Skip to main content
Top
Published in: EURASIP Journal on Wireless Communications and Networking 1/2010

Open Access 01-12-2010 | Research Article

Random Field Estimation with Delay-Constrained and Delay-Tolerant Wireless Sensor Networks

Authors: Javier Matamoros, Carles Antón-Haro

Published in: EURASIP Journal on Wireless Communications and Networking | Issue 1/2010

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

In this paper, we study the problem of random field estimation with wireless sensor networks. We consider two encoding strategies, namely, Compress-and-Estimate (C&E) and Quantize-and-Estimate (Q&E), which operate with and without side information at the decoder, respectively. We focus our attention on two scenarios of interest: delay-constrained networks, in which the observations collected in a particular timeslot must be immediately encoded and conveyed to the Fusion Center (FC); delay-tolerant (DT) networks, where the time horizon is enlarged to a number of consecutive timeslots. For both scenarios and encoding strategies, we extensively analyze the distortion in the reconstructed random field. In DT scenarios, we find closed-form expressions of the optimal number of samples to be encoded in each timeslot (Q&E and C&E cases). Besides, we identify buffer stability conditions and a number of interesting distortion versus buffer occupancy tradeoffs. Latency issues in the reconstruction of the random field are addressed, as well. Computer simulation and numerical results are given in terms of distortion versus number of sensor nodes or SNR, latency versus network size, or buffer occupancy.

1. Introduction

In recent years, research Wireless Sensor Networks (WSNs) has attracted considerable attention. This is in part motivated by the large number of applications in which WSNs are called to play a pivotal role, such as parameter estimation (i.e., moisture, temperature), event detection (leakage of pollutants, earthquakes, fires), or localization and tracking (e.g., border control, inventory tracking), to name a few [1].
Typically, a WSN consists of one Fusion Center (FC) and a potentially large number of sensor nodes capable of collecting and transmitting data to the FC over wireless links. In many cases, the underlying phenomenon being monitored can be modeled as a spatial random field. In these circumstances, the set of sensor observations are correlated, with such correlation being typically a function of their spatial locations (see, e.g., [2]). By effectively handling correlation in the data encoding process, substantial energy savings can be achieved.
In a source coding context, the work in [3] constitutes a generalization to sensor trees of Wyner-Ziv's pioneering studies [4]. The authors propose two coding strategies, namely Quantize-and-Estimate (Q&E) and Compress-and-Estimate (C&E), and analyze their performance for various networks topologies. The Q&E encoding scheme is a particularization of Wyner-Ziv's to scenarios with no side information at the decoder. Consequently, each sensor observation is encoded (and decoded) independently. Conversely, C&E turns out to be a successive Wyner-Ziv-based coding scheme and, for this reason, it is capable of exploiting spatial correlation.
In a context of random field estimation with WSNs, the pioneering work of [5] introduced the so-called "bit-conservation principle". The authors prove that, for spatially bandlimited processes, the bit budget per Nyquist-period can be arbitrarily reallocated along the quantization precision and/or the space (by adding more sensor nodes) axes, while retaining the same decay profile of the reconstruction error. In [6] and, again, for bandlimited processes with arbitrary statistical distributions, the authors propose a mathematical framework to study the impact of the random sampling effect (arising from the adoption of contention-based multiple-access schemes) on the resulting estimation accuracy. For Gaussian observations, [7] presents a feedback-assisted Bayesian framework for adaptive quantization at the sensor nodes.
From a different perspective but still in a context of random field estimation, [2] proposes a novel MAC protocol which minimizes the attempts to transmit correlated data. By doing so, not only energy but also bandwidth is preserved. Besides, in [8], the authors investigate the impact of random sampling, as opposed to deterministic sampling (i.e., equally-spaced sensors) which is difficult to achieve in practice, in the reconstruction of the field. The main conclusion is that, whereas deterministic sampling pays off in the high-SNR regime, both schemes exhibit comparable performances in the low-SNR regime.
Contribution
In this paper, we address the problem of (nonnecessarily bandlimited) random field estimation with wireless sensor networks. To that aim, we adopt the Q&E and C&E encoding schemes of [3] and analyze their performance in two scenarios of interest: delay-constrained (DC) and delay-tolerant (DT) sensor networks. In DC scenarios, the observations collected in a particular timeslot must be immediately encoded and conveyed to the FC. In DT networks, on the contrary, the time horizon is enlarged to https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq1_HTML.gif consecutive timeslots. Clearly, this entails the use of local buffers but, in exchange, the distortion in the reconstructed random field is lower. To capitalize on this, we derive closed-form expressions of the distortion attainable in DT scenarios (unlike in [2, 6, 8], we explicitly take into account quantization effects). From this, we determine the optimal number of samples to be encoded in each of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq2_HTML.gif timeslots as a function of the channel conditions of that particular timeslot. This constitutes the first original contribution of the paper. Along with that, we identify under which circumstances buffers are stable (i.e., buffer occupancy does not grow without bound) and, besides, we study a number of distortion versus buffer occupancy tradeoffs. To the best of our knowledge, such analysis has not been conducted before in a context of random field estimation. Complementarily, we analyze the latency in the reconstruction of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq3_HTML.gif consecutive realizations (i.e., those collected in one timeslot) of the random field, this being an original contribution, as well.
The paper is organized as follows. First, in Section 2, we present the signal and communication models, and provide a general framework for distortion analysis. Next, in Section 3, we focus on delay-constrained scenarios and particularize the aforementioned distortion analysis. In Sections 4 and 5 instead, we address delay-tolerant scenarios and analyze the behavior of the Q&E and C&E encoding schemes, respectively. Next, Section 6 investigates latency issues associated with DT networks. In Section 7, we present some computer simulations and numerical results and, finally, we close the paper by summarizing the main findings in Section 8.

2. Signal Model

Let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq4_HTML.gif be a one-dimensional random field defined in the range https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq5_HTML.gif , with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq6_HTML.gif denoting the spatial variable. As in [2, 8, 9], we adopt a stationary homogeneous Gaussian Markov Ornstein-Uhlenbeck (GMOU) model [10] to characterize the dynamics and spatial correlation of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq7_HTML.gif . GMOU random fields obey the following linear stochastic differential equation
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ1_HTML.gif
(1)
where, by definition, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq8_HTML.gif with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq9_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq10_HTML.gif denotes Brownian Motion with unit variance parameter, and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq11_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq12_HTML.gif are constants reflecting the (spatial) variability of the field and its noisy behaviour, respectively. According to this model, the autocorrelation function is given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq13_HTML.gif and, hence, the process is not (spatially) bandlimited.
The random field is uniformly sampled by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq14_HTML.gif sensor nodes, with intersensor distance given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq15_HTML.gif (see Figure 1). The spatial samples can thus be readily expressed as follows [11]:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ2_HTML.gif
(2)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq16_HTML.gif .

2.1. Communication Model

As shown in Figure 2, each time slot is composed of two distinctive phases namely, the sensing phase and the transmission phase. In the former, each sensor collects and stores in a local buffer a large block of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq17_HTML.gif independent and consecutive observations https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq18_HTML.gif . Next, in the transmission phase, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq19_HTML.gif is block-encoded into a length- https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq20_HTML.gif codeword https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq21_HTML.gif in codebook https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq22_HTML.gif at a rate of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq23_HTML.gif bits per sample. The encoding (quantization) process is modeled through the auxiliary random variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq24_HTML.gif with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq25_HTML.gif standing for memoryless Gaussian noise with variance https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq26_HTML.gif and statistically independent of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq27_HTML.gif (for the ease of notation, we drop the sample index.). The corresponding codeword index https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq28_HTML.gif is then conveyed to the FC, in a total of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq29_HTML.gif channel uses, over one of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq30_HTML.gif orthogonal channels (for other encoding schemes, such as Compress-and-Estimate in Section 3.2, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq31_HTML.gif denotes the index of the bin to which the codeword belongs to. For further details, see [3]). The codeword can only be reliably decoded at the FC if the encoding rate https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq32_HTML.gif satisfies
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ3_HTML.gif
(3)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq33_HTML.gif stands for the average signal-to-noise ratio experienced in the sensor-to-FC channels, and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq34_HTML.gif denote the corresponding channel squared gains. In the sequel, such gains will be modeled as independent and exponentially-distributed unit-mean random variables (i.e., Rayleigh-fading channels) and independent over time slots (block fading assumption).
From the set of decoded codewords, the FC reconstructs the random field https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq35_HTML.gif for all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq36_HTML.gif . As a result of the spatial sampling process and the channel bandwidth constraint, the reconstructed field https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq37_HTML.gif is subject to some distortion which, throughout this paper, will be characterized by the following metric
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ4_HTML.gif
(4)

2.2. Distortion Analysis: A General Framework

For the distortion metric given by (4), the optimal estimator turns out to be the posterior mean given all the codewords https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq38_HTML.gif ; that is, the MMSE estimator [12, Chapter 10]
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ5_HTML.gif
(5)
For mathematical tractability, however, only the two closest decoded codewords, namely https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq39_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq40_HTML.gif , will be used to reconstruct https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq41_HTML.gif for all the corresponding intermediate spatial points (in noiseless scenarios, that is, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq42_HTML.gif for all https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq43_HTML.gif , this approach turns out to be optimal due to the Markovian property of GMOU processes. For the general case, yet suboptimal, it capitalizes on the codewords which retain more information on the random field at the spatial point https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq44_HTML.gif ) (see Figure 1), that is
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ6_HTML.gif
(6)
For the ease of notation and without loss of generality, in the sequel, we assume https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq45_HTML.gif and, hence, the interval between observations reads https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq46_HTML.gif . From [12, Chapter 10], the distortion associated to the estimator (6) is given by
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ7_HTML.gif
(7)
For our signal model and after some algebra, the various terms in the expression above can be computed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ8_HTML.gif
(8)
It is worth noting that the variance of the quantization noise https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq47_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq48_HTML.gif are determined by the encoding strategy in use at the sensor nodes.

3. Delay-Constrained WSNs

In delay-constrained (DC) networks, the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq49_HTML.gif samples collected in the sensing phase of a given timeslot must be necessarily encoded and transmitted to the FC in the corresponding transmission phase. Bearing this in mind, we particularize the analysis of Section 2.2 and compute the average distortion for the cases of Delay-Constrained Quantize-and-Estimate (QEDC) and Compress-and-Estimate (CEDC) encoding strategies.

3.1. Quantize-and-Estimate: Average Distortion

Here, each sensor encodes its observation regardless of any side information that could be made available to the FC. From [13], the following inequality holds for the rate at the output of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq50_HTML.gif th encoder (quantizer)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ9_HTML.gif
(9)
with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq51_HTML.gif standing for the mutual information. As discussed before, the encoding (quantization) process is modeled through the auxiliary variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq52_HTML.gif with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq53_HTML.gif and statistically independent of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq54_HTML.gif (see, e.g., [3, 14] for further details). The minimum rate per sample can be expressed as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ10_HTML.gif
(10)
From (3), (9), and (10) we have that, necessarily,
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ11_HTML.gif
(11)
By taking equality in (11), the variance of the quantization noise yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ12_HTML.gif
(12)
with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq55_HTML.gif standing for the sample-to-channel uses ratio. By replacing (12) into (7), the distortion in an arbitrary spatial point https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq56_HTML.gif in the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq57_HTML.gif th segment reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ13_HTML.gif
(13)
with
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ14_HTML.gif
(14)
The average distortion (over the spatial variable https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq58_HTML.gif ) in the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq59_HTML.gif th network segment can be computed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ15_HTML.gif
(15)
and, from this, the average distortion (over channel realizations) follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ16_HTML.gif
(16)

3.2. Compress-and-Estimate: Average Distortion

In Compress-and-Estimate encoding, the FC incorporates some side information into the decoding process. This extent can be exploited by the sensors in order to encode their observations more efficiently. For simplicity, we assume that only the codeword sent by the adjacent sensor, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq60_HTML.gif will be used as side information for decoding codeword https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq61_HTML.gif (alternatively, we could use all the sensor observations but due to the spatial Markov property of the random field model, this is not expected to substantially decrease the encoding rate). Accordingly, the minimum rate per sample can be expressed as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ17_HTML.gif
(17)
where the second equality is due to the fact that, again, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq62_HTML.gif form a Markov chain. Clearly, the codeword can be reliably transmitted if and only if
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ18_HTML.gif
(18)
By taking equality in (18), the minimum variance of the quantization noise https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq63_HTML.gif follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ19_HTML.gif
(19)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq64_HTML.gif can be easily computed as:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ20_HTML.gif
(20)
From (7), the distortion at an arbitrary spatial point https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq65_HTML.gif reads:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ21_HTML.gif
(21)
with
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ22_HTML.gif
(22)
The average distortion for each network segment can be computed as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ23_HTML.gif
(23)
and, finally, the average distortion (over the channel realizations and network segments) yields:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ24_HTML.gif
(24)

4. Delay-Tolerant WSNs with Quantize-and-Estimate Encoding

Here, we impose a long-term delay constraint: the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq66_HTML.gif samples collected in https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq67_HTML.gif consecutive timeslots must be conveyed to the FC in such https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq68_HTML.gif timeslots. In other words, sensors have now the flexibility to encode and transmit a variable number of samples in each time slot (according to channel conditions) and, by doing so, attain a lower distortion.
Let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq69_HTML.gif be the number of samples encoded in https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq70_HTML.gif channel uses by sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq71_HTML.gif in time-slot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq72_HTML.gif . As in the previous section, we have that
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ25_HTML.gif
(25)
By replacing https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq73_HTML.gif from (25) into (7), the distortion per timeslot yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ26_HTML.gif
(26)
In order to minimize the average distortion over the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq74_HTML.gif timeslots at an arbitrary spatial point https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq75_HTML.gif , we need to solve the following optimization problem, implicitly, we are assuming that sensor ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq76_HTML.gif )th encodes at a constant rate over timeslots. This extent will be verified later on in this section:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ27_HTML.gif
(27)
where the constraint in (27) is introduced to ensure the stability of the system. Unfortunately, a closed-form solution for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq77_HTML.gif cannot be obtained for this problem. Instead, we attempt to solve an approximate problem in which we assume that only codeword https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq78_HTML.gif will be used by the FC to reconstruct the random field https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq79_HTML.gif in https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq80_HTML.gif . Yet, suboptimal (the FC will actually use both codewords, namely https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq81_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq82_HTML.gif ), this solution outperforms those obtained in delay-constrained scenarios (see computer simulations section). Bearing all this in mind, the new cost function which follows from (26) can be readily expressed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ28_HTML.gif
(28)
Clearly, only the second term in the summation of the cost function https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq83_HTML.gif is relevant to the optimization problem, which can be rewritten as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ29_HTML.gif
(29)
It is straightforward to show that this problem is convex. Hence, one can construct the lagrangian as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ30_HTML.gif
(30)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq84_HTML.gif is the Lagrange multiplier. By setting the first derivative of (30) w.r.t. https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq85_HTML.gif to zero we obtain
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ31_HTML.gif
(31)
with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq86_HTML.gif denoting the negative real branch of the Lambert function [15]. As for the computation of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq87_HTML.gif , the future channel gains ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq88_HTML.gif ) would be needed, in principle. However, as https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq89_HTML.gif this noncasuality requirement vanishes: by the law of large numbers, we have that
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ32_HTML.gif
(32)
and, hence, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq90_HTML.gif can be readily obtained by replacing this last expression into the constraint of (29), namely
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ33_HTML.gif
(33)
where we have defined
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ34_HTML.gif
(34)
Finally, replacing https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq91_HTML.gif into (31) yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ35_HTML.gif
(35)
and, by using https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq92_HTML.gif into (40), the quantization noise for the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq93_HTML.gif th sensor node reads:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ36_HTML.gif
(36)
which evidences that the encoding rate is constant over timeslots (as initially assumed) and over sensors too.

4.1. Average Distortion in the Reconstructed Random Field

By inserting https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq94_HTML.gif into the original cost function of (26), the distortion for an arbitrary point in the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq95_HTML.gif th network segment reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ37_HTML.gif
(37)
Interestingly, distortion is not a function of the channel gain experienced by the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq96_HTML.gif th sensor in timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq97_HTML.gif (i.e., distortion does not depend on https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq98_HTML.gif ). As a result and unlike in QEDC encoding, the distortion experienced in every timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq99_HTML.gif is identical. This can be useful in applications where a constant distortion level is needed.
After some tedious manipulations, the average distortion in the entire reconstructed random field can be expressed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ38_HTML.gif
(38)

4.2. Buffer Stability Considerations

In order to derive a closed-form solution of the optimal number of samples to be encoded in each time slot ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq100_HTML.gif ), in (32) we let the number of timeslots https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq101_HTML.gif grow to infinity. Clearly, this might lead to a situation were buffer occupancy grows without bound, that is, to buffer unstability. To avoid that, we will encode and transmit a (slightly) higher number of samples per timeslot, namely
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ39_HTML.gif
(39)
with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq102_HTML.gif . By doing so, one can prove (see the appendix) that buffers are stable. Unsurprisingly, this come at the expense of an increased distortion in the estimates (see computer simulation results in Section 7).

5. Delay-Tolerant WSNs with Compress-and-Estimate Encoding

As in previous section, we let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq103_HTML.gif be the number of samples encoded in https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq104_HTML.gif channel uses (i.e., one timeslot). Again, the rate at the output of the C&E encoder must satisfy
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ40_HTML.gif
(40)
To stress that expression (40) differs from (25) in that the C&E encoder assumes that the FC will use https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq105_HTML.gif to decode https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq106_HTML.gif and, hence, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq107_HTML.gif has been replaced by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq108_HTML.gif . Therefore, from (7) and the definition of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq109_HTML.gif in (20), we have that for the current block of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq110_HTML.gif samples the distortion reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ41_HTML.gif
(41)
By averaging over https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq111_HTML.gif timeslots, the following optimization problem results:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ42_HTML.gif
(42)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ43_HTML.gif
(43)
Solving this problem leads to a closed-form solution that is identical to that of the QEDT case, namely,
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ44_HTML.gif
(44)
Finally, replacing https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq112_HTML.gif into (40) yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ45_HTML.gif
(45)
that is, the encoding rate in CEDT networks is constant over sensors and timeslots, as implicitly assumed in the score function (43). To remark, the stability analysis of Section 4.2 also applies here.

5.1. Average Distortion in the Reconstructed Random Field

By inserting https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq113_HTML.gif into the original cost function of (43), the distortion for an arbitrary point in the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq114_HTML.gif th segment reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ46_HTML.gif
(46)
As in the QEDT case, distortion is not a function of the channel gain experienced by the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq115_HTML.gif th sensor in timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq116_HTML.gif . Hence, the distortion experienced in every timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq117_HTML.gif is identical. Therefore, the average distortion for each network segment can be computed in a closed form as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ47_HTML.gif
(47)
Finally, the average distortion in the whole reconstructed random field yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ48_HTML.gif
(48)

6. Latency Analysis

In delay-tolerant networks, each sensor encodes and transmits a variable number of samples per timeslot. As a result, the time elapsed until the FC receives the first https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq118_HTML.gif samples from all the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq119_HTML.gif sensors in the network (which allows for the reconstruction of the first https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq120_HTML.gif realizations of the random field) is unavoidably larger than in delay-constrained networks. In this section, we attempt to characterize such latency. To that aim, we start by analyzing the time needed for one sensor to transmit https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq121_HTML.gif consecutive samples of the random field. Next, we derive the latency of the QEDT and CEDT encoding strategies, respectively.

6.1. Latency Analysis for a Single Sensor Node

Let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq122_HTML.gif be the number of samples encoded in https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq123_HTML.gif channel uses in timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq124_HTML.gif . The probability that https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq125_HTML.gif samples are encoded in arbitrary timeslot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq126_HTML.gif can be expressed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ49_HTML.gif
(49)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ50_HTML.gif
(50)
Besides, we define
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ51_HTML.gif
(51)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ52_HTML.gif
(52)
On that basis, we model our system as an absorbing Markov chain [16, Chapter 8] with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq127_HTML.gif transient states ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq128_HTML.gif ) and one absorbing state ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq129_HTML.gif ) defined as follows (see, Figure 3):
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ53_HTML.gif
(53)
The transition matrix https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq130_HTML.gif of an absorbing Markov chain has the following canonical form:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ54_HTML.gif
(54)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq131_HTML.gif denotes the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq132_HTML.gif transient matrix and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq133_HTML.gif is a https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq134_HTML.gif nonzero vector (otherwise the absorbing state could never be reached from the transient states). The entries of the matrix https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq135_HTML.gif can be computed as follows:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ55_HTML.gif
(55)
The entries of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq136_HTML.gif https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq137_HTML.gif vector, which denote the probability of absorbtion from each transient states, are given by
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ56_HTML.gif
(56)
Our goal is to characterize the time elapsed until the absorbing state is reached or, in other words, the time needed to transmit https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq138_HTML.gif consecutive samples of the local observation of the random field at sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq139_HTML.gif (i.e., sensor latency). For an absorbing Markov chain, the time to absorbtion, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq140_HTML.gif , is a random variable which obeys the so-called Discrete Phase-type (DPH) distribution. From [17], the probability mass and cumulative distribution functions can be expressed as:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ57_HTML.gif
(57)
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ58_HTML.gif
(58)
where the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq141_HTML.gif vector https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq142_HTML.gif is used to define the initial conditions. Since we assume that initally no samples have been transmitted, this yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ59_HTML.gif
(59)
From all the above, the average time to absorbtion reads:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ60_HTML.gif
(60)
Alternatively, from [16, Chapter 8], one can compute
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ61_HTML.gif
(61)
the elements of which account the average time to absorbtion from state https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq143_HTML.gif . Consequently, the average sensor latency is given by its first element, namely, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq144_HTML.gif .
Finally, we need to derive a closed-form expression for the set of probabilities https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq145_HTML.gif defined in (50) and (52). From (35), we have that
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ62_HTML.gif
(62)
with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq146_HTML.gif and, hence,
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ63_HTML.gif
(63)
for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq147_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq148_HTML.gif . For Rayleigh-fading channels, the CDF of the channel gain is given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq149_HTML.gif .

6.2. Latency Analysis for QEDT Encoding

At this point, the interest lies in characterizing the time elapsed until the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq150_HTML.gif sensors in the network encode and transmit their first https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq151_HTML.gif samples of the random field. Let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq152_HTML.gif be a random variable which accounts for QEDT latency, namely
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ64_HTML.gif
(64)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq153_HTML.gif stands for the latency associated to the individual sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq154_HTML.gif as defined in the previous section. Since, on the one hand, sensors experience i.i.d fading channels and, on the other, codewords from different sensors are decoded independently, then https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq155_HTML.gif turn out to be i.i.d. DPH random variables with marginal pmf's and CDFs given by (57) and (58), respectively. From all the above, the CDF of the latency associated to QEDT encoding reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ65_HTML.gif
(65)
The probability mass function can be computed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ66_HTML.gif
(66)
and, from this last expression, the average latency yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ67_HTML.gif
(67)
Intuitively, latency is a monotonically increasing function in the number of sensors (the more sensors, the larger the time needed to collect all samples). This extent will be verified in Section 7 (Simulation and numerical results).

6.3. Latency Analysis for CEDT Encoding

The latency analysis for CEDT strategies if far more involved due to the successive encoding of data that C&E schemes entail. In general, this does not allow for the derivation of closed-form expressions and, thus, we will resort to an approximate (yet accurate) model.
In order for the FC to successfully decode the codeword received from sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq156_HTML.gif , the codeword sent by the adjacent sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq157_HTML.gif must have been decoded first. Consequently, the codeword sent by the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq158_HTML.gif th sensor will be the last one to be decoded. Since sensors experience i.i.d. fading channels (and, thus, the number of observations received from different sensors are not time-aligned), when the first https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq159_HTML.gif samples sent by sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq160_HTML.gif are ready to be decoded, a total of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq161_HTML.gif samples from sensor https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq162_HTML.gif have already been decoded on average. Accordingly, a total of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq163_HTML.gif samples from sensor #1 have already been decoded too (see Figure 4). Hence, the first https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq164_HTML.gif realizations of the entire random field can be reconstructed if, equivalently, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq165_HTML.gif samples sent by the first sensor have already been decoded by the FC. The encoding/decoding process for the first sensor is identical in C&E and Q&E schemes and, hence, in order to compute the latency for the reconstruction of the random field, it suffices to compute the time to absorbtion for an individual sensor (sensor #1) as we did in Section 6.1. The only change with respect to the model given in (54) is that the Markov chain has now a total of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq166_HTML.gif states (instead of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq167_HTML.gif ) and, hence, the size and elements of matrix https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq168_HTML.gif and vectors https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq169_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq170_HTML.gif in (57) and (58) must be adjusted accordingly.
As for parameter https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq171_HTML.gif , which exclusively depends on the pdf of the sensor-to-FC channel gains, it can only be determined empirically (see next section).

7. Simulations and Numerical Results

Figure 5 depicts the (pertimeslot) distortion in the reconstructed random field for both the QEDC and QEDT encoding strategies and different SNR values. For the QEDC strategy, we show the average value along with the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq172_HTML.gif confidence interval (to recall that, unlike in the QEDT case, the distortion in QEDC encoding varies from timeslot to timeslot). Several conclusions can be drawn. First, for each curve there exists an optimal operating point; that is, a network size for which distortion can be minimized. The intuition behind this fact is that, despite that spatial variations of the random field are better captured by a denser grid of sensors, for a total bandwidth constraint the available rate per sensor progressively diminishes, this resulting into a more rough quantization of the observations. Thus, the optimal trade-off between these two effects needs to be identified. Second, the distortion associated to delay-tolerant strategies is, as expected, lower than for delay-constrained ones. Moreover, the lower the average SNR in the sensor-to-FC channels (namely, sensors with lower transmit power), the higher the gain (up to 3 dB for SNR = 0 dB). Third, guaranteing buffer stability in the QEDT scheme only results into a marginal penalty in distortion, as shown in the curves labeled with https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq173_HTML.gif and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq174_HTML.gif . Complementarily, in Figure 6, we depict buffer occupancy for several values of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq175_HTML.gif . For https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq176_HTML.gif , the system is clearly unstable. Conversely, by letting https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq177_HTML.gif take positive values, for example, for https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq178_HTML.gif as in Figure 5, the average buffer occupancy can be kept under control (with a relatively small average buffer occupancy of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq179_HTML.gif samples, in this case). Clearly, increasing https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq180_HTML.gif has a two-fold effect: the average buffer occupancy diminishes but, simultaneously, the resulting distortion increases.
The rate at which distortion decreases for the QEDC and QEDT schemes (evaluated at their respective optimal operating points) for an increasing SNR is shown in Figure 7. For intermediate distortion values, the gap is approximately 4 dB. That is, for a prescribed distortion level, the energy consumption in delay-constrained networks is 2.5 times higher.
Figure 8 illustrates the average distortion in the reconstructed random field for the CEDC and CEDT encoding strategies. As in quantize-and-estimate encoding, there exists an optimal number of sensors nodes. Finding such https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq188_HTML.gif reveals particularly useful for random fields with low https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq189_HTML.gif per sensor, since the curve is sharper in this case. The gap between the minimum distortion attainable by the CEDC and CEDT schemes (which results from an adequate exploitation of channel fluctuation in the delay-tolerant approach) is approximately 2-3 dB. Concerning buffer occupancy-distortion tradeoffs, the same comments as in the quantize-and-estimate case apply.
Next, in Figure 9, we compare the distortion attained by QEDT/CEDT encoding strategies for random fields with low and high spatial variabilities ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq192_HTML.gif , https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq193_HTML.gif , resp.). Due to the fact that CEDT is capable of exploiting spatial correlation, it always outperforms QEDT. Moreover, the higher the spatial correlation ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq194_HTML.gif ), the larger the gap between the curves.
Finally, in Figures 10 and 11 we depict the average latency for the QEDT and CEDT strategies, respectively. Interestingly, there exists a trade-off in terms of attainable distortion versus latency. Whereas in CEDT encoding latency exhibits a linear increase in the number of sensors, in QEDT encoding latency grows logarithmically (i.e., more slowly). However, CEDT schemes attain a lower distortion than QEDT ones. Besides, in Figure 10 it is also worth noting the perfect match between simulations and numerical results and, unsurprisingly, that the higher the average https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq197_HTML.gif , the lower the latency. Also, Figure 11 reveals that by using an appropriate value of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq198_HTML.gif (i.e., https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq199_HTML.gif ), the latency associated to the approximate model described in Section 6.3 matches the actual one.

8. Conclusions

In this paper, we have extensively analyzed the problem of random field estimation with wireless sensor networks. In order to characterize the dynamics and spatial correlation of the random field, we have adopted a stationary homogeneous Gaussian Markov Ornstein-Uhlenbeck model. We have considered two scenarios of interest: delay-constrained (DC) and delay-tolerant (DT) networks. For each scenario, we have analyzed two encoding schemes, namely, quantize-and-estimate (QE) and compress-and-estimate (CE). In all cases (QEDC, QEDT, CEDC and CEDT), we have carried out an extensive analysis of the average distortion experienced in the reconstructed random field. Moreover, for the QEDT and CEDT strategies we have derived closed-form expressions of (i) the average distortion in the estimates, and (ii) the optimal number of samples of the random field to be encoded in each timeslot (under some simplifying assumptions). Interestingly, the resulting pertimeslot distortion in DT scenarios is deterministic and constant whereas, in DC scenarios, it ultimately depends on the fading conditions experienced in each timeslot. Next, we have focused on the latency associated to the QEDT and CEDT strategies. We have modeled our system as an absorbing Markov chain and, on that basis, we have fully characterized the pdf, CDF, and the average latency for the QEDT case. For CEDT encoding, we have identified an approximate system model suitable for the computation of the average latency. Simulation results reveal that, under a total bandwidth constraint, there exists an optimal number of sensors for which the distortion in the reconstructed random field can be minimized (QEDC, QEDT, CEDC and CEDT cases). This constitutes the best trade-off in terms of, on the one hand, the ability to capture the spatial variations of the random field and, on the other, the persensor channel bandwidth available to encode observations. Besides, the distortion associated to delay-tolerant strategies is, as expected, lower than for delay-constrained ones: some 2-3 dB for both the QE and CE encoding schemes. Moreover, buffer occupancy can be kept at very moderate levels (3 timeslots) with a marginal penalty in terms of distortion (less than 0.3 dB). We also observe that CE schemes effectively exploit the spatial correlation and, by doing so, attain a lower distortion than their QE counterparts (DC and DT scenarios). As far as latency is concerned, we have empirically shown that CEDT exhibits a linear increase in the number of sensors whereas in QEDT encoding latency grows logarithmically (i.e., more slowly). However, CEDT schemes attain a lower distortion than QEDT ones. Besides, for the QEDT case, there is a perfect match between simulations and the theoretical model and, for the CEDT case, latency can be accurately represented by adequately parameterizing the aforementioned approximate system model.

Acknowledgment

This work is partly supported by the Catalan Government (2009 SGR 1046), the EC-funded project NEWCOM++ (216715), and the Spanish Ministry of Science and Innovation (FPU grant AP2007-01654).
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://​creativecommons.​org/​licenses/​by/​2.​0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Appendix

Appendix

Buffer Stability Analysis
We want to prove that buffers are stable (i.e., their occupancy is bounded) for large https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq201_HTML.gif . Let https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq202_HTML.gif denote the number of samples in the buffer of the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq203_HTML.gif th sensor in time slot https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq204_HTML.gif , with initial conditions given by https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq205_HTML.gif . After https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq206_HTML.gif timeslots, the increase in the number of samples stored in the buffer can be expressed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ68_HTML.gif
(A.1)
where https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq207_HTML.gif accounts for the number of samples generated in those https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq208_HTML.gif timeslots, and https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq209_HTML.gif with
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ69_HTML.gif
(A.2)
stands for the actual number of samples encoded and transmitted by the https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq210_HTML.gif th sensor node. The probability of experiencing an increase greater than https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq211_HTML.gif in the number of samples stored reads
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ70_HTML.gif
(A.3)
for any https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq212_HTML.gif . Replacing (A.2) into this last expression yields:
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ71_HTML.gif
(A.4)
where we have defined
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ72_HTML.gif
(A.5)
For large https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq213_HTML.gif , we can resort to the central limit theorem by which
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ73_HTML.gif
(A.6)
Hence, as long as https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq214_HTML.gif takes strictly positive values ( https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq215_HTML.gif ), we have that
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ74_HTML.gif
(A.7)
This result states that, as long as we encode a slightly higher number of samples per timeslot (which depends on parameter https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq216_HTML.gif ) the probability that the increase in buffer occupancy exceeds https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq217_HTML.gif samples (for a finite value of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq218_HTML.gif ) can be made arbitrary small for large https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq219_HTML.gif . That is, buffers are stable. Conversely, https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq220_HTML.gif yields
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ75_HTML.gif
(A.8)
this meaning that, even for arbitrarily large values of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq221_HTML.gif , the probability that buffer occupancy increases beyond https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq222_HTML.gif is unavoidably https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq223_HTML.gif (i.e., unstable buffers).
In addition to this main result, the probability for buffers to drain after https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq224_HTML.gif timeslots can be expressed as
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ76_HTML.gif
(A.9)
By resorting again to the central limit theorem, we have that for any positive value of https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq225_HTML.gif
https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_Equ77_HTML.gif
(A.10)
and, thus, buffers will drain with probability one after a sufficiently large number of timeslots.
Literature
1.
go back to reference Akyildiz IF, Su W, Sankarasubramaniam Y, Cayirci E: Wireless sensor networks: a survey. Computer Networks 2002, 38(4):393-422. 10.1016/S1389-1286(01)00302-4CrossRef Akyildiz IF, Su W, Sankarasubramaniam Y, Cayirci E: Wireless sensor networks: a survey. Computer Networks 2002, 38(4):393-422. 10.1016/S1389-1286(01)00302-4CrossRef
2.
go back to reference Vuran MC, Akyildiz IF: Spatial correlation-based collaborative medium access control in wireless sensor networks. IEEE/ACM Transactions on Networking 2006, 14(2):316-329.CrossRef Vuran MC, Akyildiz IF: Spatial correlation-based collaborative medium access control in wireless sensor networks. IEEE/ACM Transactions on Networking 2006, 14(2):316-329.CrossRef
3.
go back to reference Draper SC, Wornell GW: Side information aware coding strategies for sensor networks. IEEE Journal on Selected Areas in Communications 2004, 22(6):966-976. 10.1109/JSAC.2004.830875CrossRef Draper SC, Wornell GW: Side information aware coding strategies for sensor networks. IEEE Journal on Selected Areas in Communications 2004, 22(6):966-976. 10.1109/JSAC.2004.830875CrossRef
4.
go back to reference Wyner AD, Ziv J: The rate-distortion function for source coding with side information at the decoder. IEEE Transactions on Information Theory 1976, 22(1):1-10. 10.1109/TIT.1976.1055508MathSciNetCrossRefMATH Wyner AD, Ziv J: The rate-distortion function for source coding with side information at the decoder. IEEE Transactions on Information Theory 1976, 22(1):1-10. 10.1109/TIT.1976.1055508MathSciNetCrossRefMATH
5.
go back to reference Ishwar P, Kumar A, Ramchandran K: Distributed sampling for dense sensor networks: a bit-conservation principle. Proceedings of the 2nd International Workshop on Information Processing in Sensor Networks, April 2003, Lecture Notes in Computer Science 2634: 17-31.CrossRefMATH Ishwar P, Kumar A, Ramchandran K: Distributed sampling for dense sensor networks: a bit-conservation principle. Proceedings of the 2nd International Workshop on Information Processing in Sensor Networks, April 2003, Lecture Notes in Computer Science 2634: 17-31.CrossRefMATH
6.
go back to reference Dardari D, Conti A, Buratti C, Verdone R: Mathematical evaluation of environmental monitoring estimation error through energy-efficient wireless sensor networks. IEEE Transactions on Mobile Computing 2007, 6(7):790-802.CrossRef Dardari D, Conti A, Buratti C, Verdone R: Mathematical evaluation of environmental monitoring estimation error through energy-efficient wireless sensor networks. IEEE Transactions on Mobile Computing 2007, 6(7):790-802.CrossRef
7.
go back to reference Dogandžić A, Qiu K: Decentralized random-field estimation for sensor networks using quantized spatially correlateddata and fusion-center feedback. IEEE Transactions on Signal Processing 2008, 56(12):6069-6085.MathSciNetCrossRef Dogandžić A, Qiu K: Decentralized random-field estimation for sensor networks using quantized spatially correlateddata and fusion-center feedback. IEEE Transactions on Signal Processing 2008, 56(12):6069-6085.MathSciNetCrossRef
8.
go back to reference Dong M, Tong L, Sadler BM: Impact of data retrieval pattern on homogeneous signal field reconstruction in dense sensor networks. IEEE Transactions on Signal Processing 2006, 54(11):4352-4364.CrossRef Dong M, Tong L, Sadler BM: Impact of data retrieval pattern on homogeneous signal field reconstruction in dense sensor networks. IEEE Transactions on Signal Processing 2006, 54(11):4352-4364.CrossRef
9.
go back to reference Marco D, Neuhoff DL: Reliability vs. efficiency in distributed source coding for field-gathering sensor networks. Proceedings of the 3rd International Symposium on Information Processing in Sensor Networks (IPSN '04), April 2004, Berkeley, Calif, USA 161-168. Marco D, Neuhoff DL: Reliability vs. efficiency in distributed source coding for field-gathering sensor networks. Proceedings of the 3rd International Symposium on Information Processing in Sensor Networks (IPSN '04), April 2004, Berkeley, Calif, USA 161-168.
10.
12.
go back to reference Kay SM: Fundamentals of Statistical Signal Processing: Estimation Theory, Prentice-Hall Signal Processing Series. Prentice-Hall, Englewood Cliffs, NJ, USA; 1993.MATH Kay SM: Fundamentals of Statistical Signal Processing: Estimation Theory, Prentice-Hall Signal Processing Series. Prentice-Hall, Englewood Cliffs, NJ, USA; 1993.MATH
13.
go back to reference Cover TM, Thomas JA: Elements of Information Theory, Wiley Series in Telecommunications. Wiley, New York, NY, USA; 1993. Cover TM, Thomas JA: Elements of Information Theory, Wiley Series in Telecommunications. Wiley, New York, NY, USA; 1993.
14.
go back to reference Ishwar P, Puri R, Ramchandran K, Pradhan SS: On rate-constrained distributed estimation in unreliable sensor networks. IEEE Journal on Selected Areas in Communications 2005, 23(4):765-775.CrossRefMATH Ishwar P, Puri R, Ramchandran K, Pradhan SS: On rate-constrained distributed estimation in unreliable sensor networks. IEEE Journal on Selected Areas in Communications 2005, 23(4):765-775.CrossRefMATH
15.
go back to reference Corless RM, Gonnet GH, Hare DEG, Jeffrey DJ, Knuth DE:On the Lambert function. Advances in Computational Mathematics 1996, 5(4):329-359.MathSciNetCrossRefMATH Corless RM, Gonnet GH, Hare DEG, Jeffrey DJ, Knuth DE:On the Lambert https://static-content.springer.com/image/art%3A10.1155%2F2010%2F102460/MediaObjects/13638_2010_Article_1789_IEq200_HTML.gif function. Advances in Computational Mathematics 1996, 5(4):329-359.MathSciNetCrossRefMATH
16.
go back to reference Mayer CD: Matrix Analysis and Applied Linear Algebra. SIAM; 2001. Mayer CD: Matrix Analysis and Applied Linear Algebra. SIAM; 2001.
17.
go back to reference Neuts MF: Matrix-Geometric Solutions in Stochastic Models: An Algorithmic Approach, Chapter 2: Probability Distributionsof Phase Type. Dover; 1981. Neuts MF: Matrix-Geometric Solutions in Stochastic Models: An Algorithmic Approach, Chapter 2: Probability Distributionsof Phase Type. Dover; 1981.
Metadata
Title
Random Field Estimation with Delay-Constrained and Delay-Tolerant Wireless Sensor Networks
Authors
Javier Matamoros
Carles Antón-Haro
Publication date
01-12-2010
Publisher
Springer International Publishing
DOI
https://doi.org/10.1155/2010/102460

Other articles of this Issue 1/2010

EURASIP Journal on Wireless Communications and Networking 1/2010 Go to the issue

Research Article

Field Division Routing

Premium Partner