1: \documentclass[9pt,conference]{IEEEtran}
2: %\documentclass{article}
3: %\long\def\comment#1{}
4:
5: \usepackage[dvips]{graphics}
6: \usepackage[dvips]{graphicx}
7: \usepackage{amsmath,epsfig,epsf,psfrag,amssymb,amsfonts,latexsym}
8: \usepackage{verbatim}
9: %\usepackage{spconf}
10: \usepackage[mathscr]{eucal}
11:
12: \input macros
13: %\input setup
14: \input labelfig.tex
15:
16: \newtheorem{proposition}{Proposition}
17: \newtheorem{definition}{Definition}
18: \newtheorem{theorem}{Theorem}
19: \newtheorem{corollary}{Corollary}
20: \newtheorem{model}{Model}
21:
22:
23: \newcommand{\bth}{\mbox{${\bf \theta}$}}
24: \newcommand{\bzero}{\mbox{${\bf 0}$}}
25: \newcommand{\bS}{\mbox{$\mathcal{S}$}}
26: \newcommand{\bDc}{\mbox{${\bf \Dc}$}}
27: \newcommand{\bRc}{\mbox{${\bf \Rc}$}}
28: \newcommand{\bHc}{\mbox{${\bf \Hc}$}}
29: \newcommand{\bTc}{\mbox{${\bf \Tc}$}}
30: \newcommand{\bFc}{\mbox{${\bf \Fc}$}}
31: \newcommand{\bPc}{\mbox{${\bf \Pc}$}}
32: \newcommand{\bGc}{\mbox{${\bf \Gc}$}}
33:
34: \def\nn{\nonumber}
35: \def\defeq{\stackrel{\Delta}{=}}
36: \def\scalefig#1{\epsfxsize #1\textwidth}
37: \def\Ebb{{\mathbb E}}
38: \def\Rbb{{\mathbb R}}
39:
40:
41: \newcommand{\SNR}{\mbox{SNR}}
42:
43: \newcommand{\Xmsc}{\mathscr{X}}
44:
45: \newcommand{\beq}{\begin{equation}}
46: \newcommand{\eeq}{\end{equation}}
47:
48: \newcommand{\Var}{\mbox{${\mbox{Var}}$}}
49:
50:
51: \IEEEoverridecommandlockouts
52: \flushbottom
53: \title{Sensor Configuration and Activation for Field Detection\\ in Large Sensor Arrays}
54: \author{\authorblockN{ Youngchul Sung and Lang Tong\thanks{This work was supported in part by
55: the Multidisciplinary University Research Initiative (MURI) under
56: the Office of Naval Research Contract N00014-00-1-0564. Prepared
57: through collaborative participation in the Communications and
58: Networks Consortium sponsored by the U.~S. Army Research
59: Laboratory under the Collaborative Technology Alliance Program,
60: Cooperative Agreement DAAD19-01-2-0011.}}
61: \authorblockA{School of Electrical and Computer Engineering\\
62: Cornell University\\
63: Ithaca, NY 14850, USA \\
64: Email:\{ys87,ltong\}@ece.cornell.edu}
65: \and
66: \authorblockN{H. Vincent Poor\thanks{The work of H. V. Poor was
67: supported in part by the Office of Naval Research under Grant
68: N00014-03-1-0102.}}
69: \authorblockA{ Dept. of Electrical Engineering\\
70: Princeton University \\
71: Princeton, NJ 08544 \\
72: Email:poor@princeton.edu}}
73:
74:
75:
76: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77: \begin{document}
78: \maketitle
79: %\ninept
80:
81:
82: {\footnotesize
83: \begin{abstract} The problems of sensor configuration and activation for
84: the detection of correlated random fields using large sensor
85: arrays are considered. Using results that characterize the
86: large-array performance of sensor networks in this application,
87: the detection capabilities of different sensor configurations are
88: analyzed and compared. The dependence of the optimal choice of
89: configuration on parameters such as sensor signal-to-noise ratio
90: (SNR), field correlation, etc., is examined, yielding insights
91: into the most effective choices for sensor selection and activation in various operating regimes.
92: \end{abstract} }
93: \vspace{-0.5em}
94: \section{Introduction}
95: \label{sec:intro}
96:
97: The main design criteria for sensor networks are the performance
98: in the specific task and the energy efficiency of the network. In
99: this paper, we consider optimal sensor configuration and selection
100: for densely deployed sensor networks for the detection of
101: correlated random fields. An example in which the problem of such
102: sensor selection arises is Sensor Network with Mobile Access (SENMA), as
103: shown in Fig.~\ref{fig:senma}, where a mobile access point
104: collects sensor data controlling sensor transmissions in the
105: reachback channel.
106: \begin{figure}[hbtp]
107: \centerline{
108: \begin{psfrags}
109: \psfrag{A}[c]{Access point} \psfrag{B}[l]{Sensor}
110: \psfrag{C}[c]{Sensor Network}
111: \scalefig{0.35}\epsfbox{figures/senmamin.eps}
112: \end{psfrags}
113: } \caption{Sensor network with mobile access point}
114: \label{fig:senma}
115: \end{figure}
116: To maximize the energy efficiency of such a network, one should
117: judiciously select and activate sensors to satisfy the desired
118: detection performance with the minimum amount of sensor data since
119: the number of activated sensors is directly related to the energy
120: consumption of the entire network.
121:
122: To simplify the problem for analysis, we focus on a 1-dimensional
123: space, and investigate how various parameters such as the field
124: correlation, signal-to-noise ratio (SNR), etc., affect the optimal
125: configuration
126: for different sensor schedules.
127: \begin{comment}
128: \begin{figure}[htbp]
129: \centerline{ {
130: \begin{psfrags}
131: \psfrag{0}[c]{ $0$}
132: \psfrag{L}[c]{ $x_n$}
133: \psfrag{a}[r]{ $s(x)$}
134: \psfrag{d}[c]{ $\Delta_{i}$}
135: \psfrag{xi}[c]{ $x_i$}
136: \psfrag{xi1}[c]{ $x_{i+1}$}
137: \scalefig{0.3}\epsfbox{figures/sample_ra.eps}
138: \end{psfrags}
139: } } \caption{Signal and sensor location} \label{fig:sensorlocations}
140: \end{figure}
141: \end{comment}
142: Specifically, we assume that the signal field $s(x)$ is the
143: stationary solution of the stochastic diffusion equation \cite{Cox&Miller:book}:
144: \begin{equation} \label{eq:diffusioneq}
145: \frac{ds(x)}{dx}= - A s(x)+ B u(x), ~~x\ge 0,
146: \end{equation}
147: where $A \ge 0$ and $B$ are known, and the initial condition is
148: given by $s(0) \sim \Nc(0,\Pi_0)$. Here, $x$ denotes position
149: along the linear axis of the sensor array. The process noise
150: $u(x)$ is a zero-mean white Gaussian process, independent of
151: both sensor measurement noises $\{w_i\}$ and the initial state $s(0)$.
152: We assume that each activated sensor takes a
153: measurement of the field at its location, and subsequently
154: transmits the data to the collector or fusion center\footnote{
155: We will not focus on local quantization of $y_i$ at the sensor
156: level here, nor will we consider the transmission error to the
157: fusion center. These are important design issues that must be
158: treated separately.}. The observation $y_i$ from the activated
159: sensor $i$ located at $x_i$ ($x_i < x_{i+1}$) is governed by the
160: following statistical hypotheses
161: \begin{equation} \label{eq:hypothesisscalar}
162: \begin{array}{lcl}
163: H_0 &: & y_i = w_i, ~~~~i=1,2,\cdots, n,\\
164: H_1 &: & y_i = s_i+ w_i, \\
165: \end{array}
166: \end{equation}
167: where $\{w_i\}$ are $\Nc(0,\sigma^2)$ measurement noises, with
168: known $\sigma^2$ and independent from sensor to sensor, and
169: the signal sample $s_i \defeq s(x_i)$. The dynamics of the collected signal samples $\{s_i\}$ are given by
170: \begin{eqnarray}
171: s_{i+1} &=& a_i s_i + u_i, \label{eq:statespacemodelgeneral}\\
172: a_i &=& e^{-A\Delta_{i}}, \label{eq:statespaceai}
173: \end{eqnarray}
174: where $\Delta_{i} \defeq |x_{i+1}-x_i|$
175: and $u_i \sim \Nc(0, \Pi_0(1-a_i^2))$. A similar model was derived
176: in \cite{Micheli&Jordan:02MTNS}.
177:
178: Note that $0 \le a_i \le 1$ for
179: $ 0 \le A \le \infty$ and $a_i$ determines the amount of
180: correlation between sample $s_i$ and $s_{i+1}$; $a_i=0$ implies
181: that two samples are independent while for perfectly correlated
182: signal samples we have $a_i=1$ . By the stationarity, $\Ebb
183: \{s_i^2\} = \Pi_0$ for all $i$, and the SNR for the observations
184: is given by ${\Pi_0}/{\sigma^2}$.
185:
186:
187:
188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189: \subsection{Summary of Results}
190:
191: We adopt the Neyman-Pearson formulation of fixing
192: the detector size $\alpha$ and minimizing the miss probability.
193: The miss probability $P_M(\Xmsc,n;\alpha,\mbox{SNR})$ is a
194: function of the number, $n$, and locations, $\Xmsc
195: \defeq \{x_1,\cdots,x_n\}$, of the activated sensors as well as detector size
196: $\alpha$ and SNR. Usually, the miss probability decreases
197: exponentially as $n$ increases and the error exponent is defined
198: as the decay rate
199: \begin{equation}\label{eq:exp}
200: K_\alpha(\Xmsc;\SNR)=\lim_{n\rightarrow \infty} \frac{1}{n}\log
201: P_M(\Xmsc,n;\alpha,\mbox{SNR}).
202: \end{equation}
203: The error exponent is a good performance index since it gives an
204: estimate of the number of samples required for a given detection
205: performance; faster decay rate implies that fewer samples are
206: needed for a given miss probability. Hence, the energy efficient
207: configuration for activation can be formulated to find the optimal $\Xmsc$
208: (where data should be collected) maximizing the error exponent
209: when the sample size is sufficiently large.
210:
211: Based on our
212: previous results on the behavior of the error exponent for the
213: detection of correlated random
214: fields \cite{Sung&Tong&Poor:04ITsub}, we examine several strategies
215: for sensor configuration for the testing of $H_1$ versus $H_0$, and
216: propose guidelines for the optimal configuration for different
217: operating regimes. Specifically, we consider uniform configuration,
218: periodic clustering, and periodic configuration with arbitrary sensor
219: locations within a period. We show that the optimal configuration
220: is a function of the field correlation and the SNR of the
221: observations. For uniform configuration, the optimal strategy
222: is to cover the entire signal field with the activated sensors for
223: SNR $>1$. For SNR $<1$, on the other hand, there exists an optimal
224: spacing between the activated sensors. We also derive the error
225: exponents of periodic clustering and arbitrary periodic
226: configurations. Depending on the field correlation and SNR, the
227: periodic clustering outperforms uniform configuration. Furthermore,
228: there exists an optimal cluster size for intermediate values of
229: field correlation. The closed-form error exponent obtained for the
230: vector state-space model explains the transitory error behavior
231: for different sensor configurations as the field correlation
232: changes. It is seen that the optimal periodic configuration is
233: either periodic clustering or uniform configuration for highly correlated or
234: almost independent signal fields.
235:
236:
237: \subsection{Related Work}
238:
239: The detection of Gauss-Markov processes in Gaussian noise is a
240: classical problem. See \cite{Kailath&Poor:98IT} and references
241: therein. Our work is based on the large deviations results in
242: \cite{Sung&Tong&Poor:04ITsub}, where the closed-form error
243: exponent was derived for the Neyman-Pearson detection, with a
244: fixed size, of correlated random fields using the innovations
245: approach for the log-likelihood ratio (LLR) \cite{Schweppe:65IT}.
246: There is an extensive literature on the large deviation approach
247: to the detection of Gauss-Markov
248: processes \cite{Benitz&Bucklew:90IT}-\cite{
249: %,Bahr:90IT,Bahr&Bucklew:90SP,Barone&Gigli&Piccioni:95IT,Bryc&Smolenski:93SPL,Bercu&Gamboa&Rouault:97SPA,
250: Luschgy:94SJS}.
251: The application of the large deviations principle (LDP) to
252: sensor networks has been considered by other authors as well. The
253: sensor configuration problem can be viewed as a sampling problem. To
254: this end, Bahr and Bucklew \cite{Bahr&Bucklew:90SP} optimized the
255: exponent numerically under a Bayesian formulation. For a specific
256: signal model (low pass signal in colored noise), they showed that
257: the optimal sampling depends on SNR, which we also show in this
258: paper in a different setting. Chamberland and Veeravalli have
259: also considered the detection of correlated fields in large sensor
260: networks under the formulation of LDP and a fixed
261: threshold for the LLR test with the focus on detection performance
262: under power constraint \cite{Chamberland&Veeravalli:04ITWS}.
263:
264:
265:
266: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267: \section{Preliminaries: Error Exponent and Properties}
268: \label{sec:preliminaries}
269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
270:
271: In this section we briefly present previous results
272: \cite{Sung&Tong&Poor:04ITsub} relevant to our sensor configuration
273: problem.
274: The error exponent for the Neyman-Pearson detection of the hypotheses
275: (\ref{eq:hypothesisscalar})
276: with a fixed size $\alpha \in (0,1)$ and uniformly configured
277: sensors with spacing $\Delta$ (i.e., $\Xmsc = \{
278: (i-1)\Delta\}_{i=1}^n$) is given by
279: \begin{equation}
280: K_\alpha(\Xmsc;\SNR)= -\frac{1}{2} \log\frac{\sigma^2}{R_{e}} +
281: \frac{1}{2} \frac{\tilde{R}_{e}}{R_{e}}
282: - \frac{1}{2}, \label{eq:errorexponentscalar}
283: \end{equation}
284: independently of the value of $\alpha$,
285: where $R_{e}$
286: and $\tilde{R}_{e}$ are the steady-state
287: variances of the innovations process of $\{y_i\}$ calculated under $H_1$ and $H_0$, respectively.
288: The closed-form formula (\ref{eq:errorexponentscalar}) is obtained
289: via the innovations representation of the log-likelihood ratio
290: \cite{Schweppe:65IT}, and enables us to further investigate the
291: properties of the error exponent with respect to (w.r.t.)
292: parameters such as the correlation strength and SNR. (See
293: \cite{Sung&Tong&Poor:04ITsub} for more detail.) Here, we note that
294: the error exponent for the miss probability with a fixed size does
295: not depend on the value $\alpha$ of the size. Thus, the error
296: exponent depends only on $\Xmsc$ and SNR. (For notational
297: convenience, we use $K$ for the error exponent unless the
298: arguments are needed.)
299:
300:
301: We now describe the basic properties of the error exponent
302: starting from the extreme correlation cases.
303:
304: \vspace{0.3em}
305: \begin{theorem}[Extreme correlations] \label{theo:extremecor}
306: The error exponent $K$ is a continuous function of the correlation
307: coefficient $a \defeq e^{-A\Delta}$ for a given
308: SNR. Furthermore,
309: \begin{itemize}
310: \item[(i)] for i.i.d. observations ($a=0$) the error exponent $K$ reduces to the Kullback-Leibler
311: information $D(p_0||p_1)$ where $p_0 \sim \Nc(0,\sigma^2)$ and $p_1
312: \sim \Nc(0, \Pi_0+\sigma^2)$;
313: \item[(ii)] for the perfectly correlated signal ($a=1$) the error exponent $K$ is zero for any SNR,
314: and the miss probability decays to zero with
315: $\Theta(\frac{1}{\sqrt{n}})$.
316: \end{itemize}
317: \end{theorem}
318:
319: \vspace{0.3em}
320: The above theorem reduces to the Stein's lemma
321: for the i.i.d. case.
322: For the perfectly correlated case ($a=1$), on the other hand, the miss probability does not decay exponentially;
323: rather it
324: decays in polynomial order $n^{-1/2}$.
325:
326: The error behavior for intermediate values of correlation is
327: summarized by the following theorem, and shows distinct characteristics for different SNR regimes.
328:
329: \vspace{0.3em}
330: \begin{theorem}[$K$ vs. correlation] \label{theo:etavscorrelation}
331: \begin{itemize}
332: \item[(i)] For SNR $> 1$, $K$ decreases monotonically as the
333: correlation increases (i.e. $a \uparrow 1$); \item[(ii)] For SNR
334: $< 1$, there exists a non-zero correlation value $a^*$ that
335: achieves the maximal $K$, and $a^*$ is given by the solution of
336: the following equation.
337: \begin{equation} \label{eq:optimalam}
338: [1+a^2+\Gamma(1-a^2)]^2-2(r_e+\frac{a^4}{r_e})=0,
339: \end{equation}
340: where $r_e = R_{e}/\sigma^2$. Furthermore, $a^*$ converges to one
341: as SNR approaches zero.
342: \end{itemize}
343: \end{theorem}
344: Hence, an i.i.d. signal gives the best detection performance for
345: a given SNR $> 1$. The intuition behind this result is that the
346: signal component in the observation is strong at high SNR, and the
347: new information contained in the observation provides more benefit
348: to the detector than the noise averaging effect present for
349: correlated observations. For SNR $< 1$, on the other hand, the
350: error exponent does not decrease
351: monotonically as correlation becomes strong, and there exists an optimal correlation. This is because at
352: low SNR
353: the signal component is weak in the observation and correlation between signal samples
354: provides a
355: noise averaging effect. This noise averaging will become evident in Section \ref{subsec:periodiccluster}.
356: \begin{figure}[htbp]
357: \centerline{ \SetLabels
358: \L(0.25*-0.1) (a) \\
359: \L(0.75*-0.1) (b) \\
360: \endSetLabels
361: \leavevmode
362: %\ShowGrid
363: \strut\AffixLabels{
364: \scalefig{0.23}\epsfbox{figures/plot_sep14etavsasnrhigh.eps}
365: \scalefig{0.234}\epsfbox{figures/plot_sep14etavsasnrlow.eps} } }
366: \vspace{0.5cm} \caption{$K$ vs. correlation coefficient $a$: (a)
367: SNR = 10 dB (b) SNR= -3, -6, -9 dB} \label{fig:Kvsa}
368: \end{figure}
369: Fig. \ref{fig:Kvsa} shows the error exponent as a function of the
370: correlation coefficient
371: $a$ for several values of SNR. Two plots clearly show the different error behaviors as a function
372: of correlation
373: in the high and low SNR regimes. Unit SNR is a transition point between two different
374: behavioral regimes of the error exponent as a function of correlation strength.
375:
376: The error exponent is also a function of SNR. This aspect of the
377: behavior of the error exponent is given by the following theorem.
378:
379: \vspace{0.3em}
380: \begin{theorem}[$K$ vs. SNR] \label{theo:etavsSNR}
381: The error exponent $K$ is monotone increasing as SNR increases
382: for a given correlation coefficient $0 \le a <1$. Moreover, at
383: high SNR the error exponent $K$ increases linearly with respect
384: to $\frac{1}{2}\log \mbox{SNR}$.
385: \end{theorem}
386:
387:
388: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
389: \section{Sensor Configuration}
390: \label{sec:optimalscheduling}
391: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
392:
393: In this section, we investigate several sensor configurations, and analyze
394: the corresponding detection performance via the error exponent.
395: We also provide the closed-form error-exponent for several interesting cases by extending the
396: results in the previous section. Specifically, we consider uniform
397: configuration, clustering, and periodic configuration with arbitrary
398: locations within a spatial period as described in Fig.
399: \ref{fig:schedulingprotocols}. We provide the optimal
400: configuration for uniform configuration for the detection of
401: stationary correlated fields, and investigate the benefit of other
402: configurations.
403: \begin{figure}[hbtp]
404: \centerline{
405: \begin{psfrags}
406: \psfrag{a}[c]{(a)} \psfrag{b}[c]{(b)} \psfrag{c}[c]{(c)}
407: \psfrag{d}[c]{(d)} \psfrag{Delta}[c]{$\Delta$}
408: \psfrag{MD1}[c]{$M\Delta$} \psfrag{MD2}[c]{$M\Delta$}
409: \psfrag{Xc}[l]{$\Xc$} \psfrag{sensor}[l]{\textsf{Sensor}}
410: \scalefig{0.350}\epsfbox{figures/protocolsIPSN.eps}
411: \end{psfrags}
412: } \caption{Configurations for $n$ sensor activations:
413: (a) uniform configuration (b) periodic clustering (c) arbitrary
414: periodic configuration} \label{fig:schedulingprotocols}
415: \end{figure}
416:
417:
418: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
419: \vspace{-0.5em}
420: \subsection{Uniform Configuration} \label{subsec:uniformschedule}
421:
422: For uniform configuration with spacing $\Delta$ between neighboring
423: sensors, the data model is described by the state-space model
424: (\ref{eq:statespacemodelgeneral}) with
425: \[
426: a_1=\cdots=a_n=a = e^{-A\Delta},
427: \]
428: and the results in Section \ref{sec:preliminaries} are
429: directly applicable. The key connection between sensor configuration
430: and detector performance is given by the correlation coefficient
431: $a$. First, we consider $\SNR
432: > 1$. In this case, by Theorem \ref{theo:etavscorrelation} {\em (i)}, the error exponent decreases
433: monotonically as $a$ increases, i.e., the spacing $\Delta$ decreases for a given field diffusion rate $A$.
434: Hence, when the support of the signal field $\Sc$ is finite and
435: $n$ sensors are planned to be activated in the field,
436: the optimal uniform scheme is to distribute the $n$ activated sensors to cover all the signal field, which makes
437: the observations least correlated; localizing all the scheduled sensors in a subregion of the stationary signal field
438: is not optimal. For $\SNR < 1$, on the other hand, the optimal spacing $\Delta^*$ for an infinite (in size) signal field
439: is given by
440: \begin{equation}
441: \Delta^* = - \log \frac{a^*}{A},
442: \end{equation}
443: where $a^*$ is given by the solution of (\ref{eq:optimalam}).
444: $\Delta^*$ is finite for any SNR strictly less than one since the diffusion coefficient
445: $A < 0$ and $a^* > 0$ for any SNR $< 1$. The optimal spacing as a function of SNR is shown for $A=1$
446: in Fig. \ref{fig:lowSNRoptimaldistance}.
447: \begin{figure}[hbtp]
448: \centerline{
449: \begin{psfrags}
450: \scalefig{0.30}\epsfbox{figures/plot_sep14deltaopt.eps}
451: \end{psfrags}
452: } \caption{Optimal spacing between sensors for infinite signal
453: field ( SNR $<1$, $A=1$)} \label{fig:lowSNRoptimaldistance}
454: \end{figure}
455:
456: For a finite signal field $\Sc$ with $n$ scheduled sensors,
457: $\Delta^*$ is still optimal among the class of uniform
458: configurations if $n\Delta^* < |\Sc|$, where $|\Sc|$ is the
459: spatial duration of the signal field. In this case, the sensor
460: field does not need to cover the entire signal field. However, if
461: $n\Delta^*
462: > |\Sc|$, $\Delta^*$ may no longer be the optimal spacing. As shown in Fig.
463: \ref{fig:Kvsa}, the error exponent decreases when the spacing
464: $\Delta$ deviates from $\Delta^*$. Hence, activating sensors fewer
465: than $\bar{n}
466: \defeq \lfloor \frac{|\Sc|}{\Delta^*}\rfloor$ with spacing larger
467: than $\Delta^*$ always
468: gives a worse performance than $\bar{n}$ sensors with
469: spacing $\Delta^*$. However, this may not be the case for activating more
470: sensors than $\bar{n}$ (up to $n$) by reducing the spacing from
471: $\Delta^*$. Even if the error exponent decreases by reducing the
472: spacing, more sensors are activated over the signal field.
473: Therefore, better performance is possible for the latter case
474: since the product of the error exponent and the number of samples
475: determines the miss probability approximately. Similar situation
476: also occurs at SNR $>1$ for finite signal field. Note that the
477: error exponent increases as the correlation decreases. (See Fig.
478: \ref{fig:Kvsa}.) Thus, by spreading the activated sensors with a
479: reduced number of activated sensors in the signal field, sensor
480: data become less correlated and the slope of error decay becomes
481: larger at a cost of reducing the number of observations. However,
482: the increase in the error exponent is not large enough to
483: compensate for the loss in the number of sensors in the field.
484: Fig. \ref{fig:finitefieldSNR10dB} shows the error exponent as a
485: function of the number of scheduled sensors in $\Sc$ ( $|\Sc|=1$)
486: at 10 dB SNR for several different diffusion rates. The dashed
487: line shows the decay of $n^{-1}$ for which the performance loss by
488: the decrease in the number of scheduled sensors is exactly
489: balanced. We observe that the decay of the actual error exponent
490: is slower than $n^{-1}$. Hence, when the maximum number of
491: available sensors in a finite signal field case is
492: $n$, the optimal configuration is to activate all $n$ sensors
493: covering the entire field with maximal spacing.
494: \begin{figure}[htbp]
495: \centerline{ {
496: \begin{psfrags}
497: \scalefig{0.3}\epsfbox{figures/plot_nov3finiteKvsnSNR10dB.eps}
498: \end{psfrags}
499: } } \caption{Error exponent vs. $n$ for a fixed signal field
500: ($|\Sc|=1$)} \label{fig:finitefieldSNR10dB}
501: \end{figure}
502:
503:
504: Another interesting fact about the finite signal field is the
505: asymptotic behavior when the number of sensors increases without
506: bound. In this case, the correlation coefficient converges to one,
507: i.e.,
508: \begin{equation} \label{eq:sensorplaceinfinitydensity}
509: a = \exp \left(- A \frac{|\Sc|}{n}\right) \rightarrow 1
510: ~~~\mbox{as}~ n \rightarrow \infty.
511: \end{equation}
512: By Theorem \ref{theo:extremecor} {\em (ii)}, the error probability
513: does not decay exponentially, but decays with polynomial order
514: $n^{-1/2}$ for any finite $A$ as $n\rightarrow \infty$. The
515: exception is the singular case where $A = \infty$, i.e.,
516: the signal is a white process. Therefore, for the detection of stationary correlated
517: fields, it is a better strategy to cover a larger area as long as
518: the signal field extends there than to localize
519: activated sensors more densely in a subregion.
520:
521:
522: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
523: \subsection{Periodic Clustering} \label{subsec:periodiccluster}
524:
525:
526: The uniform configuration for a finite signal field reveals that
527: there is a benefit at high SNR to making sensor spacing large to
528: obtain less correlated observations, but activating fewer sensors
529: results in a bigger loss than the gain from being less correlated.
530: This naturally leads to our second configuration:
531: periodic clustering shown in Fig. \ref{fig:schedulingprotocols}
532: (b), aiming at the benefits from both correlation and the number
533: of scheduled sensors. In this configuration, we activate $M$
534: sensors very close in location, and repeat this multiple
535: activation periodically over signal field so that the number of
536: scheduled sensors is preserved and the spacing between clusters
537: becomes larger than that of uniform configuration.
538:
539: For further analysis, we assume that the $M$ sensors within a
540: cluster are located at the same position. With the total number
541: of scheduled sensors $n=MN$, the observation vector $\ybf_n =
542: [y_1,y_2,\cdots, y_n]^T$ under $H_1$ is given by
543: \begin{equation}
544: \ybf_n = \tilde{\sbf}_n \otimes {\mathbf 1}_M + \wbf_n,
545: \end{equation}
546: where $\otimes$ is the Kronecker product,
547: \begin{equation}
548: \tilde{\sbf}_n =[s(0), s(\tilde{\Delta}),\cdots,
549: s(N\tilde{\Delta})]^T,
550: \end{equation}
551: and $\tilde{\Delta} = |\Sc|/N= M \Delta$. ($\Delta$ is the sensor
552: spacing for the uniform configuration for $n$ sensors in $\Sc$.) The
553: covariance matrix of $\ybf_n$ is given by
554: \begin{equation}\label{eq:covarianceclustering}
555: \Ebb \{\ybf_n \ybf_n^T\} = \left\{
556: \begin{array}{cc}
557: \Sigma_{s,N}(\tilde{a}) \otimes {\mathbf 1}_M{\mathbf 1}_M^T +\sigma^2\Ibf& \mbox{under}~H_1,\\
558: \sigma^2 \Ibf & \mbox{under}~H_0,
559: \end{array}
560: \right.
561: \end{equation}
562: where $\tilde{a}= \exp(-A \tilde{\Delta})$.
563: The signal covariance matrix has a
564: block Toeplitz structure due to the perfect correlation of
565: signal samples within a cluster. $ \Sigmabf_{s,N}(\tilde{a})$ in
566: (\ref{eq:covarianceclustering})
567: is a positive-semidefinite Toeplitz matrix where
568: the $k$th off-diagonal entries are given by $r_s(k)=\Pi_0
569: \tilde{a}^k$. For any $A > 0$, ~$0 \le \tilde{a} < 1$ and
570: $r_s(\cdot)$ is an absolutely summable sequence; the eigenvalues
571: of $\Sigmabf_{s,n}$ are bounded from above and
572: below\cite{Gray:72IT}. Using the convergence of the eigenvalues
573: of $\Sigmabf_{s,N}$ and the properties of the Kronecker product,
574: we obtain the error exponent for periodic clustering.
575:
576:
577: \begin{proposition}[Periodic Clustering] \label{prop:K_period_clustering}
578: For the Neyman-Pearson detector for the hypotheses
579: (\ref{eq:hypothesisscalar}) with level $\alpha \in (0,1)$ and
580: periodically clustered sensor configuration, the error exponent
581: of the miss probability is given by
582: \begin{equation} \label{eq:prop_errorexponent_periodcluster}
583: \tilde{K}= \frac{1}{M}K(\tilde{\Delta};M*\SNR),
584: \end{equation}
585: where $K(\tilde{\Delta};M*\SNR)$ is the error exponent for
586: uniform configuration with spacing $\tilde{\Delta}$ and $M*\SNR$ for
587: each sensor.
588: \end{proposition}
589:
590: {\em Proof:} See \cite{Sung&Tong&Poor:04SPsub}.
591:
592: \vspace{0.3em} The optimal detector for periodic clustering
593: consists of two steps. We first take an average of the
594: observations within each cluster, and then apply the optimal
595: detector for a single sample at each location to the
596: ensemble of $N$ average values. Intuitively, it is reasonable to
597: average the observations within a cluster since the signal
598: component is in the same direction and the noise is random. By
599: averaging, the magnitude of the signal component increases by $M$
600: times with the increase in the
601: noise power by the same factor; the
602: SNR within a cluster increases by the factor $M$. This is shown
603: in the relation (\ref{eq:prop_errorexponent_periodcluster}) to
604: uniform configuration. The error exponent
605: (\ref{eq:prop_errorexponent_periodcluster}) shows the advantage
606: and disadvantage of the periodic clustering over uniform configuration
607: covering the same signal field.
608: Clustering gives two benefits. First, the correlation between clusters is reduced for the same $A$
609: by making
610: the spacing larger, and the error decay per cluster increases.
611: Second, the SNR for each cluster also increases by $M$ times. However, the effective number
612: of signal samples
613: is also reduced by a factor of $M$ comparing with the uniform configuration. The performance of clustering is determined
614: by the dominating factor depending on
615: the diffusion rate of the underlying signal and the SNR of the observations.
616: \begin{figure}[htbp]
617: \centerline{ \SetLabels
618: \L(0.27*-0.1) (a) \\
619: \L(0.75*-0.1) (b) \\
620: %\L(0.83*-0.1) (c) \\
621: \endSetLabels
622: \leavevmode
623: %\ShowGrid
624: \strut\AffixLabels{
625: \scalefig{0.23}\epsfbox{figures/plot_nov4periodclusterSNR10dBA0p1.eps}
626: \scalefig{0.226}\epsfbox{figures/plot_nov4periodclusterSNR10dBA1.eps}
627: } }
628: \vspace{0.3cm}
629: \centerline{ \SetLabels
630: %\L(0.16*-0.1) (a) \\
631: %\L(0.50*-0.1) (b) \\
632: \L(0.5*-0.1) (c) \\
633: \endSetLabels
634: \leavevmode
635: %\ShowGrid
636: \strut\AffixLabels{
637: \scalefig{0.23}\epsfbox{figures/plot_nov4periodclusterSNR10dBA10.eps}
638: } }
639: \vspace{0.1cm} \caption{$\exp(-n\tilde{K})$ vs. $M$
640: ($|\Sc|=1$, SNR = 10 dB): (a) $A=0.1$ (b) $A=1$ (c) $A=10$}
641: \label{fig:tildeKnperiodclusterhighSNR}
642: \end{figure}
643: Fig. \ref{fig:tildeKnperiodclusterhighSNR} shows
644: $\exp(-n\tilde{K})$, which
645: is an approximate miss probability for large samples, for different diffusion rates at 10 dB SNR.
646: The total number
647: of sensors is fixed at $n=100$, and the cluster size is chosen as $M=[1,2,4,5,10]$.
648: For the highly correlated field ($A=0.1$), it is seen that the reduced correlation between sampled signals is dominant
649: and the periodic clustering gives better performance than uniform configuration (i.e., $M=1$).
650: (See Fig. \ref{fig:Kvsa}. The gain in the error exponent due to
651: reducing correlations is large in the high correlation region.)
652: On the other hand, for the almost independent signal field ($A=10$), clustering
653: gives worse performance than uniform configuration.
654: In this case, the correlation between the scheduled samples is already weak, and the increase
655: in the error exponent due to increased spacing is insignificant as shown in Fig. \ref{fig:Kvsa}. Hence,
656: the benefit of clustering results only from the increase
657: in SNR. By Theorem \ref{theo:etavsSNR}, however, the rate of increase in the error exponent due
658: to the increased SNR
659: is $\frac{1}{2}\log M$ at high SNR, which does not compensate for the loss in the number of effective samples
660: by the factor $1/M$.
661: For the signal field with intermediate correlation, there is a trade-off between the gain and loss of clustering,
662: and there exists an optimal clustering as shown in Fig.\ref{fig:tildeKnperiodclusterhighSNR} (b).
663: Hence, one should choose the optimal clustering depending on the diffusion rate and
664: the size of the underlying
665: signal field.
666: \begin{figure}[htbp]
667: \centerline{ \SetLabels
668: \L(0.27*-0.1) (a) \\
669: \L(0.75*-0.1) (b) \\
670: %\L(0.83*-0.1) (c) \\
671: \endSetLabels
672: \leavevmode
673: %\ShowGrid
674: \strut\AffixLabels{
675: \scalefig{0.23}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA0p1.eps}
676: \scalefig{0.23}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA1.eps}
677: } }
678: \vspace{0.3cm}
679: \centerline{ \SetLabels
680: %\L(0.16*-0.1) (a) \\
681: %\L(0.50*-0.1) (b) \\
682: \L(0.5*-0.1) (c) \\
683: \endSetLabels
684: \leavevmode
685: %\ShowGrid
686: \strut\AffixLabels{
687: \scalefig{0.23}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA10.eps}
688: } }
689: \vspace{0.1cm} \caption{$\exp(-n\tilde{K})$ vs. $M$ ($|\Xc|=1$,
690: SNR = -3 dB): (a) $A=0.1$ (b) $A=1$ (c) $A=10$}
691: \label{fig:tildeKnperiodclusterlowSNR}
692: \end{figure}
693: \begin{comment}
694: \begin{figure}[htbp]
695: \centerline{ \SetLabels
696: \L(0.16*-0.1) (a) \\
697: \L(0.50*-0.1) (b) \\
698: \L(0.83*-0.1) (c) \\
699: \endSetLabels
700: \leavevmode
701: %\ShowGrid
702: \strut\AffixLabels{
703: \scalefig{0.16}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA0p1.eps}
704: \scalefig{0.16}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA1.eps}
705: \scalefig{0.16}\epsfbox{figures/plot_nov4periodclusterSNRm3dBA10.eps}
706: } } \vspace{0.5cm} \caption{$\exp(-n\tilde{K})$ vs. $M$ ($|\Xc|=1$,
707: SNR = -3 dB): (a) $A=0.1$ (b) $A=1$ (c) $A=10$}
708: \label{fig:tildeKnperiodclusterlowSNR}
709: \end{figure}
710: \end{comment}
711: Fig. \ref{fig:tildeKnperiodclusterlowSNR} shows the approximate
712: miss probability of the periodic clustering for -3 dB SNR with
713: other parameters the same as in the high SNR case. It is seen
714: that periodic clustering outperforms than uniform configuration in all
715: considered correlation values since the SNR is the dominant factor
716: in the detector performance at low SNR.
717:
718:
719: Clustering also explains the polynomial behavior in asymptotic
720: error decay for the infinite density model considered in
721: (\ref{eq:sensorplaceinfinitydensity}). We can view an increase in
722: the number $n$ of sensors in a finite signal field as increasing
723: the cluster size $M$ with a fixed number $N$ of clusters. As $n$
724: increases, SNR per cluster increases linearly with $n$, and will
725: be in the high SNR regime eventually. At high SNR, the error
726: exponent increases at the rate of $\frac{1}{2}\log \SNR$ by
727: Theorem \ref{theo:etavsSNR}.
728: Hence, the overall miss probability is given approximately by
729: \begin{equation}
730: P_M \approx C \exp( - N K) = C_1 \exp( - \frac{1}{2}N \log n )\sim
731: C_2 n^{-1/2},
732: \end{equation}
733: for sufficiently large $n$, which coincides the results in Theorem
734: \ref{theo:extremecor} {\em (ii)}. Now it is clear that for highly
735: correlated cases the decay in error probability with an increasing
736: number of sensors is mainly due to the noise averaging effect
737: rather than the effect of the new information about the signal in
738: the observations.
739:
740: \subsection{Arbitrary Periodic Configuration}
741:
742: The previous section shows that periodic clustering outperforms
743: uniform configuration depending on the field correlation. However,
744: periodic clustering is limited since all the sensors within a
745: spatial period are scheduled on the same location. Considering
746: periodic structure we now generalize the locations of the
747: scheduled sensors within a period. First, we provide the
748: closed-form error exponent for the Neyman-Pearson detector for
749: stationary vector Gauss-Markov signals using noisy sensors. Using
750: the closed-form error exponent, we investigate the optimal
751: periodic configuration with arbitrary locations within a period.
752:
753: We again consider $n=MN$ sensors scheduled over $\Sc$ with $M$
754: sensors within a period, and denote the relative distance of $M$
755: sensors within a period as
756: \begin{equation}
757: x_1=0, ~x_2-x_1=\Delta_1, ~x_3-x_2 = \Delta_2, ~\cdots,
758: x_{M+1}-x_M =\Delta_M. \nonumber
759: \end{equation}
760: Hence, the interval of a period is $\Delta =
761: \Delta_1+\cdots+\Delta_M$. Define the signal sample and observation
762: vectors for period $i$ as
763: \begin{eqnarray}
764: \vec{s}_i &\defeq& [s_{1i},s_{2i},\cdots, s_{Mi}]^T, ~~i=1,\cdots, N, \label{eq:brevesbfi}\\
765: \vec{y}_i &\defeq& [y_{1i},y_{2i},\cdots, y_{Mi}]^T,
766: \end{eqnarray}
767: where $s_{mi} = s_{(i-1)M+m}$ and $y_{mi}=y_{(i-1)M+m}$. The
768: hypotheses (\ref{eq:hypothesisscalar}) can be rewritten in vector
769: form as
770: \begin{equation} \label{eq:hypothesisvector}
771: \begin{array}{lcl}
772: H_0 &: & \vec{y}_i = \vec{w}_i, ~~~~~~~~~~~~i=1,2,\cdots, N, \\
773: H_1 &: & \vec{y}_i = \vec{s}_i+ \vec{w}_i,\\
774: \end{array}
775: \end{equation}
776: where the measurement noise $\vec{w}_i \sim \Nc({\mathbf 0},
777: \sigma^2\Ibf_M)$ independent over $i$, and $\vec{s}_i$ satisfies
778: the vector state-space model
779: \begin{eqnarray} \label{eq:statespacevector}
780: \vec{s}_{i+1}&=&\Abf \vec{s}_i + \Bbf \vec{u}_i, \\
781: \vec{u}_i &\stackrel{i.i.d.}{\sim}& \Nc(0, ~\Qbf), ~~~\Qbf \ge
782: 0,\nn
783: \end{eqnarray}
784: where $\vec{u}_i$ are defined similarly to the quantities in
785: (\ref{eq:brevesbfi}). Specifically, the feedback and input
786: matrices, $\Abf$ and $\Bbf$, are given from the scalar
787: state-space model as {\scriptsize
788: \begin{equation} \label{eq:periodpatternABQ}
789: \Abf=
790: \left[ \begin{array}{cccc}
791: 0 & 0 & 0 & e^{-A\Delta_M}\\
792: 0 & 0 & 0 & e^{-A(\Delta_M+\Delta_1)}\\
793: 0 & 0 & 0 &\vdots\\
794: 0 & 0 & 0 &e^{-A(\Delta_M+\Delta_1+\cdots+\Delta_{M-1})}
795: \end{array}\right],
796: \end{equation}
797: and
798: \begin{equation}
799: \Bbf= \left[ \begin{array}{cccc}
800: 1 & 0 & 0 & 0 \\
801: e^{-A\Delta_1} & 1 & 0 & 0\\
802: \vdots & \cdots & \ddots & 0 \\
803: e^{-A(\Delta_1+\cdots+\Delta_{M-1})} & & e^{-A\Delta_{M-1}} & 1
804: \end{array}\right],
805: \end{equation}
806: } and {\scriptsize
807: \begin{equation}
808: \Qbf=\Pi_0
809: \mbox{diag}((1-e^{-2A\Delta_M}),(1-e^{-2A\Delta_1}),\cdots,(1-e^{-2A\Delta_{M-1}})).
810: \nonumber
811: \end{equation}
812: } Notice that $\Abf$, $\Bbf$, and $\Qbf$ are not varying with $i$
813: due to the periodicity. Only the last column of $\Abf$ is non-zero
814: due to the Markov property of the scalar process $\{s_i\}$, and
815: the corresponding non-zero eigenvalue of $\Abf$ is simply given by
816: $\lambda = e^{-A\Delta}$ so that
817: $|\lambda| < 1$ for arbitrary sensor locations within a period for any diffusion rate $A > 0$. Notice that the
818: eigenvalue is the same as the correlation coefficient $a$ with sampling distance $\Delta$, the period
819: of the interval.
820: The initial condition for the vector model is given by
821: \begin{equation}
822: \vec{s}_1 \sim \Nc(0, \Cbf_0),
823: \end{equation}
824: where $\Cbf_0$ is given by{\scriptsize
825: \begin{equation}
826: \Pi_0 \left[
827: \begin{array}{ccccc}
828: 1 & e^{-A\Delta_{1,2}} & e^{-A\Delta_{1,3}} & \cdots & e^{-A\Delta_{1,M}} \\
829: e^{-A\Delta_{2,1}} & 1 & e^{-A\Delta_{2,3}} & & e^{-A\Delta_{2,M}} \\
830: e^{-A\Delta_{3,1}} & e^{-A\Delta_{3,3}} & 1 & e^{-A\Delta_{3,4}} & \vdots \\
831: \vdots & & & \ddots & e^{-A\Delta_{1,M-1}} \\
832: e^{-A\Delta_{M,1}} &\cdots & & e^{-A\Delta_{M-1,1}}&1 \\
833: \end{array}
834: \right], \nonumber
835: \end{equation}
836: } and $\Delta_{i,j} \defeq |x_j - x_i|$.
837: The initial covariance matrix $\Cbf_0$ is derived from the scalar initial condition $s_1 \sim \Nc(0,\Pi_0)$,
838: and it can be shown that $\Cbf_0$ satisfies the following Lyapunov equation
839: \begin{equation}
840: \Cbf_0 = \Abf \Cbf_0 \Abf^T + \Bbf\Qbf\Bbf^T.
841: \end{equation}
842: Thus, the vector signal sequence $\{\vec{s}_i\}$ is a stationary
843: process
844: although the scalar process is not in general for the arbitrary periodic configuration.
845:
846: For the vector case, the innovations
847: approach \cite{Sung&Tong&Poor:04ITsub} to obtain the error exponent
848: is very useful, and
849: provides a closed-form formula for the error exponent of the
850: Neyman-Pearson detection of stationary vector processes in noisy
851: observations.
852:
853: \vspace{0.5em}
854: \begin{proposition}[Arbitrary Periodic Configuration] \label{theo:errorexponentNPvector}
855: For the Neyman-Pearson detector for the hypotheses
856: (\ref{eq:hypothesisvector},\ref{eq:statespacevector}) with level
857: $\alpha \in (0,1)$ (i.e. $ P_F \le \alpha$), the best error
858: exponent of the miss probability (per a vector observation) is
859: given by
860: \begin{equation} \label{eq:errorexponentNPtheovector}
861: K_v = -\frac{1}{2}\log \frac{\sigma^{2m}}{det(\Rbf_{e})}+
862: \frac{1}{2}\mbox{tr}\left(\Rbf_{e}^{-1}
863: \tilde{\Rbf}_{e}\right)-\frac{m}{2},
864: \end{equation}
865: independently of the value of $\alpha$. The steady-state
866: covariance matrices $\Rbf_{e}$
867: and $\tilde{\Rbf}_{e}$ of the innovation process calculated under $H_1$ and $H_0$, respectively,
868: are given by
869: \begin{equation}
870: \Rbf_{e} =\sigma^2\Ibf_m + \Pbf,
871: \end{equation}
872: where $\Pbf$ is the unique stabilizing
873: solution of the discrete-time Riccati equation
874: \begin{equation}
875: \label{eq:RiccatiVector} \Pbf = \Abf \Pbf \Abf^T + \Bbf\Qbf\Bbf^T -
876: \Abf \Pbf \Rbf_e^{-1} \Pbf \Abf^T,
877: \end{equation}
878: and
879: \begin{equation}
880: \tilde{\Rbf}_{e} =
881: \sigma^2(\Ibf_m+ \tilde{\Pbf}),
882: \end{equation}
883: where
884: $\tilde{\Pbf}$ is the unique positive-semidefinite solution of the
885: following Lyapunov equation
886: \begin{equation} \label{eq:theo1Lyapunov}
887: \tilde{\Pbf}= (\Abf - \Kbf_p) \tilde{\Pbf}(\Abf-\Kbf_p)^T + \Kbf_p
888: \Kbf_p^T,
889: \end{equation}
890: and $\Kbf_p =
891: \Abf\Pbf \Rbf_{e}^{-1}$.
892: \end{proposition}
893:
894: \vspace{0.2em}
895: Proposition \ref{theo:errorexponentNPvector} is
896: proved by extending the results in \cite{Sung&Tong&Poor:04ITsub}
897: with modification from scalar to vector observations.
898:
899: Using (\ref{eq:errorexponentNPtheovector}), we now investigate the
900: large sample detection performance for arbitrary periodic
901: configuration. First, we consider the case of $M=2$ in which we have
902: freedom to schedule one intermediate sensor at an arbitrary
903: location within an interval $\Delta$. Periodic clustering and
904: uniform configuration are special cases of this configuration with
905: $\Delta_1=0$ and $\Delta_1=\Delta/2$, respectively.
906: \begin{figure}[htbp]
907: \centerline{ \SetLabels
908: \L(0.25*-0.1) (a) \\
909: \L(0.76*-0.1) (b) \\
910: \endSetLabels
911: \leavevmode
912: %\ShowGrid
913: \strut\AffixLabels{
914: \scalefig{0.23}\epsfbox{figures/plot_nov6periodpatterKvsD1A1.eps}
915: \scalefig{0.23}\epsfbox{figures/plot_nov6periodpatterKvsD1A8.eps}
916: } } \vspace{0.3cm} \centerline{ \SetLabels
917: \L(0.25*-0.1) (c) \\
918: \L(0.76*-0.1) (d) \\
919: \endSetLabels
920: \leavevmode
921: %\ShowGrid
922: \strut\AffixLabels{
923: \scalefig{0.23}\epsfbox{figures/plot_nov6periodpatterKvsD1A15.eps}
924: \scalefig{0.23}\epsfbox{figures/plot_nov6periodpatterKvsD1A100.eps}
925: } } \vspace{0.5cm} \caption{$K_v$ vs. $\Delta_1$ ($M=2$,
926: $\Delta=0.02$, SNR = 10 dB): (a) $A=1$ (b) $A=8$ (c) $A=15$ (d)
927: $A=100$} \label{fig:periodpatternM2highSNR}
928: \end{figure}
929: Fig.
930: \ref{fig:periodpatternM2highSNR} shows the error exponent for
931: different diffusion rates at high SNR. We observe an
932: interesting behavior w.r.t. the diffusion rate. For the highly
933: correlated field ($A=1$), periodic clustering ($\Delta_1=0$) gives the
934: best performance while uniform configuration provides the worst.
935: However, as the field correlation becomes weak ($A=8$), we observe
936: a second lobe grow at the uniform configuration point ($\Delta_1 =
937: \Delta/2$). The value of the second lobe becomes larger than
938: that corresponding clustering as the correlation becomes weaker
939: ($A=15$), and eventually the error exponent decreases
940: monotonically as the configuration deviates from uniform
941: configuration to periodic clustering. This behavior of the error exponent
942: clearly shows that the optimal configuration depends on the field
943: correlation. Consistent with the results in the previous sections,
944: one should reduce the correlation between observations for highly
945: correlated fields while the uniform configuration is best for almost
946: independent signal fields. Interestingly, the optimal configuration
947: for $M=2$ at high SNR is either the clustering or the uniform
948: configuration depending on the field correlation; no configuration
949: in-between is optimal! Fig. \ref{fig:periodpatternM2lowSNR} shows
950: the error exponent for $M=2$ at low SNR. It is seen that
951: clustering is always the best strategy for all values of field correlation
952: considered since the increase in the effective SNR due to noise
953: averaging is the dominant factor in detection performance at low
954: SNR. It is also seen that the location of the intermediate sensor
955: is not important for the highly uncorrelated field ($A=1000$)
956: unless it is very close to the first sensor within a period. This
957: is intuitively obvious since the intermediate sensor provides an
958: almost independent observation (for which the location does not
959: matter) as it separates from the first and the noise averaging is
960: not available between the independent samples.
961: \begin{figure}[htbp]
962: \centerline{ \SetLabels
963: \L(0.27*-0.1) (a) \\
964: \L(0.75*-0.1) (b) \\
965: %\L(0.83*-0.1) (c) \\
966: \endSetLabels
967: \leavevmode
968: %\ShowGrid
969: \strut\AffixLabels{
970: \scalefig{0.24}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA1.eps}
971: \scalefig{0.235}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA100.eps}
972: } }
973: \vspace{0.3cm}
974: \centerline{ \SetLabels
975: %\L(0.16*-0.1) (a) \\
976: %\L(0.50*-0.1) (b) \\
977: \L(0.5*-0.1) (c) \\
978: \endSetLabels
979: \leavevmode
980: %\ShowGrid
981: \strut\AffixLabels{
982: \scalefig{0.24}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA1000.eps}
983: } }
984: \vspace{0.1cm} \caption{$K_v$ vs. $\Delta_1$ ($M=2$,
985: $\Delta=0.02$, SNR = -3 dB): (a) $A=1$ (b) $A=100$ (c) $A=1000$}
986: \label{fig:periodpatternM2lowSNR}
987: \end{figure}
988: \begin{comment}
989: \begin{figure}[htbp]
990: \centerline{ \SetLabels
991: \L(0.16*-0.1) (a) \\
992: \L(0.50*-0.1) (b) \\
993: \L(0.83*-0.1) (c) \\
994: \endSetLabels
995: \leavevmode
996: %\ShowGrid
997: \strut\AffixLabels{
998: \scalefig{0.16}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA1.eps}
999: \scalefig{0.16}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA100.eps}
1000: \scalefig{0.16}\epsfbox{figures/plot_nov6periodpatterKvsD1SNRm3dBA1000.eps}
1001: } } \vspace{0.5cm} \caption{$K_v$ vs. $\Delta_1$ ($M=2$,
1002: $\Delta=0.02$, SNR = -3 dB): (a) $A=1$ (b) $A=100$ (c) $A=1000$}
1003: \label{fig:periodpatternM2lowSNR}
1004: \end{figure}
1005: \end{comment}
1006:
1007: For the case of $M=3$, due to the periodicity, the location of one
1008: sensor in a period is fixed and we can choose the locations $(x_2,
1009: x_3)$ of the two other sensors arbitrarily such that $ 0 \le x_2
1010: \le \Delta$ and $0 \le x_3 \le \Delta$. Fig.
1011: \ref{fig:periodpatternM3highSNR} shows the error exponent as a
1012: function of $(x_2,x_3)$ for $M=3$ at 10 dB SNR. Similar behavior
1013: is seen as in the case of $M=2$.
1014: \begin{figure}[htbp]
1015: \centerline{ \SetLabels
1016: \L(0.25*-0.1) (a) \\
1017: \L(0.76*-0.1) (b) \\
1018: \endSetLabels
1019: \leavevmode
1020: %\ShowGrid
1021: \strut\AffixLabels{
1022: \scalefig{0.2}\epsfbox{figures/plot_nov4periodpatternKvSNR10dBM3A1.eps}
1023: \hspace{0.5cm}
1024: \scalefig{0.2}\epsfbox{figures/plot_nov4periodpatternKvSNR10dBM3A5.eps}
1025: } } \vspace{0.5cm} \centerline{ \SetLabels
1026: \L(0.25*-0.1) (c) \\
1027: \L(0.76*-0.1) (d) \\
1028: \endSetLabels
1029: \leavevmode
1030: %\ShowGrid
1031: \strut\AffixLabels{
1032: \scalefig{0.2}\epsfbox{figures/plot_nov4periodpatternKvSNR10dBM3A6.eps}
1033: \hspace{0.5cm}
1034: \scalefig{0.2}\epsfbox{figures/plot_nov4periodpatternKvSNR10dBM3A9.eps}
1035: } } \vspace{0.5cm} \caption{$K_v$ vs. $(x_2,x_3)$ ($M=3$,
1036: $\Delta=0.03$, SNR = 10 dB): (a) $A=1$ (b) $A=5$ (c) $A=6$ (d)
1037: $A=9$ (red:high value, blue:low value)}
1038: \label{fig:periodpatternM3highSNR}
1039: \end{figure}
1040: For the highly correlated signal ($A=1$), we see the maximal value
1041: of the error exponent at $(0,0)$,$(0,\Delta)$,$(\Delta,0)$, and
1042: $(\Delta,\Delta)$, which all correspond to periodic clustering.
1043: Hence, periodic clustering is the best among all configurations.
1044: In this case, it is seen that uniform configuration is the worst
1045: configuration. As the field correlation becomes weak, however,
1046: the best configuration moves to uniform configuration eventually, as
1047: seen in Fig \ref{fig:periodpatternM3highSNR} (d). It is seen that
1048: placing two sensors clustered and one in the middle of the spatial
1049: period is optimal for transitory values of field correlation as
1050: shown in Fig. \ref{fig:periodpatternM3highSNR} (b) and (c). Figure \ref{fig:optimalconfigM3} summarizes the optimal configuration for different field correlation.
1051: \begin{figure}[hbtp]
1052: \centerline{
1053: \begin{psfrags}
1054: \psfrag{D1}[c]{$\Delta$}
1055: \psfrag{D2}[c]{$\Delta/2$}
1056: \psfrag{D3}[c]{$\Delta/3$}
1057: \psfrag{PC}[l]{{\small Periodic clustering}}
1058: \psfrag{Uniform}[l]{{\small Uniform configuration}}
1059: \psfrag{sensor}[l]{{\small Activated sensor}}
1060: \psfrag{FC}[l]{{\small Field correlation}}
1061: \psfrag{Strong}[l]{{\small Strong}}
1062: \psfrag{Weak}[l]{{\small Weak}}
1063: \scalefig{0.38}\epsfbox{figures/optimalproto.eps}
1064: \end{psfrags}
1065: } \caption{Optimal configuration for $M=3$ (SNR = 10 dB)}
1066: \label{fig:optimalconfigM3}
1067: \end{figure}
1068: Other
1069: results confirm that periodic clustering gives the best
1070: configuration for most values of field correlations at low SNR.
1071:
1072: \subsection{Sensor Placement}
1073:
1074: So far we have assumed that sensors have already been deployed and
1075: considered the activation for sensing and transmissions from the selected
1076: sensors. A related problem is sensor placement. When $n$ sensors
1077: are planned to be deployed over a signal field for the detection
1078: application, how should we place the $n$ sensors in the field? The
1079: results in the previous sections provide the answer for this
1080: problem as well.
1081:
1082: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1083: \section{Conclusion}
1084: \label{sec:conclusion}
1085: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1086:
1087: We have considered energy-efficient sensor activation for large sensor
1088: networks deployed to detect correlated random fields.
1089: Using our results on large-sample error behavior in this application,
1090: we have analyzed and compared the detection capabilities of
1091: different sensor configuration strategies. The optimal configuration
1092: is a function of the field correlation and the SNR of sensor
1093: observations. For uniform configuration, the scheduled sensors
1094: should be maximally separated to cover the entire signal field for
1095: SNR $>1$. For SNR $<1$, on the other hand, there exists an optimal
1096: spacing between the scheduled sensors. We have also derived the
1097: error exponents of periodic clustering and arbitrary periodic
1098: configuration. Periodic clustering may outperform uniform configuration
1099: depending on the field correlation and SNR. Furthermore, there
1100: exists an optimal cluster size for intermediate values of
1101: correlation. The closed-form error exponent obtained for the
1102: vector state-space model explains the transitory error behavior
1103: from periodic clustering to uniform configuration.
1104:
1105:
1106:
1107:
1108: %%%%%%%%%% References %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1109:
1110: {\scriptsize
1111: \bibliographystyle{plain}
1112: \begin{thebibliography}{9}
1113:
1114:
1115: \bibitem{Poor:book}
1116: H.~V.~Poor, {\em An Introduction to Signal Detection and Estimation}, 2nd Edition, Springer, New York, 1994.
1117:
1118: \bibitem{Helstrom:book}
1119: C.~W.~Helstrom, {\em Elements of Signal Detection and Estimation},
1120: Prentice-Hall, Englewood Cliffs, NJ, 1994.
1121:
1122:
1123: \bibitem{Cox&Miller:book}
1124: D. Cox and H. Miller, {\em The Theory of Stochastic Processes}, John Wiley \& Sons Inc., New York, 1965.
1125:
1126: \bibitem{Micheli&Jordan:02MTNS}
1127: M.~Micheli and M. I. Jordan, ``{Random sampling of a continuous-time stochastic
1128: dynamical system,}'' {\em Proc. of the 15th International Symposium on the Mathematical Theory of Networks and Systems (MTNS 2002),} (University of Notre Dame, South Bend, Indiana), August 2002.
1129:
1130:
1131: \bibitem{Kailath&Poor:98IT}
1132: T.~Kailath and H.~V.~Poor, ``{Detection of stochastic processes,}''
1133: {\em IEEE Trans. on Information Theory}, vol. 44, no. 6, pp. 2230-2259, Oct. 1998.
1134:
1135:
1136:
1137: \bibitem{Schweppe:65IT} F. C. Schweppe, `` Evaluation of likelihood functions for Gaussian signals,'' {\em IEEE Trans. on Information Theory}, vol.IT-1, pp.61-70, 1965.
1138:
1139:
1140: \bibitem{Gray:72IT}
1141: R. Gray, ``{On the asymptotic eigenvalue distribution of Toeplitz
1142: matrices,}'' {\em IEEE Trans. on Information Theory,} vol. 18,
1143: no. 6, pp. 725-730, Nov 1972.
1144:
1145:
1146:
1147:
1148: \bibitem{Benitz&Bucklew:90IT}
1149: G. R. Benitz and J. A. Bucklew, ``{Large deviation rate calculations for
1150: nonlinear detectors in Gaussian noise,}''
1151: {\em IEEE Trans. on Information Theory}, vol. 36, no. 2, pp.358-371, March 1990.
1152:
1153: \bibitem{Bahr:90IT}
1154: R. K. Bahr, ``{Asymptotic analysis of error probabilities for the nonzero-mean Gaussian hypothesis testing problem,}''
1155: {\em IEEE Trans. on Information Theory}, vol. 36, no. 3, pp.597-607, March 1990.
1156:
1157:
1158:
1159: \bibitem{Bahr&Bucklew:90SP}
1160: R. K. Bahr and J. A. Bucklew, ``{Optimal sampling schemes for the Gaussian
1161: hypothesis testing problem,}''
1162: {\em IEEE Trans. on Acoustics, Speech, and Signal Processing}, vol. 38, no. 10, pp.1677-1686, Oct. 1990.
1163:
1164:
1165: \bibitem{Barone&Gigli&Piccioni:95IT}
1166: P. Barone, A. Gigli, and M. Piccioni, ``Optimal importance sampling for some quadratic forms of ARMA processes,''
1167: {\em IEEE Trans. on Information Theory,} vol. 41, no. 6, pp. 1834 - 1844, 1995.
1168:
1169:
1170: \bibitem{Bryc&Smolenski:93SPL}
1171: W. Bryc and W. Smolenski, ``On the large deviation principle for a quadratic functional of the autoregressive process,'' {\em Statistics and Probability Letters,} vol. 17, pp. 281-285, 1993.
1172:
1173:
1174: \bibitem{Bercu&Gamboa&Rouault:97SPA}
1175: B. Bercu, F. Gamboa, and A. Rouault, ``Large deviations for quadratic forms of stationary Gaussian processes,'' {\em Stochastic Processes and their Applications,} vol. 71, pp. 75-90, 1997.
1176:
1177:
1178:
1179: \bibitem{Luschgy:94SJS} H. Luschgy, ``Asymptotic behavior of Neyman-Pearson tests for autoregressive processes,'' {\em Scand. J. Statist.}, vol. 21, pp. 461-473, 1994.
1180:
1181: \bibitem{Vajda:book} I. Vajda, {\em Theory of Statistical Inference and Information,} Kluwer Academic Publishers, Dordrecht, 1989.
1182:
1183:
1184: \bibitem{Chamberland&Veeravalli:04ITWS}
1185: J.-F. Chamberland and V. V. Veeravalli, ``Design of sensor
1186: networks for detection applications via large-deviation theory,"
1187: {\em Proc. 2004 IEEE Information Theory Workshop,} San Antonio,
1188: TX, Oct. 2004.
1189:
1190: \bibitem{Sung&Tong&Poor:04ITsub} Y. Sung, L. Tong, and H. V. Poor, ``Neyman-Pearson detection of
1191: Gauss-Markov signals in noise: Closed-form error exponent and
1192: properties,'' submitted to {\em IEEE Trans. on Information Theory}, Nov. 2004.
1193:
1194: \bibitem{Sung&Tong&Poor:04SPsub} Y. Sung, L. Tong, and H. V. Poor, ``Sensor scheduling and activation for field
1195: detection in large sensor arrays,'' preprint, Nov. 2004.
1196:
1197:
1198:
1199:
1200:
1201: \end{thebibliography}
1202: }
1203:
1204:
1205:
1206:
1207:
1208:
1209:
1210: \end{document}
1211:
1212:
1213:
1214:
1215:
1216:
1217: