda_1+\Lambda_2) x^{2/\alpha}}
572: \frac{dx}{1+x}.
573: \end{equation*}
574:
575: \noindent In this way we can define the game between the two networks like so.
576: \begin{definition} (Variable Rate Random Access Game)
577: A strategy for network $i$ in the Variable Rate Random Access Game is a choice of $\Lambda_i\in[0,\pi\lambda_i d^2]$.
578: The payoff functions are
579: \begin{align*}
580: U_1(\Lambda_1,\Lambda_2) &= \Lambda_1 \int_0^{\infty} e^{-(\Lambda_1+\Lambda_2) x^{2/\alpha}} \frac{dx}{1+x} \\
581: U_2(\Lambda_1,\Lambda_2) &= \Lambda_2 \int_0^{\infty} e^{-(\Lambda_1+\Lambda_2) x^{2/\alpha}} \frac{dx}{1+x}.
582: \end{align*}
583: \end{definition}
584:
585: From the above definition we see that the Fixed-Rate game is derived from the Variable-Rate game by merely applying a
586: step-function lower bound to the players utility functions, with the width of the step-function optimized, i.e.
587: \begin{align*}
588: &\int_0^{\infty} e^{-(\Lambda_1+\Lambda_2) x^{2/\alpha}} \frac{dx}{1+x} \\
589: &\quad\quad\quad\quad \ge \max_{\beta_i>0}
590: e^{-(\Lambda_1+\Lambda_2)\beta_i^{2/\alpha}}\int_{0}^{\beta_i}\frac{dx}{1+x} \\
591: &\quad\quad\quad\quad= \max_{\beta_i>0} \log(1+\beta_i)e^{-(\Lambda_1+\Lambda_2)\beta_i^{2/\alpha}}.
592: \end{align*}
593:
594: \noindent A plot comparing the expression on the left as a function of $\Lambda_1+\Lambda_2$, to the expression on the
595: right as a function of $\Lambda_1+\Lambda_2$, for $\alpha = 4$, is presented in figure \ref{fig:vr_vs_fr}. The figure
596: suggests that both expressions share a similar functional dependency on $\Lambda_1+\Lambda_2$. It is therefore natural
597: to wonder whether, as a consequence of this close relationship, the N.E. of the Variable-Rate game bears any
598: relationship to the N.E. of the Fixed-Rate game?
599:
600: \begin{figure}
601: \centering
602: \includegraphics[width=420pt]{vr_vs_fr.eps}
603: \caption{The utility functions used in the Fixed-Rate model are lower bounds of those used in the Variable-Rate model.
604: In this sample plot $\alpha = 4$. The y-axis represents $U_i/\Lambda_i$. }\label{fig:vr_vs_fr}
605: \end{figure}
606:
607: As in the Fixed-Rate game, the utility functions of the Variable-Rate game can be explicitly evaluated when $\Lambda_1$
608: and $\Lambda_2$ are large yielding
609: \begin{align*}
610: U_1(\Lambda_1,\Lambda_2) &\approx \frac{\Lambda_1}{(\Lambda_1+\Lambda_2)^{\alpha/2}}\cdot \Gamma(\alpha/2+1) \\
611: U_2(\Lambda_1,\Lambda_2) &\approx \frac{\Lambda_2}{(\Lambda_1+\Lambda_2)^{\alpha/2}}\cdot \Gamma(\alpha/2+1).
612: \end{align*}
613:
614: \noindent Comparing with equations (\ref{eqn:utility_fns1}) and (\ref{eqn:utility_fns2}) we see that for large
615: $\Lambda_1,\Lambda_2$, the utility functions of the Variable-Rate game have exactly the same functional dependency on
616: $\Lambda_1,\Lambda_2$ as the utility functions of the Fixed-Rate game, differing only in an $\alpha$-dependent
617: constant. These constants are plotted in figure \ref{fig:consts}. The plots illustrate the benefit in total system
618: throughput that stems from playing the Variable-rate game in place of the Fixed-Rate game.
619:
620: \begin{figure}
621: \centering
622: \includegraphics[width=420pt]{consts.eps}
623: \caption{The total system throughputs at equilibrium for the Variable-Rate and Fixed-Rate games differ only by
624: $\alpha$-dependent constants. These constants are plotted above. For the Variable-Rate game the constant is
625: $\Gamma(\alpha/2+1)$ versus $(\alpha/2)^{\alpha/2}e^{-\alpha/2}$ for the Fixed-Rate game.} \label{fig:consts}
626: \end{figure}
627:
628: As anticipated by the above discussion, the N.E. behavior of the Variable-Rate game parallels that of the Fixed-Rate
629: game. The same two modes are present, $2<\alpha<4$ and $\alpha>4$. These give rise to the same three spreading regimes,
630: the only difference being that the boundaries delineating them are shifted slightly. The N.E. values
631: $(\Lambda_1^*,\Lambda_2^*)$ in each regime take on a similar form.
632: \begin{theorem} \label{thm:vr_soln}
633: The Variable-Rate Random Access Game has a unique N.E. $(\Lambda_1^*,\Lambda_2^*)$ which lies in one of three regions.
634: Let $\Lambda''(\alpha)$ be the unique solution of
635: \begin{equation}\label{eqn:Lambda_doubleprime}
636: \int_0^{\infty} \frac{1 - \Lambda''x^{2/\alpha}}{1+x} e^{-2\Lambda''x^{2/\alpha}} dx = 0.
637: \end{equation}
638:
639: \noindent for $\alpha > 4$, and equal to positive infinity for $\alpha\le 4$.
640: \begin{itemize}
641: \item (Full/Full reuse) If $N_1 \le \Lambda''(\alpha)$ and $N_2 \le$ the unique solution over $\Lambda$ of
642: \begin{equation} \label{eqn:vr_thm}
643: \int_0^{\infty} \frac{1 - \Lambda x^{2/\alpha}}{1+x} e^{-(N_1+\Lambda)x^{2/\alpha}} dx = 0,
644: \end{equation}
645: \noindent then $(\Lambda_1^*,\Lambda_2^*) = (N_1,N_2)$.
646: \item (Full/Partial reuse) If $N_1 \le \Lambda''(\alpha)$ and $N_2 >$ the unique solution of equation (\ref{eqn:vr_thm}) then
647: $\Lambda_1^* = 1$ and $\Lambda_2^*$ is equal to this unique solution.
648: \item (Partial/Partial reuse) If $N_1 > \Lambda''(\alpha)$ then $(\Lambda_1^*,\Lambda_2^*) = (\Lambda''(\alpha),\Lambda''(\alpha))$.
649: \end{itemize}
650: \end{theorem}
651:
652: \noindent A regime map is provided in figure \ref{fig:vr_regimes}. As is evident from the above theorem, it is not
653: possible to characterize the behavior of the N.E. for the Variable-Rate game as explicitly as for the Fixed-Rate game.
654: This is in part due to the more complex representation of the utility functions in terms of integrals, and in part due
655: to the fact that the the function $\Lambda''(\alpha)$ cannot be represented in terms of the function
656: $\Lambda'(\alpha)$, as in the case of the Fixed-Rate game, where one function equals the square-root of the other
657: evaluated at $\alpha/2$.
658:
659: \begin{figure}
660: \centering
661: \includegraphics[width=420pt]{vr_regimes.eps}
662: \caption{The three regimes for the N.E. of the Variable-Rate Game.} \label{fig:vr_regimes}
663: \end{figure}
664:
665: For large $N_1$ however, we can use the approximation adopted in theorem \ref{thm:apx_soln} to explicitly characterize
666: the behavior of the N.E. in the full/partial reuse regime.
667: \begin{theorem}\label{thm:apx_soln_vr} (Variable-Rate Random Access N.E. for $2 < \alpha <4$ and $N_1 \gg 1$)
668: In the limit $N_1 \rightarrow \infty$ the N.E. occurs at
669: \begin{align*}
670: &(\Lambda_1^*,\Lambda_2^*) \\
671: &\quad\quad = \left\{
672: \begin{array}{ll}
673: (N_1,N_2), & \hbox{$N_1\le N_2\le \dfrac{2}{\alpha-2}N_1$} \\
674: \left(N_1,\dfrac{2}{\alpha-2}N_1\right), & \hbox{$\dfrac{2}{\alpha-2}N_1\le N_2$.}
675: \end{array}
676: \right.
677: \end{align*}
678: \end{theorem}
679:
680: \begin{figure}
681: \centering
682: \includegraphics[width=420pt]{Lambda_prime.eps}
683: \caption{The functions $\Lambda''(\alpha)$ and $\Lambda'(\alpha)$ for the Variable-Rate model versus the equivalent
684: functions $\Lambda^*(\alpha)$ and $\sqrt(\Lambda^*(\alpha/2))$ for the Fixed-Rate model.}\label{fig:Lambda_prime}
685: \end{figure}
686:
687: \noindent Thus for $2 < \alpha < 4$ and large $N_1$, the behavior of the N.E. in the Variable-Rate game is identical to
688: that of the Fixed-Rate game. As discussed earlier, the values of $U_1$ and $U_2$ at equilibrium are equal to those of
689: the Fixed-Rate game times a constant $(\alpha/2)^{\alpha/2}e^{-\alpha/2}/\Gamma(\alpha/2+1)$.
690:
691: \subsection{Explanation of Behavior}
692: The intuition behind our result is the following. The average throughput per link is essentially equal to the product
693: of the fraction of time transmissions are scheduled, and the average number of bits successfully communicated per
694: transmission. Adjusting the transmit density has a linear effect on the former term, but a non-linear effect on the
695: latter. The latter depends on the $\SIR$ and the $\SIR$ essentially depends on the pathloss exponent via
696: \begin{equation*}
697: {\SIR} \approx \left(\frac{\text{distance to transmitter}}{\text{distance to interferer}}\right)^{\alpha}
698: \end{equation*}
699:
700: \noindent When the nearest interferer is closer than the transmitter, the ratio inside the parentheses is less than
701: one, and a large value of $\alpha$ substantially hurts the $\SIR$, dragging it to near zero and causing the link
702: capacity to drop to near zero. However, when the nearest interferer is further away than the transmitter, the ratio is
703: greater than one and a large value of $\alpha$ substantially improves the $\SIR$, resulting in a large link capacity.
704: Thus for large $\alpha$ the average number of bits successfully communicated per transmission is very sensitive to
705: whether or not the transmission disc is empty.
706:
707: This insensitivity for sufficiently small $\alpha$ means that increasing the transmit density in network $i$ causes a
708: linear increase in the fraction of time transmissions are scheduled, but has little effect on the number of bits
709: successfully communicated per transmission, up until the point where the transmit density of network $i$ starts to
710: dwarf the transmit density of network $j$. Thus network $i$ will wish to increase its transmit density until it is
711: sufficiently larger than network $j$'s. Likewise network $j$ will wish to increase its transmit density until it is
712: sufficiently larger than network $i$'s. Ultimately this results in either
713: \begin{enumerate}
714: \item a full/full reuse solution, which occurs when the sparser network max's out and winds up simultaneously scheduling
715: all of its transmissions, and the denser network is insufficiently dense such that its optimal transmit density based
716: on the sparser networks choice, results in it simultaneously scheduling all of its transmissions, or
717: \item a full/partial reuse solution, which occurs when the sparser network max's out and winds up simultaneously scheduling
718: all of its transmissions but the denser network is sufficiently dense such that its optimal transmit density based on
719: the sparser networks choice, results in it simultaneously scheduling only a fraction of its transmissions.
720: \end{enumerate}
721:
722: The opposite effect occurs for sufficiently large $\alpha$, where the average number of bits successfully communicated
723: per transmission depends critically on whether or not there is an interferer inside the transmission disc. In this
724: scenario network $i$ will set its active density to a level lower than network $j$'s, in order to capitalize on those
725: instances in which network $j$ happens to not schedule any transmissions nearby to one of network $i$'s receivers,
726: resulting in the successful communication of a large number of bits. Likewise network $j$ will set its active density
727: to a level lower than network $i$'s, and the system converges to the partial/partial reuse regime.
728:
729: \section{Variable Transmission Range}
730: One of our initial assumptions was that all tx-rx pairs have the same transmission range $d$. In this section we
731: consider the scenario where the transmission ranges of all tx-rx pairs in the system are i.i.d. random variables $D_j$.
732: When the variance of $D_j$ is large, some form of power control may be required to ensure long range transmissions are
733: not unfairly penalized. A natural form of power control involves tx nodes setting their transmit powers such that all
734: transmissions are received at the same $\SNR$. This means transmit power scales proportional to $D_j^{\alpha}$. Denote
735: the distance from the $k$th tx node to the $j$th rx node $D_{ij}$. Then the interference power from the $k$th tx node
736: impinging on the $j$th rx node is proportional to $D_{kk}^{\alpha}/D_{kj}^{\alpha}$. In the fixed transmission range
737: scenario this quantity was proportional to $1/D_{kj}^{\alpha}$. There we assumed the bulk of the interference was
738: generated by the dominant interferer. Denote the scheduled set of tx nodes at time $t$ by ${\cal A}(t)$. This
739: assumption essentially evoked the following approximation
740: \begin{equation*}
741: \sum_{k\in\cal A(t)} 1/D_{kj}(t)^{\alpha} \approx \max_{k\in \cal A(t)} 1/D_{kj}(t)^{\alpha}.
742: \end{equation*}
743:
744: \noindent The equivalent approximation for the variable transmission range problem is
745: \begin{equation*}
746: \sum_{k\in\cal A(t)} D_{kk}^{\alpha}/D_{kj}(t)^{\alpha} \approx \max_{k\in \cal A(t)}
747: D_{kk}^{\alpha}/D_{kj}(t)^{\alpha}.
748: \end{equation*}
749:
750: \noindent Thus for variable range transmission the dominant interferer is not necessarily the nearest to the receiver.
751: Under this assumption the $\SIR$ at the $j$th rx node at time $t$ is then
752: \begin{equation*}
753: {\SIR}_j(t) = \frac{D_{k^*k^*}^{-\alpha}}{D_{k^*j^*}^{-\alpha}}
754: \end{equation*}
755:
756: \noindent where $k^*$ is the index of the tx node that is closest to the $j$th receiver relative to its transmission
757: range.
758:
759: Let us compute the throughput for the variable transmission range model under the Fixed-Rate Random Access protocol.
760: \begin{equation*}
761: {\mathbb E} {\overline R} = p_i {\mathbb P} \left( {\SIR}_j(t) > \beta_i\right)\log(1+\beta_i).
762: \end{equation*}
763:
764: \noindent \noindent The probability the $\SIR$ is greater than the threshold
765: \begin{align*}
766: {\mathbb P} ({\SINR_j(t)> \beta_i}) &= {\mathbb P} \left( D_{kj} > \beta_i^{1/\alpha} D_{kk}, \forall k\right) \\
767: &= \left( \int_{0}^{\sqrt{n/\pi}/\beta_i^{1/\alpha}} {\mathbb P} \left( D_{kj} > \beta_i^{1/\alpha} x \right) {\mathbb
768: P}_{D_{kk}}(x)dx\right)^{n\lambda} \\
769: &= \left( \int_{0}^{\sqrt{n/\pi}/\beta_i^{1/\alpha}} \left(1-\frac{p_i\pi\beta_i^{2/\alpha}x^2}{n}\right) {\mathbb
770: P}_{D_{kk}}(x)dx\right)^{n\lambda} \\
771: &= \left( 1 - \frac{p_i\pi\beta_i^{2/\alpha}}{n} \int_{0}^{\sqrt{n/\pi}/\beta_i^{1/\alpha}} x^2 {\mathbb
772: P}_{D_{kk}}(x)dx\right)^{n\lambda} \\
773: &= \left( 1 - \frac{p_i\pi\beta_i^{2/\alpha}{\mathbb E}D_{kk}^2}{n}\right)^{n\lambda} \\
774: &\rightarrow e^{-p_i\pi \beta_i^{2/\alpha}{\mathbb E}D_{kk}^2}
775: \end{align*}
776:
777: \noindent as $n\rightarrow \infty$. For notational simplicity let $d\equiv D_{kk}$. If we define $N_i$ as the average
778: number of nodes per transmission disc, where the average is taken over both the geographical distribution of the nodes
779: {\it and the distribution of the size of the transmission disc}, i.e.
780: \begin{equation*}
781: N_i = \pi \lambda_i {\mathbb E}_d^2
782: \end{equation*}
783:
784: \noindent we wind up with ${\mathbb E}{\overline R} \rightarrow p_i \log(1+\beta_i) e^{-N_ip_i\beta_i^{2/\alpha}}$,
785: which is the same result as the fixed-transmission range model. It is straightforward to extend the analysis to the
786: case of two competing networks. The throughput per user in network 1 is then
787: \begin{equation*}
788: {\mathbb E}{\overline R} \rightarrow p_1 \log(1+\beta_1) e^{-(N_1p_1+N_2p_2)\beta_1^{2/\alpha}}.
789: \end{equation*}
790:
791: \noindent Likewise for network 2. From this we see that all results for the fixed-transmission range model extend to
792: the variable-transmission range model by simply replacing $d^2$ by ${\mathbb E} d^2$.
793:
794: \section{Simulations}
795: \subsection{Random Access Protocol}
796:
797: In order to get a sense of the typical behavior of the players in the (Variable-Rate) Random Access game, and to
798: justify the validity of the Dominant Interferer assumption, we simulate the behavior of the following greedy algorithm
799: with the interference computed based on all transmissions in the network, not just the strongest. \\
800:
801: \noindent {\bf Inputs:} $p_1(0),p_2(0),\Delta$ \\
802: {\bf Outputs:} ${\bf p}_i = [p_i(1),\dots,p_i(500)]$, for $i=1,2$. \\
803:
804:
805: \noindent {\bf For} $t = 1$ to $500$ \\
806:
807: Form estimate $\overline{R}_1(p_1(t-1) + \Delta,p_2(t-1))$
808:
809: Form estimate $\overline{R}_1(p_1(t-1) - \Delta,p_2(t-1))$ \\
810:
811: {\bf If} $\overline{R}_1(p_1(t-1) + \Delta,p_2(t-1)) > \overline{R}_1(p_1(t-1) - \Delta,p_2(t-1))$
812:
813: \hspace{3 mm} $p_1(t) = \min (p_1(t-1) + \Delta,1)$
814:
815: {\bf Else}
816:
817: \hspace{3 mm} $p_1(t) = \max (p_1(t-1) - \Delta, 0)$
818:
819: {\bf End} \\
820:
821: Form estimate $\overline{R}_2(p_1(t), p_2(t-1) + \Delta)$
822:
823: Form estimate $\overline{R}_2(p_1(t), p_2(t-1) - \Delta)$ \\
824:
825: {\bf If} $\overline{R}_2(p_1(t), p_2(t-1) + \Delta) > \overline{R}_2(p_1(t),p_2(t-1) - \Delta)$
826:
827: \hspace{3 mm} $p_2(t) = \min( p_2(t-1) + \Delta, 1)$
828:
829: {\bf Else}
830:
831: \hspace{3 mm} $p_2(t) = \max(p_2(t-1) - \Delta,0)$
832:
833: {\bf End} \\
834:
835: \noindent {\bf End} \\
836:
837: Each update time $t$, network 1 temporarily sets its access probability to $p_1(t-1)+\Delta$ and measures the resulting
838: throughput, averaged over 200 transmission times. This is denoted $\overline{R}_1(p_1(t-1) + \Delta,p_2(t-1))$. It then
839: repeats this measurement for an access probability of $p_1(t-1)-\Delta$. This is denoted $\overline{R}_1(p_1(t-1) -
840: \Delta,p_2(t-1))$. It then either permanently increases its access probability to $p_1(t) = p_1(t-1)+\Delta$ or
841: permanently decreases it to $p_1(t) = p_1(t-1) - \Delta$ depending on which option it estimates will lead to a higher
842: throughput. Now network 2 performs the same operation. It uses a total of 400 time slots to measure the effect of
843: increasing versus decresing its access probability and then either sets $p_2(t) = p_2(t-1)+\Delta$ or $p_2(t) =
844: p_2(t-1)-\Delta$. If $\Delta$ is small, then both networks can perform these measurement operations simultaneously
845: without significantly affecting the outcome.
846:
847: The topology used in the simulations consisted of 400 tx nodes from network 1 and 200 tx nodes from network 2, all
848: i.i.d. uniformly distributed in a square of unit area. For each tx node, its corresponding rx node was located at a
849: point randomly chosen at uniform from a disc of radius 0.15. This corresponds to $N_1=400/\sqrt(800)\approx 14.14$ and
850: $N_2=200/\sqrt(800) \approx 7.28$. A step-size of $\Delta = 0.02$ was used. When computing the throughputs, in order to
851: avoid boundary effects, only transmissions emanating from those tx nodes in the interior of the network were counted.
852: The results for $\alpha = 2.5, 3.5$ and $4.5$ are displayed in figure \ref{fig:RA_sim}. The observed behavior
853: corresponds to the analytical results. For the values of $N_1$ and $N_2$ used, the N.E. lies in the Full/Full regime
854: for $\alpha=2.5$, the Full/Partial regime for $\alpha=3.5$, and the Partial/Partial regime for $\alpha=4.5$, as can be
855: seen from figure \ref{fig:regimes}.
856:
857: \begin{figure}
858: \centering
859: \includegraphics[width=210pt]{RAa2_5.eps}\\
860: \includegraphics[width=210pt]{RAa3_5.eps}\\
861: \includegraphics[width=210pt]{RAa4_5.eps}
862: \caption{Simulations of greedy algorithm under Random Access protocol.}\label{fig:RA_sim}
863: \end{figure}
864:
865: \subsection{Carrier Sensing Multiple Access based protocol}
866:
867: The high level conclusion from our analysis of the Random Access protocol, is that the Nash Equilibrium is cooperative
868: in nature for a sufficiently high pathloss exponent. Ideally we would like to be able to draw this conclusion for a
869: more sophisticated class of scheduling protocols employing carrier sensing. Due to the analytical intractability of the
870: problem, we present simulation results to illustrate this effect. We assume both networks operate under the following
871: protocol. We present a centralized version of it due to space constraints, but claim there exists a distributed version
872: that performs identically in most cases. During the scheduling phase, each tx-rx pair is assigned a unique token at
873: random from $\{1,\dots,n\}$. Tx nodes proceed with their transmission so long as they will not cause excessive
874: interference to any rx node with a higher priority token. More precisely, a transmission is scheduled so long as for
875: each rx node with higher priority, the difference between its received signal power in dB and the interference power
876: from the lower priority tx node in dB, exceeds a silencing threshold $\gamma_i$ ($i=1$ for network 1 and $i=2$ for
877: network 2). Thus a game between the two networks can be defined where the strategies are the choices of silencing
878: thresholds $\gamma_1$ and $\gamma_2$. We refer to this as the {\it CSMA game}. The silencing threshold for the CSMA
879: game essentially plays the same role as the access probability in the Random Access game -it determines the degree of
880: spatial reuse. A high value of $\gamma$ leads to a low density of transmissions, a low value of $\gamma$ leads to a
881: high density.
882:
883: We simulate the behavior that arises when both networks optimize their silencing thresholds in a greedy manner.
884: Analogously to before, we have the following algorithm. \\
885:
886: \noindent {\bf Inputs:} $p_1(0),p_2(0),\Delta$ \\
887: {\bf Ouputs:} ${\bf p}_i = [p_i(1),\dots,p_i(500)]$, for $i=1,2$. \\
888:
889: \noindent {\bf For} $t = 1$ to $500$ \\
890:
891: Form estimate $\overline{R}_1(\gamma_1(t-1) + \Delta,\gamma_2(t-1))$
892:
893: Form estimate $\overline{R}_1(\gamma_1(t-1) - \Delta,\gamma_2(t-1))$ \\
894:
895: {\bf If} $\overline{R}_1(\gamma_1(t-1) + \Delta,\gamma_2(t-1)) > \overline{R}_1(\gamma_1(t-1) - \Delta,\gamma_2(t-1))$
896:
897: \hspace{3 mm} $\gamma_1(t) = \min (\gamma_1(t-1) + \Delta,30{\sf dB})$
898:
899: {\bf Else}
900:
901: \hspace{3 mm} $\gamma_1(t) = \max (\gamma_1(t-1) - \Delta, -30{\sf dB})$
902:
903: {\bf End} \\
904:
905: Form estimate $\overline{R}_2(\gamma_1(t), \gamma_2(t-1) + \Delta)$
906:
907: Form estimate $\overline{R}_2(\gamma_1(t), \gamma_2(t-1) - \Delta)$ \\
908:
909: {\bf If} $\overline{R}_2(\gamma_1(t), \gamma_2(t-1) + \Delta) > \overline{R}_2(\gamma_1(t),\gamma_2(t-1) - \Delta)$
910:
911: \hspace{3 mm} $\gamma_2(t) = \min( \gamma_2(t-1) + \Delta, 30{\sf dB})$
912:
913: {\bf Else}
914:
915: \hspace{3 mm} $\gamma_2(t) = \max(\gamma_2(t-1) - \Delta,-30{\sf dB})$
916:
917: {\bf End} \\
918:
919: \noindent {\bf End} \\
920:
921: The topology used in the setup is identical to before, the only exception being that at each iteration of the
922: algorithm, 10 old tx-rx pairs leave each network, and 10 new pairs join in i.i.d. locations drawn uniformly at random.
923: This is to ensure sufficient averaging.
924:
925: In a similar fashion to before, each network estimates the effect of either increasing or decreasing the silencing
926: threshold and then makes a permanent choice. For the same parameter values, the results of the simulation are displayed
927: in figure \ref{fig:CSMA_sim}. On the y-axes of these plots we have drawn the fraction of nodes simultaneously scheduled
928: at each iteration, which we denote $f_1$ and $f_2$, rather than the silencing thresholds $\gamma_i$, in order to draw a
929: simple visual comparison with figure \ref{fig:RA_sim}. For this reason there is more fluctuation in the results, as the
930: fraction of simultaneously scheduled transmissions varies not only due to the up/down movements of the silencing
931: thresholds, but also due to the changing topology.
932:
933: We conclude from these plots that for small values of $\alpha$ (namely $\alpha=2.5$) the system converges to a
934: competitive equilibrium, where both networks simultaneously schedule a large fraction of their transmissions, and for
935: large values of $\alpha$ (namely $\alpha=3.5$ and $4.5$) the system converges to a near cooperative equilibrium, where
936: both networks schedule a small fraction of their transmissions.
937:
938: \begin{figure}
939: \centering
940: \includegraphics[width=210pt]{CSMAa2_5.eps}\\
941: \includegraphics[width=210pt]{CSMAa3_5.eps}\\
942: \includegraphics[width=210pt]{CSMAa4_5.eps}
943: \caption{Simulations of greedy algorithm under Carrier Sensing Multiple Access protocol}\label{fig:CSMA_sim}
944: \end{figure}
945:
946: \section{Conclusion}
947: This work studied spectrum sharing between wireless devices operating under a random access protocol. The crucial
948: assumption made was that nodes belonging to the same network or coalition cooperate with one another. Competition only
949: exists between nodes belonging to rival networks. It was found that cooperation between devices within each network
950: created the necessary incentive to prevent total anarchy. For pathloss exponents greater than four, we showed that
951: contrary to ones intuition, there can be a natural incentive for devices to cooperate to the extent that each occupies
952: only a fraction of the available bandwidth. Such results are optimistic and encouraging. We demonstrated via
953: simulations that it may be possible to extend them to more complex operating protocols such as those that employ
954: carrier-sensing to determine when the medium is free. More generally one wonders whether a multi-stage game capturing
955: the system dynamics under such a protocol can be formulated, and whether the desirable properties of the single-stage
956: game continue to hold. It would also be worthwhile investigating the incentives wireless links have to form coalitions,
957: as in this work it was in essence assumed that coalitions had been pre-determined.
958:
959: \section{Proofs}
960: \subsection{Theorem \ref{thm:op_access_prob}}
961: The limiting expression for the average throughput is
962: \begin{equation*}
963: {\mathbb E}{\overline R}(p_i,\beta_i) \rightarrow p_i \log(1+\beta_i) e^{-N_ip_i\beta_i^{2/\alpha}}.
964: \end{equation*}
965:
966: \noindent Given a $\beta_i$ there is a single maximum over $p_i$. By differentiation we have
967: \begin{equation*}
968: \frac{\partial {\mathbb E}{\overline R}}{\partial{p_i}} = \left(1-p_iN_i\beta_i^{2/\alpha}\right)
969: \log(1+\beta_i)e^{-N_ip_i\beta_i^{2/\alpha}},
970: \end{equation*}
971:
972: \noindent so $p_i^* = \min(1,1/N_i\beta_i^{2/\alpha})$. Thus
973: \begin{equation*}
974: {\mathbb E}{\overline R}(p_i^*,\beta_i) = \left\{
975: \begin{array}{ll}
976: \log(1+\beta_i)e^{-N_i\beta_i^{2/\alpha}}, &\hbox{$\beta_i\le N_i^{-2/\alpha}$}, \\
977: \dfrac{\log(1+\beta_i)e^{-1}}{N_i\beta_i^{2/\alpha}}, &\hbox{$\beta_i > N_i^{-2/\alpha}$}.
978: \end{array}
979: \right.
980: \end{equation*}
981:
982: \noindent Both of these functions have one maximum, but the maximum of the second function is always greater than the
983: maximum of the first as it represents the solution to the unconstrained problem
984: \begin{equation*}
985: \max_{p_i>0,\beta_i>0}{\mathbb E}{\overline R}(p_i,\beta_i),
986: \end{equation*}
987:
988: \noindent whereas the maximum of the first represents the solution to
989: \begin{equation*}
990: \max_{p_i=1,\beta_i>0}{\mathbb E}{\overline R}(p_i,\beta_i).
991: \end{equation*}
992:
993: \noindent Thus if the maximum of the second function occurs for $\beta_i > N_i^{-2/\alpha}$, it is the maximum of the
994: entire function, but if it occurs for $\beta_i \le N^{-2/\alpha}$, the maximum of the entire function is the maximum of
995: the first function over the domain $\beta_i \le N^{-2/\alpha}$. By differentiation we find the maximum of the second
996: function occurs at the unique solution of
997: \begin{equation*}
998: \frac{\alpha}{2} = (1+1/\beta_i)\log(1+\beta_i)
999: \end{equation*}
1000:
1001: \noindent which is $\beta_i={\Lambda^*}^{-\alpha/2}$. Thus if $N_i>\Lambda^*$, the solution is $p^* = \Lambda^*/N_i$
1002: and $\beta_i^* = {\Lambda^*}^{-\alpha/2}$. If $N_i\le \Lambda^*$ the solution is $p^* = 1$ and $\beta_i^*$ equal to the
1003: unique solution of
1004: \begin{equation}\label{eqn:2ndfn_max}
1005: \frac{\alpha}{2N_i\beta_i^{2/\alpha}} = (1+1/\beta_i)\log(1+\beta_i)
1006: \end{equation}
1007:
1008: \noindent or $N_i^{-2/\alpha}$, whichever is smaller. But as $(1+1/x)\log(1+x)$ is a monotonically increasing function,
1009: from the definition of $\Lambda^*$ we have
1010: \begin{equation*}
1011: \frac{\alpha}{2} \le \left(1+N_i^{\alpha/2}\right) \log \left( 1 + N_i^{-\alpha/2} \right)
1012: \end{equation*}
1013:
1014: \noindent whenever $N_i\le \Lambda^*$. Combining this with equation (\ref{eqn:2ndfn_max}) and using the monotonicity of
1015: $(1+1/x)\log(1+x)$ we see that the unique solution to equation (\ref{eqn:2ndfn_max}) is always smaller than
1016: $N_i^{-2/\alpha}$. This establishes the desired result.
1017:
1018: \subsection{Theorem \ref{thm:opt_p2}}
1019: First we show that any N.E. must lie on the boundary of the strategy space, i.e. $\Lambda_i=1$ for some $i$. The
1020: utility functions are smooth and continuous. Differentiating $U_1$ with respect to $\Lambda_1$ yields
1021: \begin{equation*}
1022: \frac{\partial U_1}{\partial \Lambda_1} = e^{-\Lambda_2/\Lambda_1}\left(
1023: \frac{\Lambda_2}{\Lambda_1}-\frac{\alpha}{2\left( 1 + \Lambda_1^{\alpha/2}\right)
1024: \log\left(1+\Lambda_1^{-\alpha/2}\right)} + 1\right)
1025: \end{equation*}
1026:
1027: \noindent Consider the function $f(x)=(1+x)\log(1+1/x)$. As $f(x)$ is monotonically decreasing for $x\ge 0$ and
1028: $\lim_{x\rightarrow \infty} f(x) = 1$ we have $f(x)> 1$ for all $x\ge 0$. Thus for $\alpha<4$, ${\partial
1029: U_1}/{\partial \Lambda_1}
1030: > 0$ whenever $\Lambda_1<\Lambda_2$ and similarly ${\partial U_2}/{\partial \Lambda_2} > 0$ whenever $\Lambda_2<\Lambda_1$. Thus for a N.E. to occur
1031: in the interior of the strategy space we must have both $\Lambda_1>\Lambda_2$ and $\Lambda_2>\Lambda_1$. As these
1032: conditions are mutually exclusive at least one of the constraints of the strategy space must be active at the N.E.. In
1033: essence each network is trying to set it's active density higher than the other's. Eventually at least one network maxs
1034: out.
1035:
1036: First consider the case where the solution to
1037: \begin{equation}\label{eqn:opt_p2_again}
1038: N_1 = x\left(\frac{\alpha}{2\left(1+{x}^{\alpha/2}\right)\log \left(1+{x}^{-\alpha/2}\right)} - 1\right).
1039: \end{equation}
1040:
1041: \noindent occurs for $x < N_2$. Suppose $\Lambda_2^* = N_2$. Then as $N_1 \le N_2$ we have $\Lambda_1\le \Lambda_2^*$,
1042: hence $\partial U_1/\partial \Lambda_1 > 0$ for all $\Lambda_1$ on the interior and hence $\Lambda_1^* = N_1$. Now the
1043: function $U_2(\Lambda_1^*,\Lambda_2)$ has a unique maximum for $\Lambda_2$. This maximum satisfies $\partial
1044: U_2(\Lambda_1^*,\Lambda_2)\partial \Lambda_2=0$, which is equation (\ref{eqn:opt_p2_again}) with $\Lambda_2^*$
1045: substituted for $x$. But the solution equation (\ref{eqn:opt_p2_again}) satisfies $x < N_2$, so $\Lambda_2^* < N_2$, a
1046: contradiction. Thus the constraint $\Lambda_2 \le N_2$ must be inactive. Suppose instead that the constraint
1047: $\Lambda_1^* = N_1$ is active. Then by the same arguments the unique $\Lambda_2^*$ satisfies equation
1048: (\ref{eqn:opt_p2_again}). This establishes the solution and it's uniqueness, for the first case.
1049:
1050: Second consider the case where the solution to equation (\ref{eqn:opt_p2_again}) occurs for $x\ge N_2$. Then it is
1051: straightforward to check using similar arguments above, that the unique solution satisfies
1052: $(\Lambda_1^*,\Lambda_2^*)=(N_1,N_2)$. This establishes the result.
1053:
1054: \subsection{Theorem \ref{thm:asymp_soln}}
1055: For $\alpha>2$ the solution to equation (\ref{eqn:opt_p2}) only goes to infinity for $N_1\rightarrow \infty$. In this
1056: limit $(1+{\Lambda_2^*}^{\alpha/2})\log(1+{\Lambda_2^*}^{-\alpha/2})\rightarrow 1$ and
1057: \begin{equation*}
1058: \Lambda_2^* \rightarrow \frac{2}{\alpha-2}\Lambda_1^*
1059: \end{equation*}
1060:
1061: \noindent if $N_1\le (\alpha/2-1)N_2$. Otherwise, $\Lambda_2^*=N_2$. Computing $p_i^*=\Lambda_i^*/N_i$ produces the
1062: stated result.
1063:
1064: \subsection{Theorem \ref{thm:asymp_util}}
1065: This follows by direct substitution.
1066:
1067: \subsection{Theorem \ref{thm:a_ge_4}}
1068: The proof of this result is more involved than the proof of theorem 2.3. There are three regimes.
1069:
1070: First consider the joint spread regime where $N_1 > \sqrt{\Lambda^*(\alpha/2)}$. We show that a N.E. cannot occur on
1071: the boundary of the strategy space. Suppose $\Lambda_1^*=N_1$. Then $\Lambda_1^*>\sqrt{\Lambda^*(\alpha/2)}$. As
1072: $\sqrt{\Lambda^*(\alpha/2)}$ is the solution to the equation
1073: \begin{equation*}
1074: \frac{\alpha}{2\left( 1 + {\sqrt{\Lambda^*(\alpha/2)}}^{\alpha/2}\right) \log \left( 1 +
1075: {\sqrt{\Lambda^*(\alpha/2)}}^{-\alpha/2} \right)}- 1 = 1.
1076: \end{equation*}
1077:
1078: \noindent and $(1+x)\log(1+1/x)$ is a monotonically decreasing function for $x>0$, we have
1079: \begin{equation*}
1080: \frac{\alpha}{2\left( 1 + {\Lambda_1^*}^{\alpha/2}\right) \log \left( 1 + {\Lambda_1^*}^{-\alpha/2} \right)} - 1 > 1.
1081: \end{equation*}
1082:
1083: \noindent If $\Lambda_1^*$ lies on the boundary of the strategy space then $\partial U_1(\Lambda_1^*,\Lambda_2^*)
1084: /\partial \Lambda_1 > 0$ which implies
1085: \begin{align*}
1086: \Lambda_2^* &> \Lambda_1^*\left( \frac{\alpha}{2\left( 1 + {\Lambda_1^*}^{\alpha/2}\right) \log \left( 1 +
1087: {\Lambda_1^*}^{-\alpha/2} \right)} -
1088: 1\right) \\
1089: &> \Lambda_1^* \\
1090: &> \sqrt{\Lambda^*(\alpha/2)}.
1091: \end{align*}
1092:
1093: \noindent This in turn implies
1094: \begin{equation*}
1095: \frac{\alpha}{2\left( 1 + {\Lambda_2^*}^{\alpha/2}\right) \log \left( 1 + {\Lambda_2^*}^{-\alpha/2} \right)} - 1 > 1.
1096: \end{equation*}
1097:
1098: \noindent At equilibrium $\partial U_2(\Lambda_1^*,\Lambda_2) / \partial \Lambda_2 \ge 0$ so
1099: \begin{align*}
1100: \Lambda_1^* &\ge \Lambda_2^*\left( \frac{\alpha}{2\left( 1 + {\Lambda_2^*}^{\alpha/2}\right) \log \left( 1 +
1101: {\Lambda_2^*}^{-\alpha/2} \right)} -
1102: 1\right) \\
1103: &> \Lambda_2^*,
1104: \end{align*}
1105:
1106: \noindent a contradiction. Thus $\Lambda_1 < N_1$. Now assume $\Lambda_2 = N_2$. By assumption $N_2 \ge N_1$ so
1107: $\Lambda_2^*
1108: > \sqrt{\Lambda^*(\alpha/2)}$. By repeating the same arguments we can generate the same style of contradiction and thus $\Lambda_2
1109: < N_2$. This establishes that a N.E. cannot occur on the boundary of the strategy space. In essence each network is
1110: trying to undercut the active density of the other. This drags the equilibrium away from the boundary.
1111:
1112: Now we establish any N.E. must be symmetric, i.e. $\Lambda_1^*=\Lambda_2^*$. Suppose a N.E. $(\Lambda_1^*,\Lambda_2^*)$
1113: with $\Lambda_1^* \neq \Lambda_2^*$ exists. Then as it must lie on the interior of the strategy space and as the
1114: utility functions are symmetric, $(\Lambda_2^*,\Lambda_1^*)$ must also be a N.E.. On the interior of the strategy space
1115: the N.E. criterion is $\partial U_1(\Lambda_1^*,\Lambda_2^*) /\partial \Lambda_1 = 0$ and so the function
1116: $\Lambda_1^*(\Lambda_2)$ is monotonically increasing in $\Lambda_2$. But this implies we cannot have N.E. at both
1117: $(\Lambda_1^*,\Lambda_2^*)$ and $(\Lambda_2^*,\Lambda_1^*)$, a contradiction. Thus $\Lambda_1^*=\Lambda_2^*$.
1118:
1119: By differentiating the utility functions this implies that at any N.E. $\Lambda_1^*$ satisfies
1120: \begin{equation*}
1121: \frac{\alpha}{2\left( 1 + {\Lambda_1^*}^{\alpha/2}\right) \log \left( 1 + {\Lambda_1^*}^{-\alpha/2} \right)} - 1 = 1
1122: \end{equation*}
1123:
1124: \noindent with $\Lambda_2^* = \Lambda_1^*$. But this is equivalent to $\Lambda_1^* = \sqrt{\Lambda^*(\alpha/2)}$. Thus
1125: the N.E. is unique and occurs at $(\Lambda_1^*,\Lambda_2^*) = (\sqrt{\Lambda^*(\alpha/2)},\sqrt{\Lambda^*(\alpha/2)})$.
1126:
1127: Next consider the partial spread and full spread regimes where $N_1 \le \sqrt{\Lambda^*(\alpha/2)}$. We first show that
1128: $\Lambda_1^* = N_1$. Suppose $\Lambda_1^* < N_1$. Then $\Lambda_1^* < \sqrt{\Lambda^*(\alpha/2)}$ which implies
1129: \begin{equation*}
1130: \frac{\alpha}{2\left( 1 + {\Lambda_1^*}^{\alpha/2}\right) \log \left( 1 + {\Lambda_1^*}^{-\alpha/2} \right)} - 1 < 1.
1131: \end{equation*}
1132:
1133: \noindent At equilibrium $\partial U_1/\partial \Lambda_1 = 0$ so
1134: \begin{align*}
1135: \Lambda_2^* &= \Lambda_1^*\left( \frac{\alpha}{2\left( 1 + {\Lambda_1^*}^{\alpha/2}\right) \log \left( 1 +
1136: {\Lambda_1^*}^{-\alpha/2} \right)} -
1137: 1\right) \\
1138: &< \Lambda_1^* \\
1139: &< \sqrt{\Lambda^*(\alpha/2)}.
1140: \end{align*}
1141:
1142: \noindent But this in turn implies
1143: \begin{equation*}
1144: \frac{\alpha}{2\left( 1 + {\Lambda_2^*}^{\alpha/2}\right) \log \left( 1 + {\Lambda_2^*}^{-\alpha/2} \right)} - 1 < 1
1145: \end{equation*}
1146:
1147: \noindent which in conjunction with the equilibrium condition $\partial U_2/\partial \Lambda_2 = 0$ implies
1148: $\Lambda_1^*<\Lambda_2^*$, a contradiction. Thus $\Lambda_1^* = N_1$. Now we can solve for $\Lambda_2^*$ to conclude
1149: that $\Lambda_2^*$ is the unique solution to equation (\ref{eqn:opt_p2}) or $N_2$, whichever is smaller. This concludes
1150: the proof.
1151:
1152: \subsection{Theorem \ref{thm:vr_soln}}
1153:
1154: We first tackle the full spread and partial spread regimes. It is shown in lemma \ref{lem:4} in the appendix that
1155: $\Lambda''(\alpha)$ is undefined for $\alpha\le 4$ and recall $\Lambda''(\alpha)$ is defined to be positive infinity
1156: for $\alpha > 4$. Consider the case where $N_1\le \Lambda''(\alpha)$. We show that $\Lambda_1^*=N_1$. Suppose the
1157: contrary, that $\Lambda_1^*<N_1$. Then $\Lambda_1^* < \Lambda''(\alpha)$. Define the function
1158: \begin{equation*}
1159: f(s_1,s_2) \triangleq \dfrac{\int_0^{\infty} e^{-(s_1+s_2)x^{2/\alpha}} \dfrac{dx}{1+x}}{s_1\int_0^{\infty}
1160: x^{2/\alpha}e^{-(s_1+s_2)x^{2/\alpha}} \dfrac{dx}{1+x} },
1161: \end{equation*}
1162:
1163: \noindent for $s_1$ and $s_2$ positive. In Lemma \ref{lem:1} in the appendix it is shown that $f(s,s)$ is a
1164: monotonically decreasing function in $s$. By rearranging equation (\ref{eqn:Lambda_doubleprime}) one can check that
1165: $f(\Lambda''(\alpha),\Lambda''(\alpha))=1$. Thus $f(\Lambda_1^*,\Lambda_1^*)>1$. For a given $\Lambda_2$ it is shown in
1166: Lemma \ref{lem:2} in the appendix that the utility function $U_1(\Lambda_1,\Lambda_2)$ is a smooth continuous function
1167: of $\Lambda_1$ with a unique maximum (and likewise for $U_2$ given $\Lambda_1$). As $\Lambda_1^*<N_1$ the maximum with
1168: respect to $\Lambda_1$ occurs at $\partial U_1(\Lambda_1^*,\Lambda_2)/\partial \Lambda_1 = 0$. By rearranging equation
1169: (\ref{eqn:vr_thm}) one can check that this condition is equivalent to $f(\Lambda_1^*,\Lambda_2)=1$. This means
1170: $f(\Lambda_1^*,\Lambda_2^*) < f(\Lambda_1^*,\Lambda_1^*)$. In lemma \ref{lem:2} we show that $f(s_1,s_2)$ is a
1171: monotonically increasing function in $s_2$ given $s_1$. Thus we have $\Lambda_2^*<\Lambda_1^*$ and so also
1172: $\Lambda_2^*<\Lambda''(\alpha)$ and $\Lambda_2^*<N_1$. Thus $f(\Lambda_2^*,\Lambda_2^*)>1$. As by assumption $N_1\le
1173: N_2$, we have $\Lambda_2^*<N_2$ and so the maximum of $U_2$ occurs at $\partial U_2(\Lambda_1,\Lambda_2^*)/\partial
1174: \Lambda_2 = 0$. This means $f(\Lambda_2^*,\Lambda_1^*) = 1 < f(\Lambda_2^*,\Lambda_2^*)$ which implies
1175: $\Lambda_1^*<\Lambda_2^*$. This is a contradiction. Thus we must have $\Lambda_1^* = N_1$ at a N.E.. By maximizing over
1176: $\Lambda_2$ via differentiation of $U_2$, we see that $\Lambda_2^*$ equals the solution of (\ref{eqn:vr_thm}) or $N_2$
1177: whichever is smaller. Lemma \ref{lem:2} establishes the solution of (\ref{eqn:vr_thm}) always exists and is unique.
1178:
1179: Now consider the case where $N_1 > \Lambda''(\alpha)$. We first show that a N.E. cannot occur on the boundary of the
1180: strategy space. Suppose $\Lambda_1^*=N_1$. Then $\Lambda_1 > \Lambda''(\alpha)$. This implies
1181: $f(\Lambda_1^*,\Lambda_1^*) <1$. As $\Lambda_1^*=N_1$ the maximum of $U_1$ occurs at $\partial
1182: U_1(\Lambda_1^*,\Lambda_2)/\partial \Lambda_1 \ge 0$ for a given $\Lambda_2$. This means $f(\Lambda_1^*,\Lambda_2^*)
1183: \ge 1 > f(\Lambda_1^*,\Lambda_1^*)$, thus $\Lambda_2^* > \Lambda_1^*$. We also then have $\Lambda_2^*>
1184: \Lambda''(\alpha)$. From the optimality condition for network 2 we then have $\partial
1185: U_2(\Lambda_1^*,\Lambda_2)/\partial \Lambda_2\ge 0$. This means $f(\Lambda_2^*,\Lambda_1^*) \ge 1 >
1186: f(\Lambda_2^*,\Lambda_2^*)$ which implies $\Lambda_1^*
1187: > \Lambda_2^*$. This is a contradiction. Thus we must have $\Lambda_1^* < N_1$. As $N_2\ge N_1>\Lambda''(\alpha)$ we can repeat the
1188: argument for $\Lambda_2^*$ to conclude that we must also have $\Lambda_2^*< N_2$. This proves a N.E. can only occur on
1189: the interior of the strategy space.
1190:
1191: Now we establish any N.E. must be symmetric, i.e. $\Lambda_1^*=\Lambda_2^*$. Suppose a N.E. $(\Lambda_1^*,\Lambda_2^*)$
1192: with $\Lambda_1^* \neq \Lambda_2^*$ exists. Then as it must lie on the interior of the strategy space and as the
1193: utility functions are symmetric, $(\Lambda_2^*,\Lambda_1^*)$ must also be a N.E.. On the interior of the strategy space
1194: the N.E. criterion is $\partial U_1(\Lambda_1^*,\Lambda_2^*) /\partial \Lambda_1 = 0$ and so the function
1195: $\Lambda_1^*(\Lambda_2)$ is monotonically increasing in $\Lambda_2$ by lemma \ref{lem:3}. But this implies we cannot
1196: have N.E. at both $(\Lambda_1^*,\Lambda_2^*)$ and $(\Lambda_2^*,\Lambda_1^*)$, a contradiction. Thus
1197: $\Lambda_1^*=\Lambda_2^*$.
1198:
1199: Finally one can verify that $\Lambda_1^*=\Lambda_2^*=\Lambda''(\alpha)$ is a N.E. by differentiating the utility
1200: functions. Thus the unique N.E. is $(\Lambda_1^*,\Lambda_2^*) = (\Lambda''(\alpha),\Lambda''(\alpha))$.
1201:
1202: In essence what is going on here is that in the absence of strategy space constraints, when
1203: $\Lambda_i<\Lambda''(\alpha)$, network $i$ wants to set $\Lambda_i>\Lambda_j$ and when $\Lambda_i>\Lambda''(\alpha)$
1204: network $i$ wants to set $\Lambda_i<\Lambda_j$. Thus the natural equilibrium is at $(\Lambda_1^*,\Lambda_2^*) =
1205: (\Lambda''(\alpha),\Lambda''(\alpha))$. The problem for $\alpha<4$ is $\Lambda''(\alpha)$ is infinite and the sparser
1206: network winds up maxing out at $\Lambda_1^*=N_1$. When $\alpha>4$ the function $\Lambda''(\alpha)$ is finite and it is
1207: possible to have $N_1>\Lambda''(\alpha)$, i.e. both networks have a sufficiently high density of nodes so as not to be
1208: constrained by the strategy space. In this case they get to set their access probabilities so as to achieve the natural
1209: equilibrium.
1210:
1211: \section{Appendix}
1212: \begin{lemma}\label{lem:1}
1213: The function $f(s,s)$ is monotonically decreasing in $s$.
1214: \end{lemma}
1215: \begin{proof}
1216: By changing variables we can rewrite $f(s,s)$ as
1217: \begin{equation*}
1218: f(s,s) \triangleq \dfrac{\int_0^{\infty}g_s(x)dx}{\int_0^{\infty}xg_s(x)dx}
1219: \end{equation*}
1220:
1221: \noindent where
1222: \begin{equation*}
1223: g_s(x) = \dfrac{x^{\alpha/2-1}e^{-2x}}{s^{\alpha/2}+x^{\alpha/2}}.
1224: \end{equation*}
1225:
1226: \noindent Choose a pair of values for $s_1$ and $s_2$ satisfying $0\le s_1\le s_2$ and then observe that for any
1227: $x_1\le x_2$ the following inequality holds
1228: \begin{equation*}
1229: \dfrac{g_{s_2}(x_2)}{g_{s_1}(x_2)} \ge \dfrac{g_{s_2}(x_1)}{g_{s_1}(x_1)}.
1230: \end{equation*}
1231:
1232: \noindent Thus
1233: \begin{align*}
1234: \int_0^{\infty}\int_0^{x_2} (x_2-x_1)g_{s_1}(x_1)g_{s_2}(x_2)dx_1dx_2 &= \int_0^{\infty}\int_{x_1}^{\infty} (x_2-x_1)g_{s_1}(x_1)g_{s_2}(x_2)dx_2dx_1 \\
1235: &= \int_0^{\infty}\int_{x_2}^{\infty} (x_1-x_2)g_{s_1}(x_2)g_{s_2}(x_1)dx_1dx_2 \\
1236: &\ge \int_0^{\infty}\int_{x_2}^{\infty} (x_1-x_2)g_{s_1}(x_1)g_{s_2}(x_2)dx_1dx_2.
1237: \end{align*}
1238:
1239: \noindent Then
1240: \begin{equation*}
1241: \int_0^{\infty}\int_0^{\infty} (x_2-x_1)g_{s_1}(x_1)g_{s_2}(x_2)dx_1dx_2 \ge 0
1242: \end{equation*}
1243:
1244: \noindent and so
1245: \begin{equation*}
1246: \int_0^{\infty}\int_0^{\infty} x_2g_{s_1}(x_1)g_{s_2}(x_2)dx_1dx_2 \ge \int_0^{\infty}\int_0^{\infty}
1247: x_1g_{s_1}(x_1)g_{s_2}(x_2)dx_1dx_2
1248: \end{equation*}
1249:
1250: \noindent which implies
1251: \begin{equation*}
1252: \int_0^{\infty} g_{s_1}(x)dx \int_0^{\infty} xg_{s_2}(x)dx \ge \int_0^{\infty} xg_{s_1}(x)dx \int_0^{\infty}
1253: g_{s_2}(x)dx
1254: \end{equation*}
1255:
1256: \noindent and thus $f(s_1,s_1)\ge f(s_2,s_2)$.
1257: \end{proof}
1258:
1259: \begin{lemma}\label{lem:2}
1260: The function
1261: \begin{equation*}
1262: U_i(\Lambda_1,\Lambda_2) = \Lambda_i\int_0^{\infty} e^{-(\Lambda_1+\Lambda_2)x^{2/\alpha}}\dfrac{dx}{1+x}
1263: \end{equation*}
1264:
1265: \noindent is smooth and continuous in $\Lambda_i$ with a unique maximum $\Lambda_i^*$.
1266: \end{lemma}
1267:
1268: \begin{proof}
1269: As the integral is well-defined for all positive $\Lambda_1$ and $\Lambda_2$, the function is smooth and continuous by
1270: inspection. To see that a unique maximum exists set the derivative to zero to obtain $f(\Lambda_i,\Lambda_j) = 1$. For
1271: fixed $\Lambda_j$ it is straightforward to show $f(\Lambda_i,\Lambda_j)$ is monotonically decreasing in $\Lambda_i$
1272: using arguments similar to those in lemma \ref{lem:1}. For $\Lambda_i\rightarrow 0$ we find
1273: $f(\Lambda_i,\Lambda_j)\rightarrow\infty$ and for $\Lambda_i \rightarrow \infty$ we find $f(\Lambda_i,\Lambda_j)
1274: \rightarrow 2/\alpha$ which is always less than 1 for $\alpha>2$. Thus there always exists a single $\Lambda_i^*$
1275: satisfying $f(\Lambda_i^*,\Lambda_j)=1$ and hence a unique maximum always exists.
1276: \end{proof}
1277:
1278: \begin{lemma}\label{lem:3}
1279: The function $f(s_1,s_2)$ is monotonically increasing in $s_2$ given $s_1$.
1280: \end{lemma}
1281:
1282: \begin{proof}
1283: The proof mirrors that of lemma \ref{lem:1}, the only difference being that now
1284: \begin{equation*}
1285: \dfrac{g_{s_2}(x_2)}{g_{s_1}(x_2)} \le \dfrac{g_{s_2}(x_1)}{g_{s_1}(x_1)},
1286: \end{equation*}
1287:
1288: \noindent i.e. the inequality goes the other way. We omit the details for brevity.
1289: \end{proof}
1290:
1291: \begin{lemma}\label{lem:4}
1292: The function $\Lambda''(\alpha)$ is undefined for $\alpha \le 4$ and uniquely defined for $\alpha > 4$.
1293: \end{lemma}
1294:
1295: \begin{proof}
1296: $\Lambda''(\alpha)$ is the solution to $f(\Lambda''(\alpha),\Lambda''(\alpha))=1$. The function $f(s,s)$ is a
1297: monotonically decreasing in $s$ by lemma \ref{lem:1}. By taking $s\rightarrow 0$ we find $f(s,s)\rightarrow \infty$ and
1298: by taking $s\rightarrow \infty$ we find $f(s,s)\rightarrow 4/\alpha$. Thus when $\alpha\le 4$ there is no $s$ for which
1299: $f(s,s)=1$ and hence $\Lambda''(\alpha)$ is undefined. When $\alpha > 4$ there is a single $s$ at which $f(s,s)$
1300: crosses the value 1 and hence $\Lambda''(\alpha)$ is uniquely defined.
1301: \end{proof}
1302:
1303: \begin{thebibliography}{15}
1304:
1305: \bibitem{spectrum3} L. Berlemann, G. R. Hiertz, B. H. Walke, S. Mangold, ``Radio resource sharing games: enabling QoS
1306: support in unlicensed bands,'' \emph{IEEE Network}, vol. 19, iss. 4, pp. 59-65, July-Aug. 2005.
1307:
1308: \bibitem{gg2}
1309: S.T. Chung, S. Kim, J. Lee, J.M. Cioffi, ``A game-theoretic approach to power allocation in frequency-selective
1310: Gaussian interference channels,'' \emph{Proc. of IEEE International Symposium on Information Theory}, pp. 316-316, July
1311: 2003.
1312:
1313: \bibitem{gg3}
1314: R. Etkin, A. Parekh, D. Tse, ``Spectrum sharing for unlicensed bands,'' \emph{IEEE J. Sel. Areas Comm.}, vol. 25, is.
1315: 3, pp. 517-528, Apr. 2007.
1316:
1317: \bibitem{spectrum2}
1318: M. F\'{e}legyh\'{a}zi, M. Cagalj, S. S. Bidokhti, J.-P. Hubaux, ``Non-cooperative multi-radio channel allocation in
1319: wireless networks,'' \emph{Proc. INFOCOM}, pp. 1442-1450, May 2007.
1320:
1321: \bibitem{spectrum4}
1322: M. M. Halld\'{o}rsson, J. Y. Halpern, Li (Erran) Li, Vahab S. Mirrokni, ``On spectrum sharing games,'' \emph{Proc. of
1323: Annual ACM Symposium on Principles of Distributed Computing}, pp. 107-114, 2004.
1324:
1325: \bibitem{jindal1}
1326: N. Jindal, J. G. Andrews, S. Weber, ''Optimizing the SIR operating point of Spatial Networks,'' \emph{Workshop on
1327: Information Theory and its Applications, U.C. San Diego}, available at arXiv:cd.IT/0702030v1, Feb 2007.
1328:
1329: \bibitem{jindal2}
1330: N. Jindal, J. G. Andrews, S. Weber, ''Bandwidth partitioning in Decentralized Wireless Networks,'' submitted to
1331: \emph{IEEE Trans. Wireless Comm.}, Nov. 2007.
1332:
1333: \bibitem{jindal3}
1334: S. Weber, J. G. Andrews, ''The Effect of Fading, Channel Inversion, and Threshold Scheduling on {\it Ad Hoc}
1335: Networks,'' \emph{IEEE Trans Info. Theory}, vol. 53, no. 11, Nov. 2007.
1336:
1337: \bibitem{jindal4}
1338: N. Jindal, S. Weber, J. G. Andrews, ``Fractional Power Control for Decentralized Wireless Networks,'' submitted to
1339: \emph{IEEE Trans. Wireless Comm.}, Dec. 2007.
1340:
1341: \bibitem{baccelli}
1342: F. Baccelli, B. Blaszczyszyn, P. M\"{u}hlethaler, ``An Aloha Protocol for Multihop Mobile Wireless Networks,''
1343: \emph{IEEE Trans. Info. Theory}, vol. 52, pp. 421-436, Feb. 2006.
1344:
1345: \bibitem{ra2}
1346: A.B. MacKenzie, S.B. Wicker, ``Selfish users in Aloha: a game-theoretic approach,'' \emph{Proc. of IEEE Vehicular
1347: Technology Conference}, vol. 3, pp. 1354-1357, Oct. 2001.
1348:
1349: \bibitem{spectrum5}
1350: M. H. Manshaei, M. F\'{e}legyh\'{a}zi, J. Freudiger, J.-P. Hubaux, P. Marbach, ``Spectrum Sharing Games of Network
1351: Operators and Cognitive Radios,'' \emph{Cognitive Wireless Networks: Concepts, Methodologies and Visions}, Springer,
1352: 2007.
1353:
1354: \bibitem{coal1}
1355: S. Mathur, L. Sankaranarayanan, N. B. Mandayam, ``Coalitional games in receiver cooperation for spectrum sharing,''
1356: \emph{Proc. of IEEE Inform. Sci. and Systems}, pp. 949-954, March 2006.
1357:
1358: \bibitem{ra1}
1359: O. Simeone, Y. Bar-Ness, ``A game-theoretic view on the interference channel with random access,'' \emph{Proc. of IEEE
1360: New Frontiers in Dynamic Spectrum Access Networks}, pp. 13-21, April 2007.
1361:
1362: \bibitem{gg1}
1363: Wei Yu, G. Ginis, J.M. Cioffi, ``Distributed multiuser power control for digital subscriber lines,'' \emph{IEEE J. Sel.
1364: Areas Comm.} vol. 20, no. 5, June 2002.
1365:
1366: \bibitem{spectrum1}
1367: F. Zuyuan, B. Bensaou,``Fair bandwidth sharing algorithms based on game theory frameworks for wireless ad-hoc
1368: networks,'' \emph{Proc. 23rd Annual Joint Conference of the IEEE Computer and Communications Societies (INFOCOM)}, vol.
1369: 2, pp. 1284-1295, Mar. 2004.
1370:
1371: \end{thebibliography}
1372:
1373: \end{document}
1374: