12
2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information. This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEE Transactions on Control of Network Systems 1 Distributed Optimal Control of Sensor Networks for Dynamic Target Tracking Greg Foderaro, Member, IEEE, Pingping Zhu, Member, IEEE, Hongchuan Wei, Student Member, IEEE, Thomas A. Wettergren, Senior Member, IEEE, and Silvia Ferrari, Senior Member, IEEE Abstract—This paper presents a distributed optimal control approach for managing omnidirectional sensor networks de- ployed to cooperatively track moving targets in a region of inter- est. Several authors have shown that, under proper assumptions, the performance of mobile sensors is a function of the sensor distribution. In particular, the probability of cooperative track detection, also known as track coverage, can be shown to be an integral function of a probability density function representing the macroscopic sensor network state. Thus, a mobile sensor network deployed to detect moving targets can be viewed as a multiscale dynamical system in which a time-varying probability density function can be identified as a restriction operator, and optimized subject to macroscopic dynamics represented by the advection equation. Simulation results show that the distributed control approach is capable of planning the motion of hundreds of cooperative sensors, such that their effectiveness is significantly increased compared to that of existing uniform, grid, random and stochastic gradient methods. Index Terms—Optimal Control, Distributed Control, Mobile Sensor Networks, Track Coverage, Target Tracking, Multiscale Dynamical Systems. I. I NTRODUCTION T HIS paper presents a distributed optimal control (DOC) approach for optimizing the trajectories of a network of many cooperative mobile sensors deployed to perform track detection in a region of interest (ROI). Considerable attention has been given to the problem of controlling mobile sensors in order to maximize coverage in a desired ROI, as required when no prior target information is available [1]–[10]. When prior information such as target measurements or expert knowledge are available, optimal control and information-driven strategies have been shown to significantly outperform other methods [9]–[16]. Due to the computational complexity associated with solving the optimality conditions and evaluating information theoretic functions, however, these methods typically do not scale to networks with hundreds of sensors because the com- putation they require increases exponentially with the number of agents [17]. Distributed optimal control has been recently shown to over- come the computational complexity associated with classical optimal control for systems in which the network performance or cost function is a function of a suitable restriction operator, G. Foderaro, H. Wei are with the Department of Mechanical Engineering and Materials Science, Duke University, Durham, NC, 27708. P. Zhu, S. Wahab, and S. Ferrari are with the Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY, 14853 USA e- mail: ([email protected], [email protected], [email protected]). T. A. Wettergren is with the Naval Undersea Warfare Center, Newport, RI, 02841 USA e-mail: ([email protected]) such as a probability density function (PDF) or maximum likelihood estimator (MLE) [18]–[21]. Several authors have shown that, in many instances, the performance of networks of cooperative agents, such as sensors and robotic vehicles, is a function of a PDF representing the density of the agents over the ROI [22]–[26]. Thus, one approach that has been proposed to deploy many cooperative agents is to sample a known PDF to obtain a set of agent positions in the ROI [26]. Another approach is to use a known PDF to perform locational optimization, and obtain a corresponding network representation using centroidal Voronoi partitions [27], [28]. Alternatively, agent trajectories can be computed using a hierarchical control approach that first establishes a virtual, adaptive network boundary, and then computes the agent control inputs to satisfy the boundary in a lower-dimensional space [29]. While these existing approaches are effective at reducing the dimensionality of an otherwise intractable optimal con- trol problem, they assume that the optimal PDF (or virtual boundary) are given a priori. As a result, the agents may be unable to reach the desired PDF when in the presence of dynamic constraints and/or inequality constraints on the state and controls. Conversely, if a conservative PDF is given to guarantee reachability, the network performance may be suboptimal. Furthermore, because existing methods assume stationary agent distributions, they cannot fully exploit the capabilities of mobile sensors, or take into account time- varying environmental conditions. The DOC approach, re- cently developed by the authors in [19], overcomes these limitations by optimizing a time-varying agent PDF subject to the agent dynamics. To date, DOC optimality conditions have been derived and used to solve network control problems in multi-agent path planning and navigation [19], [21]. This paper presents con- servation law results that show the closed-loop DOC system is Hamiltonian. Based on these results, an efficient numerical so- lution is obtained using a finite volume discretization scheme that has a computational complexity far reduced compared to classical optimal control. The DOC method is then applied to a network control problem in which omnidirectional sensors are deployed to cooperatively detect moving targets in an obstacle- populated ROI. Several authors have shown that the tracking and detection capability of many real sensor networks, such as passive acoustic sensors, can be represented in closed-form by assuming the sensors are omnidirectional and prior targets measurements can be assimilated into Markov motion models (see [10], [11], [30]–[33] and references therein). In this paper,

Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

1

Distributed Optimal Control of Sensor Networks forDynamic Target Tracking

Greg Foderaro, Member, IEEE, Pingping Zhu, Member, IEEE, Hongchuan Wei, Student Member, IEEE,Thomas A. Wettergren, Senior Member, IEEE, and Silvia Ferrari, Senior Member, IEEE

Abstract—This paper presents a distributed optimal controlapproach for managing omnidirectional sensor networks de-ployed to cooperatively track moving targets in a region of inter-est. Several authors have shown that, under proper assumptions,the performance of mobile sensors is a function of the sensordistribution. In particular, the probability of cooperative trackdetection, also known as track coverage, can be shown to be anintegral function of a probability density function representingthe macroscopic sensor network state. Thus, a mobile sensornetwork deployed to detect moving targets can be viewed as amultiscale dynamical system in which a time-varying probabilitydensity function can be identified as a restriction operator, andoptimized subject to macroscopic dynamics represented by theadvection equation. Simulation results show that the distributedcontrol approach is capable of planning the motion of hundredsof cooperative sensors, such that their effectiveness is significantlyincreased compared to that of existing uniform, grid, random andstochastic gradient methods.

Index Terms—Optimal Control, Distributed Control, MobileSensor Networks, Track Coverage, Target Tracking, MultiscaleDynamical Systems.

I. INTRODUCTION

THIS paper presents a distributed optimal control (DOC)approach for optimizing the trajectories of a network of

many cooperative mobile sensors deployed to perform trackdetection in a region of interest (ROI). Considerable attentionhas been given to the problem of controlling mobile sensors inorder to maximize coverage in a desired ROI, as required whenno prior target information is available [1]–[10]. When priorinformation such as target measurements or expert knowledgeare available, optimal control and information-driven strategieshave been shown to significantly outperform other methods[9]–[16]. Due to the computational complexity associated withsolving the optimality conditions and evaluating informationtheoretic functions, however, these methods typically do notscale to networks with hundreds of sensors because the com-putation they require increases exponentially with the numberof agents [17].

Distributed optimal control has been recently shown to over-come the computational complexity associated with classicaloptimal control for systems in which the network performanceor cost function is a function of a suitable restriction operator,

G. Foderaro, H. Wei are with the Department of Mechanical Engineeringand Materials Science, Duke University, Durham, NC, 27708.

P. Zhu, S. Wahab, and S. Ferrari are with the Sibley School of Mechanicaland Aerospace Engineering, Cornell University, Ithaca, NY, 14853 USA e-mail: ([email protected], [email protected], [email protected]).

T. A. Wettergren is with the Naval Undersea Warfare Center, Newport, RI,02841 USA e-mail: ([email protected])

such as a probability density function (PDF) or maximumlikelihood estimator (MLE) [18]–[21]. Several authors haveshown that, in many instances, the performance of networksof cooperative agents, such as sensors and robotic vehicles,is a function of a PDF representing the density of the agentsover the ROI [22]–[26]. Thus, one approach that has beenproposed to deploy many cooperative agents is to sample aknown PDF to obtain a set of agent positions in the ROI[26]. Another approach is to use a known PDF to performlocational optimization, and obtain a corresponding networkrepresentation using centroidal Voronoi partitions [27], [28].Alternatively, agent trajectories can be computed using ahierarchical control approach that first establishes a virtual,adaptive network boundary, and then computes the agentcontrol inputs to satisfy the boundary in a lower-dimensionalspace [29].

While these existing approaches are effective at reducingthe dimensionality of an otherwise intractable optimal con-trol problem, they assume that the optimal PDF (or virtualboundary) are given a priori. As a result, the agents maybe unable to reach the desired PDF when in the presenceof dynamic constraints and/or inequality constraints on thestate and controls. Conversely, if a conservative PDF is givento guarantee reachability, the network performance may besuboptimal. Furthermore, because existing methods assumestationary agent distributions, they cannot fully exploit thecapabilities of mobile sensors, or take into account time-varying environmental conditions. The DOC approach, re-cently developed by the authors in [19], overcomes theselimitations by optimizing a time-varying agent PDF subjectto the agent dynamics.

To date, DOC optimality conditions have been derived andused to solve network control problems in multi-agent pathplanning and navigation [19], [21]. This paper presents con-servation law results that show the closed-loop DOC system isHamiltonian. Based on these results, an efficient numerical so-lution is obtained using a finite volume discretization schemethat has a computational complexity far reduced compared toclassical optimal control. The DOC method is then applied to anetwork control problem in which omnidirectional sensors aredeployed to cooperatively detect moving targets in an obstacle-populated ROI. Several authors have shown that the trackingand detection capability of many real sensor networks, suchas passive acoustic sensors, can be represented in closed-formby assuming the sensors are omnidirectional and prior targetsmeasurements can be assimilated into Markov motion models(see [10], [11], [30]–[33] and references therein). In this paper,

Page 2: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

2

the DOC approach is used to optimize the detection capabilityof this class of sensor networks for hundreds of cooperativeagents, while also minimizing their energy consumption andavoiding collisions with the obstacles. The results show thatthe DOC approach significantly improves the probability ofdetection compared to other scalable strategies known asuniform, grid, random and stochastic gradient methods [34],[35].

II. PROBLEM FORMULATION AND ASSUMPTIONS

This paper considers the problem of optimizing the stateand control trajectories of a network of N mobile sensorsused to detect a moving target in an obstacle-populated ROI,A = [0, L] × [0, L] ⊂ R2, during a fixed time interval,t ∈ (T0, Tf ], where T0 and Tf are both given. Each sensor ismounted on a robot or vehicle whose motion is governed bya small system of ODEs,

si(t) = f [si(t),ui(t), t], si(T0) = si0 , i = 1, . . . , N (1)

where si(t) = [xTi (t), θi(t)]

T ∈ S is the ith vehicle state com-prised of the vehicle position, xi = [xi, yi]

T ∈ A, and headingangle θi ∈ [0, 2π), ui ∈ U ⊂ Rm is the vehicle control vector,and U is the space of m admissible control inputs. Here, thesuperscript “T ” denotes the transpose of matrices and vectors.Each sensor is assumed to be omnidirectional, with a constanteffective range r ∈ R, defined as the maximum range at whichthe received signal exceeds a desired threshold [9]. Then, thefield-of-view (FOV) of every sensor can be modeled by a diskC(xi, r) ⊂ R2, with radius r and center at xi.

Since xi, i = 1, . . . , N , is a time-varying continuous vector,let ℘x denote the time-varying PDF of xi, defined as a non-negative function that satisfies the normalization property,∫

A℘x(x, t)dx = 1 (2)

and such that the probability of event x ∈ B ⊂ A is,

P (x ∈ B, t) =∫B℘x(x, t)dx (3)

where, B is any subset of the ROI, and, for brevity, ℘x

is abbreviated to ℘ in the remainder of this paper. By thisapproach, each sensor can be viewed as a fluid particle in theLagrangian approach, and ℘ can be viewed as the forwardPDF of particle position [36]. Therefore, N℘ represents thedensity of sensors in A.

There is considerable precedence in both target tracking lit-erature and practice for modeling target dynamics by Markovmotion models that assimilate multiple, distributed sensormeasurements [37], [38]. These tracking algorithms have theability to incrementally update the target model over timeand output Markov transition probability density functions(PDFs) that describe the uncertainty associated with the targetbased on prior sensor measurements. This paper shows that thetarget PDFs obtained by the tracking algorithms can be usedas feedback to a distributed optimal control algorithm, suchthat the sensor motion can be planned in order to maximizethe expected number of target detections over a desired timeinterval. Subsequently, the target PDFs can be updated to

reflect the new knowledge obtained by the sensor networkcontrolled via DOC.

Let the target (T) motion be described by the unicyclekinematic equations,

xT (t) =

[xT (t)yT (t)

]=

[vT (t) cos θT (t)vT (t) sin θT (t)

], t ∈ (T0, Tf ]

(4)where xT (t) = [xT (t), yT (t)]

T ∈ A is the target state, vT (t)is the target velocity, and θT (t) is the target heading angle.Because many vehicles and targets of interest move at constantheading over some period of time, Markov motion modelsassume that the target heading and velocity are constant duringa sequence of time sub-intervals, (tj , tj+1] ⊂ (T0, Tf ], j =1, . . . ,m, that together comprise an exact cover of (T0, Tf ]. Atany time tj , j = 1, . . . ,m, the target may change both headingand velocity, and, thus, tj is also referred to as maneuveringtime. Then, letting xTj

, xT (tj), θTj, θT (tj), and vTj

,vT (tj), and integrating (4) with respect to time, the target canbe described by the motion model,

xTj+1 = xTj + [vTj cos θTj vTj sin θTj ]T∆tj , (5)

where ∆tj = tj+1 − tj , and j = 1, . . . ,m.Because the actual target track is unknown a priori, the

track parameters vTj, θTj

, and xTjcan be viewed as random

variables [37], [38]. Then, assuming for simplicity that they areindependent random variables, prior target information can beprovided in terms of target PDFs, fTj (x), fΘj (θ), and fVj (v),which are typically computed by target tracking algorithmsbased on prior sensor measurements [39] or, in the absence ofprior information, are assumed uniform.

For an omnidirectional sensor, the probability of targetdetection for a sensor at x can be described by the Booleandetection model

Ps[x(t),xT (t)] =

1, ‖x(t)− xT (t)‖ ≤ r

0, ‖x(t)− xT (t)‖ > r,(6)

where ‖·‖ denotes the Euclidean norm and a perfect detectionmodel is adopted [26], [31]. It means that the target can bedeclared a successful detection once the target belongs to anyFOV of the N sensors.

The problem considered in this paper is to optimally controlthe N omnidirectional sensors such that obstacle collisions areavoided, the probability of track detection in A is maximized,and the energy consumption is minimized subject to theequation of motion (1). The next sections shows how thisproblem can be formulated as a DOC problem and, then,solved efficiently for up to hundreds of sensors using theconservation law analysis and numerical method presented inSections V-VI.

III. PROBABILITY OF TARGET DETECTION

In this section, an objective function representing thequality-of-service of the sensor network is obtained from theprobability of target track detection and the coverage cone inthe spatio-temporal space, Ω = A× (T0, Tf ], which was firstpresented in [11]. Based on the Markov motion model, thetarget track as it evolves from time tj to t, tj < t ≤ tj+1,

Page 3: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

3

can be represented by a vector mj(t) defined in Ω and withorigin at zj = [xT

Tj, tj ]

T ∈ Ω, as shown in Fig. 1. From(6), the ith sensor is able to detect a target if and onlyif ‖xT (t) − xi(t)‖ ≤ r. Thus, as the sensor moves overtime along a trajectory xi(t), the set of all tracks detected iscontained by a time-varying three-dimensional coverage coneK(t) in Ω defined according to the following remark, takenfrom [11]:

Remark 3.1: The coverage cone defined as,

K(t) =

[x y z]T ∈ Ω ⊂ R3

∣∣ z > tj , t ∈ (tj , tj+1]∥∥∥∥[x y]T − (z − tj)

(t− tj)

[xi(t)− xTj

]− xTj

∥∥∥∥ ≤ (z − tj)

(t− tj)r

(7)

contains the set of all target tracks that intersect the ith sensor’sFOV, C(t), at any time t ∈ (tj , tj+1].

For more details and the proof of remark 3.1, the reader isreferred to [11].

v

tj

xi(t) xT(t)

Fig. 1. Example of 3-D spatio-temporal coverage cone, where the blue curveis the trajectory of the target xT (t). The heading-cone (yellow), velocity-cone (cyan), representation of coverage cone (magenta), and the correspondingangles are all indicated.

An example of three-dimensional (3D) spatio-temporal cov-erage cone, K(t), is shown in Fig. 1 as a magenta cone.Because (7) is a circular cone that is possibly oblique, it isdifficult to define a Lebesgue measure of the tracks containedby K(t) that can be computed analytically from the sensorposition and the Markov parameters. However, by extendingthe approach in [11] to a moving sensor, K(t) can be rep-resented by a pair of two-dimensional (2D) cones, referredto as heading cone and velocity cone, for which a Lebesguemeasure of the tracks detected by a sensor at xi(t) can beprovided in terms of unit vectors.

Let the 2D heading cone Kθ be defined as the projectionof K(t) onto the plane,

Ψθ = [x y z]T ∈ Ω | z = tj. (8)

such that Kθ (shown in yellow in Fig. 1) contains all possibleheadings of a target detected by the ith sensor at any timet ∈ (tj , tj+1]. Since Kθ is a 2D cone, it can be expressed asa linear combination of two unit vectors on the heading planewith respect to a local coordinate frame, Fj , such that,

Kθ[xi(t), zj ] = c1h(j)i (t) + c2l

(j)i (t) | c1, c2 ≥ 0, (9)

where,

h(j)i (t) =

cosα(j)i (t) − sinα

(j)i (t)

sinα(j)i (t) cosα

(j)i (t)

0 0

d(j)(t)∥∥d(j)(t)∥∥

cosλ(j)i (t)

sinλ(j)i (t)0

,l(j)i (t) =

cosα(j)i (t) sinα

(j)i (t)

− sinα(j)i (t) cosα

(j)i (t)

0 0

d(j)i (t)∥∥∥d(j)i (t)

∥∥∥≡

cos γ(j)i (t)

sin γ(j)i (t)0

(10)

d(j)i (t) ≡ xi(t)− xTj

and α(j)i (t) = sin−1(r/‖d(j)

i (t)‖).

Now, let the velocity cone Kv be defined as the intersectionof K with the velocity plane,

Ψv = [x y z]T ∈ Ω | x sin θTj− y cos θTj

= [sin θTjcos θTj

] xTj, z ≥ tj.

such that Kv represents the speeds of all targets with headingθTj (contained in Kθ) that are detected by the ith sensor att ∈ (tj , tj+1]. The velocity cone Kv can be represented bytwo unit vectors defined with respect to Fj , such that,

Kv[xi(t), zj ] = c1ξ(j)i (t) + c2ω(j)i (t) | c1, c2 ≥ 0, (11)

where,

ξ(j)i (t)=

sin η(j)i (t) cos θTj

sin η(j)i (t) sin θTj

cos η(j)i (t)

(j)i (t)=

sinµ(j)i (t) cos θTj

sinµ(j)i (t) sin θTj

cosµ(j)i (t)

,η(j)i (t)=tan−1

[ 1

t− tj

([cos θTj sin θTj ][xi(t)− xTj ]

−√r2 −

([sin θTj − cos θTj ][xi(t)− xTj ]

)2)],

µ(j)i (t)=tan−1

[ 1

t− tj

([cos θTj

sin θTj][xi(t)− xTj

]

+

√r2 −

([sin θTj

− cos θTj][xi(t)− xTj

])2)]

.

(12)An example of these coverage cone representations is illus-trated in Fig. 1.

As proven in [11], the pair of 2D time-varying cones,Kθ,Kv, can be used to represent all tracks contained by the3D time-varying coverage cone K. It follows that the proba-bility of a detection by the ith sensor at time t ∈ (tj , tj+1] isthe probability that the Markov parameters are contained by

Page 4: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

4

the heading and velocity cones, i.e.,

Pd(t) ≡P [mj(t) ∈ K(t)]

=

∫AfTj (x)

∫ λ(j)i (t)

γ(j)i (t)

fΘj (θ)

∫ tanµ(j)i (t)

tan η(j)i (t)

fVj (v)dvdθdx

(13)where the Markov motion PDFs are known from the trackingalgorithms (Section II).

IV. DISTRIBUTED OPTIMAL CONTROL PROBLEM

The control of the N omnidirectional sensors is achieved bymaximizing the probability of target detection, minimizing theenergy consumption, and avoiding collisions with obstaclesin the ROI. The energy consumption can be modeled asa quadratic function of the vehicle control vector u. Byintroducing a repulsive potential function Urep generated fromthe obstacle geometries [19], [40], the obstacle avoidanceobjective can be expressed as the product of ℘ and Urep.Then, the total sensor network performance can expressed asthe integral cost function,

J =

m∑j=1

∫ tj+1

tj

∫A[wr℘(x, t)Urep − wd℘(x, t)Pd(t) (14)

+weuTRu]dxdt ,

m∑j=1

∫ tj+1

tj

∫AL ℘(x, t),u, tdxdt

and must be minimized with respect to the network state, ℘,and control law, u = c[℘(x, t)], subject to (1),(2)-(3). Theconstant weights wd, wr, and we, are chosen by the userbased on the desired tradeoff between the sensing, obstacle-avoidance, and energy objectives, and R is a diagonal positive-definite matrix.

Because the dynamic constraints (1) are a function of thesensor (microscopic) state and control, x and u, the next stepis to determine the macroscopic evolution equation for ℘ from(1). It was shown in [19], [21] that if agents are never creatednor destroyed and are advected by a known velocity field (1),then the evolution of the PDF ℘ can be described by theadvection equation. The advection equation is a hyperbolicpartial differential equation (PDE) that governs the motion ofa conserved, scalar quantity, such as a PDF, through a knownvelocity field [41]. In the DOC problem considered in thispaper, the PDF ℘ is advected by the velocity field v = x ∈ Rn

obtained from (1), resulting in the macroscopic dynamics,

∂℘

∂t= −∇ · ℘(x, t)v = −∇ · ℘(x, t)f [x,u, t] (15)

The gradient, ∇, represents a row vector of partial derivativeswith respect to x, and ∇ ·F denotes the dot product between∇ and a vector field F, or the divergence of F.

Because the initial agent distribution is usually given, basedon the initial positions of the sensors in the ROI, the PDE (15)is subject to the initial condition

℘[x, T0] = ℘0(x) (16)

Also, in order to guarantee that agents are neither creatednor destroyed in A, the PDE (15) is subject to the boundary

condition,

℘[x ∈ ∂A, t] = 0, ∀t ∈ (T0, Tf ] (17)

the state constraints,

℘[x 6∈ A, t] = 0, ∀t ∈ (T0, Tf ], (18)

and the normalization condition (2).Furthermore, consider a square area A′ ⊂ A with side

length ∆x. With the assumption of no overlap between FOVs,the density of the sensors in the area A′ satisfy the followinginequality,

(∆x)2

πr2≥ N(∆x)2℘(x′), x′ ∈ A′ (19)

where ℘(x′) is assumed constant for a small ∆x. The righthand side (RHS) in (19) is the number of the sensors in thearea A, and the left hand side (LHS) in (19) is a upper bound.Therefore, the following constraint is obtained,

℘(x, t) ≤ 1

Nπr2. (20)

The analysis presented in the next section shows that theclosed-loop DOC problem is a Hamiltonian system and, thus,the agent PDF ℘ is conserved over time. As a result, numericalsolutions of the DOC problem can be obtained using conserva-tive numerical algorithms, such as finite volume (FV), that areknown to be computationally efficient and allow for coarse-grain discretizations without dissipation errors [42].

V. CONSERVATION LAW ANALYSIS

Hamiltonian systems are characterized by a constant ofmotion, or Hamiltonian function, by which optimal trajectoriescan be shown to have vanishing variations along this constantof motion, according to Pontryagin’s minimum principle [43],[44]. Because in the DOC problem the coarse dynamics aredescribed by the advection equation (15), the open-loop systemis inherently conservative [45]. The goal of this section is toshow that the controlled dynamics (or closed-loop system)is also conservative, by proving that it satisfies Hamiltonequations,

∂ψ

∂q= −dp

dt,∂ψ

∂p=dq

dt(21)

where ψ = ψ(p,q, t) is the Hamiltonian function, q = q(t) ∈Rn are the generalized coordinates, and p = p(t) ∈ Rn arethe generalized momenta.

For simplicity, the proof is presented for n = 2, wherex = [x y]T denotes the position of the ith agent in R2.Then, the Hamiltonian function is determined by recasting thedetailed equation (1) into a three-dimensional time-invariantODE. Letting x = [x y t]T and u(x) = u(t), (1) can bewritten as, [

x(x, u) y(x, u) t]T

= f(x, u) (22)

where, X is transformed into the time-space domain X =X × (T0, Tf ]. It also follows that the macroscopic evolutionequation (15) can be rewritten as,

∂℘(x)

∂t+∂[℘(x)x(x, u)]

∂x+∂[℘(x)y(x, u)]

∂y= 0 (23)

Page 5: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

5

where, now ℘ is only a function of x.Now, let A ≡ [Ax Ay At] = A(x) denote the vector

potential of the product (℘u), i.e.:

℘(x)u(x) = ∇×A(x) (24)

By performing a coordinate transformation to a canonicalreference frame defined such that Ay = 0, A can be used torelate the two-dimensional time-varying system to the three-dimensional time-invariant form, such that the Hamiltonianfunctions for the two forms are equivalent [45], [46]. Thecoordinate transformation is then given by F : x → x, wherex = [x p t]T , and,

p = −Ax[x, y(x, p, t), t] (25)

The resulting vector potential is A = [Ax(x, y(x, p, t), t) 0At(x, y(x, p, t), t)], which is governed by

℘ x =∂At

∂y, ℘ y =

∂Ax

∂t− ∂At

∂x, ℘ = −∂Ax

∂y(26)

where, the function y(x, p, t) is implicitly defined in (25).Then, the equivalent system is,

dx

dt= f(x) =

[∂At

∂p− ∂At

∂x1

]T(27)

and the time scales in the physical and canonical forms arealso equivalent.

Finally, choose the Hamiltonian function,

ψ(x, p, t) = At[x, y(x, p, t), t]. (28)

By substituting (28) into (27), Hamilton equations in (21) aresatisfied as follows,

∂ψ

∂x= −dp

dt,

∂ψ

∂p=dx

dt(29)

and are equivalent to a two-dimensional time-varying systemin canonical space X = F(X ), with Hamiltonian function ψ.Furthermore, this Hamiltonian formulation is unconditionallyvalid for any system governed by (1) and (15), and is mathe-matically equivalent to Lagrangian fluid transport for unsteadyflow in two dimensions, proving the conservative property of(15) [45].

VI. NUMERICAL SOLUTION OF DOC PROBLEM

The necessary conditions for optimality conditions for DOCproblems in the form of (14)-(18) were recently derived in[19]. These optimality conditions amount to a set of parabolicPDEs without a known analytical solution. This paper presentsa direct DOC solution method that parameterizes the agentPDF by a finite Gaussian mixture model, and discretizes thecontinuous DOC problem about a finite set of collocationpoints to obtain a nonlinear program (NLP) that is solvednumerically using sequential quadratic programming (SQP).Based on the conservation analysis results in Section V, thediscretized DOC problem can be obtained using an efficientFV discretization scheme that has a computational complexityfar reduced compared to classical optimal control.

It is assumed that the optimal sensor PDF can be ap-proximated by a finite Gaussian mixture model (GMM) [47]

obtained from the linear superposition of L time-varyingcomponents with density,

fj(x, t) =1

(2π)n/2|Σj |1/2e[−(1/2)(x−µj)

TΣ1j (x−µj)] (30)

where j = 1, . . . , L, | · | denotes the matrix determinant, (·)−1

denotes the matrix inverse, µj ∈ Rn is a time-varying meanvector, Σj ∈ Rn×n is a time-varying covariance matrix, and Lis an integer chosen by the user. Thus, at any time t ∈ (T0, Tf ]the optimal agent distribution can be represented as,

℘(x, t) =L∑

j=1

wj(t)fj(x, t) (31)

where the mixing proportions or weights, w1, ..., wL, obey 0 ≤wj ≤ 1 ∀j and

∑Lj=1 wj = 1 at all times [34].

An approximately optimal agent distribution ℘∗ can beobtained by determining the optimal trajectories of the mixturemodel parameters wj , µj , and Σj , for j = 1, . . . , L. Let ∆tdenote a constant discretization time interval, and k denote adiscrete time index, such that ∆t = (Tf − T0)/K, and thustk = k∆t, for k = 0, . . . ,K . Assume that the microscopiccontrol inputs u are piecewise-constant during every timeinterval ∆t, and that,

℘k , ℘(x, tk) ≈L∑

j=1

wj(tk)fj(x, tk) (32)

≡L∑

j=1

wjk1

(2π)n/2|Σjk|1/2e[−(1/2)(x−µjk)

TΣ1jk (x−µjk)]

represents the agent distribution at tk. Then, the weights wjk

and the elements of µjk and Σjk, for all j and k, are organizedinto a vector ζ of parameters to be determined such that theDOC cost function (14) is minimized, the DOC constraints(15)-(18) are satisfied, and such that the component densities,f1, . . ., fL, are nonnegative and obey the normalizationcondition for all k.

Since ℘ is a conserved quantity of a Hamiltonian system(Section V), the evolution equation (15) can be discretizedusing a conservative FV discretization algorithm that doesnot suffer from dissipative error when using a coarse-grainedstate discretization [42]. The FV algorithm adopted in thispaper partitions the state space X into finite volumes definedby a constant discretization interval ∆x ∈ Rn that are eachcentered about a collocation point xl ∈ X ⊂ Rn, l = 1, ..., X .

Now, let ℘l,k and ul,k denote FV approximations of℘(xl, tk) and c[℘(xl, tk)], respectively. Then, the FV approx-imation of the evolution equation (15) is obtained by applyingthe divergence theorem to (15) in every finite volume, suchthat ℘k+1 = ℘k +∆tρk, where,

ρk , −∫S

[℘k f(℘l,k,ul,k, tk)] · n dS (33)

and S and n denote the finite volume boundary and unitnormal, respectively. To ensure numerical stability, the dis-cretization intervals ∆t and ∆x are chosen to satisfy theCourant-Friedrichs-Lewy condition [42].

Then, letting ∆x(j) denote the jth element of ∆x, the dis-

Page 6: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

6

cretized DOC problem can be written as the finite-dimensionalNLP,

min JD =n∑

j=1

∆x(j)

X∑l=1

[φl,K +∆t

K∑k=1

L (℘l,k,ul,k, tk)]

sbj to ℘k+1 − ℘k −∆tρk = 0, k = 1, . . . ,Kn∑

j=1

∆x(j)

X∑l=1

℘l,k − 1 = 0, k = 1, . . . ,K (34)

℘l,0 = g0(xl), ∀xl ∈ X℘l,k = 0, ∀xl ∈ ∂X , k = 1, . . . ,K

℘k ≤∆x(j)

Nπr2k = 1, . . . ,K

where ℘l,0 is the initial distribution at xl, φl,K , φ(℘l,K) isthe terminal constraint. In addition, the inequality results fromthe geometric constraint in (20).

From (32) it can be seen that ℘l,k and ul,k are functionssolely of the mixture model parameters ζ, and thus theelements of ζ constitute the NLP variables. Also, since ℘is modeled by a Gaussian mixture, the state constraint (18)is always satisfied and needs not be included in the NLP.The solution ζ∗ of the NLP in (34) is obtained using anSQP algorithm that solves the Karush-Kuhn-Tucker (KKT)optimality conditions by representing (34) as a sequence ofunconstrained quadratic programming (QP) subproblems withobjective function JS(ζ) = JD(ζ) +

∑ λξ(ζ), where ξ

denotes the th constraint in (34), and λ denotes a vector ofmultipliers of proper dimensions.

At every major iteration ` of the SQP algorithm, the Hessianmatrix H = ∂JS/∂ζ is approximated using the Broyden-Fletcher-Goldfarb-Shanno (BFGS) rule,

H`+1 = H` +q`q

T`

qT` ∆ζ`

− HT` ∆ζT

` ∆ζ`H`

∆ζT` H`∆ζ`

(35)

where ∆ζ` = ζ` − ζ`−1, and q` is the change in thegradient ∇JS = ∂JS/∂ζ at the `th iteration [48]. The Hessianapproximation (35) is then used to generate a QP subproblem,

min h(d`) = (1/2) dT` H`d` +∇JT

S d`

sbj to ∇ξT d` + ξ = 0, ∀(36)

in the search direction d`. The optimal search direction d∗` is

computed from the above QP using an off-the-shelf QP solver[49], such that ζ`+1 = ζ` + α`d

∗` .

The step-length α` is determined by an approximate linesearch in the direction d∗

` , aimed at producing a sufficientdecrease in the merit function,

Ψ(ζ`) = J(ζ`) +∑

rT`, ξ(ζ`) (37)

based on the Armijo condition, and a penalty parameter r`,defined in [48]. The algorithm terminates when the KKTconditions are satisfied within a desired tolerance.

The NLP solution ζ∗ provides the optimal agent PDF ℘∗

according to (32), and ℘∗ can be used to obtain a microscopiccontrol law u∗(tk) = c[℘∗(x, tk)] for each sensor using thepotential field approach presented in [19]. Other PDF-based

control approaches, such as Voronoi diagrams [27], [28], orvirtual boundary methods [29] can potentially also be usedsince the optimal and reachable PDF is now known from ℘∗.

VII. COMPUTATIONAL COMPLEXITY ANALYSIS

The computational complexity of the direct DOC methodpresented in the previous section is compared to that of a directmethod for classical optimal control (OC) taken from [50]. Thedirect method in [50] obtains an NLP representation of theclassical optimal control problem by discretizing N -coupledODEs in the form (1) and the corresponding integral costfunction about a finite set of collocation points. Subsequently,the NLP solution can be obtained using an SQP algorithm withthe computational complexity shown in Table I (Classical OC).This classical direct method was also used in [17] to optimizethe track coverage of a mobile sensor network for N < 100.

Similarly to classical OC, the computational complexityof the SQP algorithm for DOC, described in Section VI,can be analyzed by determining the computation requiredby three most expensive steps, namely, the Hessian update(35), the solution of the QP subproblem (36), and the line-search minimization of the merit function (37). As shownin Table I, the solution of the QP subproblem, which iscarried out by a QR decomposition of the active constraintsusing Householder Triangularization [48], is the dominantcomputation in determining ℘∗.

It can be easily shown (Section VIII) that the computationrequired to obtain the microscopic control law from ℘∗ growslinearly with N . Thus, the computation required by the DOCdirect method exhibits cubic growth only with respect to K,and quadratic growth with respect to z. On the other hand,the computation required by the classical OC direct methodexhibits cubic growth with respect to K and N , and becomesprohibitive for N >> 1. Thus, for sensor networks withX << nN and z << mN , the DOC approach can bringabout considerable computational savings.

TABLE ICOMPUTATIONAL COMPLEXITY OF SQP SOLUTION

DOC Classical OCHessian update O(zXK2) O(nmN2K2)QP subproblem O(z2XK3) O(nm2N3K3)

Line search O(XK) O(nNK)

VIII. SIMULATION RESULTS

The effectiveness of the DOC approach presented in theprevious sections is demonstrated on a network of N = 250omnidirectional sensors that are each installed on a vehiclewith nonlinear unicycle kinematics,

x = v cos θ y = v sin θ θ = ω (38)

and deployed in an obstacle-populated workspace A =[0, L] × [0, L] shown in Fig. 2(b), with L = 16 km, overa time interval (T0, Tf ], with T0 = 0 and Tf = 15 hr.The sensor configuration, q = [x y θ]T , consists of the x, y-coordinates, and heading angle θ. The sensor control vector is

Page 7: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

7

u = [v ω]T , where v is the linear velocity, and ω is the angularvelocity. The sensors are assumed to have constant linearvelocities of v = 0.5 km/hr, and maximum angular velocitiesof ωmax = 0.52 rad/s, such that ω ∈ [−ωmax, + ωmax]. Itis assumed that the sensors are deployed in A with an initialdistribution ℘0 (Fig. 2(a)) and, thus, at t = T0 they are locatedat a set of initial positions sampled from ℘0. The number ofindependent elementary detections required to declare a targettrack detection is chosen to be k = 3, and the sensor effectiverange is r = 0.2 km.

The PDF of the initial target position (j = 0) is plotted inFig. 2(b), and is modeled by the Gaussian mixture,

f(xT0)=

3∑`=1

w`

(2π)n/2 det(Σ`)1/2e[−(1/2)(xT0

−µ`)TΣ1

` (xT0−µ`)]

(39)where µ1 = [0.5 7.5]T km, µ2 = [0.75 7]T km, and µ3 =[1.5 8]T km, and Σ1 = 0.1 I2, Σ2 = 0.1 I2, and Σ3 =.3 I2. The mixing proportions are w1 = 0.2, w2 = 0.2, andw3 = 0.6. The Markov model PDFs are shown in Table II,and the evolution of the target PDF over time obtained bynumerical integration is plotted in Fig. 3. The cost functionweights are chosen to be ws = 1, wr = 0.02, and we =0.1, based the relative importance of the sensing, obstacle-avoidance, and energy objectives, respectively.

(a)

(b)

Fig. 2. Initial sensor distribution in (a) and PDF of initial target distributionin (b) in an ROI with three obstacles.

(a)

(b)

(c)

Fig. 3. Evolution of target PDF at three instants in time.

The optimal time-varying PDF ℘∗ is obtained using thedirect DOC method presented in Section VI, where the chosennumber of mixture components is z = 9, the state space isdiscretized into X = 900 collocation points, for ∆t = 1 hr,and K = 15. Given ℘∗ and the estimated sensor PDF ℘, anattractive potential,

U(x, tk) ,1

2[℘(x, tk + δt)− ℘∗(x, tk + δt)]2 (40)

can be used to generate virtual forces that pull the sensorstoward ℘∗, for a small time increment δt [19]. At any time

Page 8: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

8

TABLE IIMARKOV MOTION MODEL PROBABILITY DENSITY FUNCTIONS (PDFS)

Sub-interval, (tj , tj+1] (hr) Heading PDF, fΘj(θ) Velocity PDF, fVj

(v)

j = 1 : (0, 5] (hr) N (µ, σ), µ = 2π/9 , σ = π/24 U(V), V = [.9, .925]

j = 2 : (5, 10] (hr) N (µ, σ), µ = −π/6 , σ = π/24 N (µ, σ), µ = .8 , σ = 0.025

j = 3 : (10, 15] (hr) Mult2(wi;µi, σi), w1 = 0.5 , µ1 = −π/3 , U(V), V = [1.2, 1.25]

σ1 = π/32, w2 = 0.5, µ2 = π/3 , σ2 = π/32

tk, the sensor PDF can be estimated efficiently from measure-ments of the microscopic sensor state using Kernel densityestimation [20]. Then, the microscopic feedback control law,

u∗(tk) = [v Q(θ, φ)]T (41)

can be shown to minimize (40) and provide closed-loopstability, provided φ = −∇U(x, tk), and,

Q(·) , a(θ)− a[Θ(φ)]sgna[Θ(φ)]− a(θ)

represents the minimum differential between the actual head-ing angle and the desired heading angle, where sgn(·) is thesign function, and a(·) is an angle wrapping function [40].

The optimal sensor PDF and microscopic sensor state andFOVs obtained by DOC are plotted in Fig. 4, at three samplemoments in time. The probabilities of detection of these fourmethods are presented in Fig. 9. From these simulations itcan be seen that the sensors are maximizing the probabilityof detection by anticipating the target motion forecast, whilealso avoiding obstacles and minimizing energy consumption.As shown in Fig. 5, ℘∗ can be used to generate control lawswith a cost linear in the number of sensors, N .

The performance of the DOC method is compared tofour existing sensor network deployment strategies knownas stochastic gradient, uniform, grid, and random strategies.Uniform, grid, and random strategies are static deploymentsin which N sensor positions are obtained using finite-mixturesampling [34]. The uniform deployment is obtained by sam-pling a uniform distribution over the obstacle-free space inA. The grid deployment is obtained by sampling a Gaussianmixture with z = 11 components centered on a grid, andthe random deployment samples a Gaussian mixture withz = 15 components randomly centered in A. In these staticdeployment strategies, collisions can be avoided by removingcomponents that overlap obstacles, and by requiring sampledpositions to be at a desired minimum distance from the nearestobstacle, as shown by the deployment examples in Fig. 6.

The stochastic gradient method presented in [35] is alsosimulated here for comparison. This method obtains the con-trol law for each sensor from the gradient of a function of thesensor initial and goal state in A. Uncertainties in the statemeasurements or environmental dynamics result in the controllaw that is obtained from the stochastic gradient descent of anappropriately chosen function. For the example in Figs. 2(b),the initial sensor states are sampled from ℘0, and the goalstates are sampled from a time-invariant goal sensor PDF that

minimizes the cost function (14) at Tf , and is plotted in Fig.7. By this approach, each sensor seeks to move toward theclosest goal state not occupied by another sensor, and avoidsobstacles by means of a repulsive potential term, denotedby Urep. Then, a feedback control law for sensors with theunicycle kinematics in (38) can be obtained in the form (41),letting U = wa‖x∗ − x0‖ + wbUrep, where x∗ is the goalstate, x0 is the initial state, and wa = 1 and wb = 2.5 areweighting constants. The results obtained by the stochasticgradient method are plotted in Fig. 8 at three sample momentsin time.

For each deployment strategy, the sensor performance isassessed by evaluating and averaging the actual number oftarget track detections obtained by twenty simulated sensornetworks. The cost function (14) is also evaluated by estimat-ing the sensor PDF from the microscopic sensor states usingkernel density estimation with a standard Gaussian kernel atevery time step in (T0, Tf ]. The performance comparisonresults, summarized in Fig. 5, show that the DOC methodsignificantly outperforms all other strategies by providing aprobability of detection that is up to three times as large asthe peak performance by other methods. These results arerepresentative of a number of simulations involving differentsensor initial conditions and different target PDFs.

IX. CONCLUSION

This paper presents a DOC approach for controlling anetwork of mobile omnidirectional sensors deployed to co-operatively track and detect moving targets in a region ofinterest. Several authors have shown that the performance ofcooperative multiagent networks, such as a sensor networks,can in many cases be represented as a function of the agentPDF. Existing approaches, however, assume that the optimalor goal PDF is known a priori. This paper shows that theDOC approach can be used to optimize a time-varying agentPDF subject to the agent dynamic or kinematic equations. Thepaper also shows that since the closed-loop DOC problem hasa Hamiltonian structure, an efficient direct method of solutioncan be obtained using a finite volume discretization schemethat has a computational complexity far reduced comparedto that of classical OC. The numerical simulation resultsshow that the direct DOC method presented in this paper isapplicable to networks with hundreds of sensors, and, as aresult, the network performance can be significantly increased

Page 9: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

9

(a)

(b)

(c)

Fig. 4. Evolution of optimal sensor PDF, ℘∗, microscopic state (red dots),and FOVs (red circles) at three instants in time

compared to existing stochastic gradient, uniform, grid, andrandom deployment strategies.

ACKNOWLEDGMENT

This research was funded by the ONR Code 321.

REFERENCES

[1] S. Martinez and F. Bullo, “Optimal sensor placement and motioncoordination for target tracking,” Automatica, vol. 42, pp. 661–668,2006.

Run

tim

e(s

ec)

N

Fig. 5. Time required to compute the microscopic control law in (41) as afunction of the number of sensors.

[2] H. Choset, “Coverage for robotics: A survey of recent results,” Annalsof Mathematics and Artificial Intelligence, vol. 31, no. 1-4, pp. 113–126,2001.

[3] T. Clouqueur, V. Phipatanasuphorn, P. Ramanathan, and K. Saluja,“Sensor deployment for detection of targets traversing a region,” MobileNetworks and Applications, vol. 8, pp. 453–461, August 2003.

[4] K. Chakrabarty, S. S. Iyengar, H. Qi, and E. Cho, “Grid coverage forsurveillance and target location in distributed sensor networks,” IEEETransactions on Computers, vol. 51, no. 12, pp. 1448–1453, 2002.

[5] K. C. Baumgartner, S. Ferrari, and T. Wettergren, “Robust deployment ofdynamic sensor networks for cooperative track detection,” IEEE Sensors,vol. 9, no. 9, pp. 1029–1048, 2009.

[6] K. C. Baumgartner and S. Ferrari, “A geometric transversal approachto analyzing track coverage in sensor networks,” IEEE Transactions onComputers, vol. 57, no. 8, pp. 1113–1128, 2008.

[7] K. A. C. Baumgartner and S. Ferrari, “Optimal placement of a movingsensor network for track coverage,” in Proceedings of the 2007 AmericanControls Conference, New York, NY, July 2007, pp. 4040–4046.

[8] S. Ferrari, R. Fierro, and D. Tolic, “A geometric optimization approachto tracking maneuvering targets using a heterogeneous mobile sensornetwork,” Proc. IEEE Conference on Decision and Control, Shanghai,China, 2009.

[9] K. A. C. Baumgartner, S. Ferrari, and A. V. Rao, “Optimal control of amobile sensor network for cooperative target detection,” IEEE Journalof Oceanic Engineering, vol. 34, no. 4, pp. 678–697, 2009.

[10] B. Bernard and S. Ferrari, “A geometric transversals approach to trackcoverage of maneuvering targets,” Proc. IEEE Conference on Decisionand Control, Atlanta, GA, pp. 1243–1249, 2010.

[11] H. Wei and S. Ferrari, “A geometric transversals approach to analyzingthe probability of track detection for maneuvering targets,” Computers,IEEE Transactions on, vol. 63, no. 11, pp. 2633–2646, 2014.

[12] S. Ferrari and G. Daugherty, “A q-learning approach to automatedunmanned air vehicle demining,” The Journal of Defense Modeling andSimulation, vol. 9, pp. 83–92, 2011.

[13] G. Zhang and S. F. M. Qian, “Information roadmap method for roboticsensor path planning,” Journal of Intelligent and Robotic Systems,vol. 56, pp. 69–98, 2009.

[14] S. Ferrari, R. Fierro, B. Perteet, C. Cai, and K. Baumgartner, “Ageometric optimization approach to detecting and intercepting dynamictargets using a mobile sensor network,” SIAM Journal on Control andOptimization, vol. 48, no. 1, pp. 292–320, 2009.

[15] C. Cai and S. Ferrari, “Information-driven sensor path planning byapproximate cell decomposition,” IEEE Transactions on Systems, Man,and Cybernetics - Part B, vol. 39, no. 2, 2009.

[16] ——, “Comparison of information-theoretic objective functions for deci-sion support in sensor systems,” in Proc. American Control Conference,New York, NY, 2007, pp. 63–133.

[17] K. C. Baumgartner, S. Ferrari, and A. Rao, “Optimal control of an un-derwater sensor network for cooperative target tracking,” IEEE Journalof Oceanic Engineering, vol. 34, no. 4, pp. 678–697, 2009.

[18] K. Rudd, G. Foderaro, and S. Ferrari, “A generalized reduced gradientmethod for the optimal control of multiscale dynamical systems,” in

Page 10: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

10

(a)

(b)

(c)

Fig. 6. Grid (a), random (b), and uniform (c) sensor deployments (red dots)and corresponding FOVs (red circles).

Fig. 7. Goal sensor PDF for stochastic gradient method.

Proceedings of the IEEE Conference on Decision and Control, Florence,Italy, 2013.

[19] G. Foderaro, S. Ferrari, and T. Wettergren, “Distributed optimal controlfor multi-agent trajectory optimization,” Automatica, vol. 50, pp. 149–154, 2014.

[20] G. Foderaro, S. Ferrari, and M. Zavlanos, “A decentralized kernel densityestimation approach to distributed robot path planning,” in Proceedingsof the Neural Information Processing Systems Conference, Lake Tahoe,NV, 2012.

[21] G. Foderaro and S. Ferrari, “Necessary conditions for optimality fora distributed optimal control problem,” Proc. IEEE Conference onDecision and Control, Atlanta, GA, pp. 4831–4838, 2010.

[22] S. Ferrari, G. Zhang, and T. Wettergren, “Probabilistic track coveragein cooperative sensor networks,” IEEE Transactions on Systems, Man,and Cybernetics-Part B, vol. 40, no. 6, 2010.

[23] S. Berman, A. Halasz, M. Hsieh, and V. Kumar, “Navigation-basedoptimization of stochastic strategies for allocating a robot swarm amongmultiple sites,” Proc. Conference on Decision and Control, pp. 4376–4381, 2008.

[24] J. Oyekan and H. Hu, “Ant robotic swarm for visualizing invisiblehazardous substances,” Robotics, vol. 2, pp. 1–18, 2013.

[25] S. Berman, A. Halasz, M. Hsieh, and V. Kumar, “Optimized stochasticpolicies for task allocation in swarms of robots,” IEEE Transactions onRobotics, vol. 25, no. 4, pp. 927–937, 2009.

[26] T. Wettergren and R. Costa, “Optimal placement of distributed sensorsagainst moving targets,” ACM Transactions on Sensor Networks, vol. 5,no. 3, 2009.

[27] J. Cortes, C. Martinez, T. Karatas, and F. Bullo, “Coverage controlfor mobile sensing networks,” IEEE Transactions on Robotics andAutomation, vol. 20, no. 2, pp. 243–255, 2004.

[28] S. Martinez, J. Cortes, and F. Bullo, “Motion coordination with dis-tributed information,” IEEE Control Systems Magazine, vol. 27, no. 4,pp. 75–88, 2007.

[29] N. Ayanian and V. Kumar, “Abstractions and controllers for groupsof robots in environments with obstacles,” Proc. IEEE InternationalConference on Robotics and Automation, Anchorage, Alaska, 2010.

[30] H. Wei and S. Ferrari, “A geometric transversals approach to sensormotion planning for tracking maneuvering targets,” Automatic Control,IEEE Transactions on, vol. 60, no. 10, pp. 2773–2778, Oct 2015.

[31] T. A. Wettergren, “Performance of search via track-before-detect fordistributed sensor networks,” IEEE Transactions on Aerospace andElectronic Systems, vol. 44, no. 1, pp. 314–325, January 2008.

[32] T. A. Wettergren, R. L. Streit, and J. R. Short, “Tracking with distributedsets of proximity sensors using geometric invariants,” IEEE Transactionson Aerospace and Electronic Systems, vol. 40, no. 4, pp. 1366–1374,October 2004.

[33] R. Costa and T. A. Wettergren, “Assessing design tradeoffs in deployingundersea distributed sensor networks,” in Proceedings of OCEANS 2007,Vancouver, BC, September 2007, pp. 1–5.

[34] G. McLachlan, Finite Mixture Models. New York, NY: Wiley Inter-science, 2000.

[35] J. L. Ny and G. J. Pappas, “Sensor-based robot deployment algorithms,”IEEE Conference on Decision and Control, pp. 5486–5492, 2010.

Page 11: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

11

(a)

(b)

(c)

Fig. 8. Stochastic gradient sensor deployment (blue dots), correspondingFOVs (blue circles), and goal states (red dots)

[36] S. B. Pope, Turbulent Flows. New York, NY: Cambridge UniversityPress, 2000.

[37] Y. Bar-Shalom, X. Li, and T. Kirubarajan, Estimation with Applicationsto Tracking and Navigation: Algorithms and Software for InformationExtraction. New York, NY: J. Wiley and Sons, 2001.

[38] Y. Bar-Shalom and X.-R. Li, Estimation and tracking: principles,techniques, and software. Artech House Norwood, 1993, vol. 393.

[39] Y. Bar-Shalom and E. W.D. Blair, Multitarget-Multisensor Tracking:Applications and Advances, Vol. III. Boston, MA: Artech House, 2000.

[40] J. C. Latombe, Robot Motion Planning. Kluwer Academic Publishers,1991.

[41] J. P. Boyd, Chebyshev and Fourier Spectral Methods, II Ed. New York,NY: Dover, 2001.

Fig. 9. Comparison of probability of detection

[42] J. C. Tannehille, D. A. Anderson, and R. H. Pletcher, ComputationalFluid Mechanics and Heat Transfer. Taylor and Francis, 1997.

[43] I. M. Gelfand and S. V. Fomin, Calculus of Variations. EnglewoodCliffs, NJ: Prentice-Hall, 1963.

[44] D. E. Kirk, Optimal Control Theory: An Introduction. EnglewoodCliffs, NJ: Prentice-Hall, 1970.

[45] M. F. M. Speetjens, “Topology of advective-diffusive scalar transport inlaminar flows,” Physical Review E, vol. 77, no. 2, 2008.

[46] K. Bajer, “Hamiltonian formulation of the equations of streamlines inthree-dimensional steady flows,” Chaos, Solitons and Fractals, vol. 4,no. 6, pp. 895–911, 1994.

[47] S. Ferrari, G. Foderaro, P. Zhu, and T. A. Wettergren, “Distributedoptimal control of multiscale dynamical systems: A tutorial,” IEEEControl Systems Magazine, vol. 36, no. 2, pp. 102–116, 2016.

[48] M. J. D. Powell, “A fast algorithm for nonlinearly constrained optimiza-tion calculations,” Numerical Analysis, vol. 630, 1978.

[49] Mathworks, Matlab Optimization Toolbox. [Online]. Avaliable:http://www.mathworks.com, 2004, function: fmincon.

[50] J. Betts, “Survey of numerical methods for trajectory optimization,”Journal of Guidance, Control, and Dynamics, vol. 21, no. 2, pp. 193–207, 1998.

Greg Foderaro received the B.S. degree in Mechan-ical Engineering from Clemson University in 2009and the Ph.D. degree in Mechanical Engineering andMaterials Science from Duke University in 2013. Heis currently a staff engineer at Applied Research As-sociates, Inc. His research interests are in underwatersensor networks, robot path planning, multiscale dy-namical systems, pursuit-evasion games, and spikingneural networks. He is a member of the IEEE.

Pingping Zhu received the B.S. degree in electron-ics and information engineering and the M.S. degreein the institute for pattern recognition and artificialintelligence from Huazhong University of Scienceand Technology, and M.S. and Ph.D. degrees inelectrical and computer engineering from Universityof Florida. He is currently a research associate at thedepartment of mechanical and aerospace engineeringin Cornell University. Prior to that, he was a post-doctoral associate at the department of mechanicalengineering and material science in Duke university.

His research interests include approximate dynamic programming, optimalcontrol of mobile sensor networks, signal processing, machine learning, andneural networks. He is a member of the IEEE.

Page 12: Distributed Optimal Control of Sensor Networks for Dynamic ...lisc.mae.cornell.edu/LISCpapers/TCNSdraft_FoderaroDOCforTracking… · Index Terms Optimal Control, Distributed Control,

2325-5870 (c) 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TCNS.2016.2583070, IEEETransactions on Control of Network Systems

12

Hongchuan Wei Hongchuan earned his B.S. degreein Vehicle Engineering from Tsinghua University,China in 2010 and is pursuing a Ph.D. degree atDuke University. His research interest includes sen-sor network deployment, which consists of optimallyplacing a set of sensors in a region of interest (ROI)such that the probability of detection of targets trav-eling through the ROI is maximized. The probabilityof detection is a function of the target movementand sensors field of view. He is also interested inutilizing information theoretic functions to acquire

optimal sensor controls in order to learn the behaviors of targets described bynon-parametric Bayesian models.

Thomas Wettergren received the BS in Electri-cal Engineering and the PhD in Applied Mathe-matics, both from Rensselaer Polytechnic Institute.He joined the Naval Undersea Warfare Center inNewport in 1995, where he has served as a researchscientist in the torpedo systems, sonar systems, andundersea combat systems departments. He currentlyserves as the US Navy Senior Technologist (ST) forOperational and Information Science, with a concur-rent title as a Senior Research Scientist at the Center.He also is an Adjunct Professor of Mechanical

Engineering at Pennsylvania State University. His personal research interestsare in planning and control of distributed systems, applied optimization, multi-agent systems and search theory. Dr. Wettergren is a senior member of theIEEE and a member of the Society for Industrial and Applied Mathematics(SIAM).

Silvia Ferrari is a Professor of MAE at CornellUniversity. Prior to that, she was Professor of En-gineering and Computer Science, and Founder andDirector of the NSF Integrative Graduate Educationand Research Traineeship (IGERT) and Fellowshipprogram on Wireless Intelligent Sensor Networks(WISeNet) at Duke University. She is the Direc-tor of the Laboratory for Intelligent Systems andControls (LISC), and her principal research interestsinclude robust adaptive control of aircraft, learningand approximate dynamic programming, and opti-

mal control of mobile sensor networks. She received the B.S. degree fromEmbry-Riddle Aeronautical University and the M.A. and Ph.D. degrees fromPrinceton University. She is a senior member of the IEEE, and a member ofASME, SPIE, and AIAA. She is the recipient of the ONR young investigatoraward (2004), the NSF CAREER award (2005), and the Presidential EarlyCareer Award for Scientists and Engineers (PECASE) award (2006).