autonomous surface vehicle multistep look-ahead measurement...

14
Autonomous Surface Vehicle Multistep Look-Ahead Measurement Location Planning for Optimal Localization of Underwater Acoustic Transponders JESSE R. GARCIA , Student Member, IEEE JAY A. FARRELL , Fellow, IEEE ZAHER M. KASSAS , Senior Member, IEEE University of California, Riverside, CA, USA MICHAEL T. OUIMET SPAWAR Systems CenterPacific, San Diego, CA, USA An underwater vehicle may utilize underwater transponders (UTs) for navigation in the absence of global navigation satellite sys- tem signals. However, the position of each UT must be known by the underwater vehicle. The problem of an autonomous surface vehicle (ASV) optimally planning measurement locations to localize a set of arbitrarily predeployed acoustic UTs is considered. The ASV is as- sumed to make noisy range measurements to the UTs. A maximum a posteriori estimator is derived to localize the UTs. In addition, a multi- step look-ahead (MSLA) ASV optimal measurement location planning (OMLP) strategy is developed. This planning strategy prescribes fu- ture multistep measurement locations. A physical interpretation of the proposed planner in the single-step, single transponder case is pro- vided. Simulation results are presented demonstrating the tradeoff between expected localization performance and computational time associated with various look-ahead horizons and travel distances. Ex- perimental results are given illustrating the proposed MSLA OMLP Manuscript received August 3, 2018; revised December 26, 2018; released for publication December 28, 2018. Date of publication April 4, 2019; date of current version December 5, 2019. DOI. No. 10.1109/TAES.2019.2909253 Refereeing of this contribution was handled by Z. Davis. This work was supported in part by the Office of Naval Research under Grant N00014-16-1-2768. Authors’ addresses: J. R. Garcia, J. A. Farrell, and Z. M. Kassas are with the Department of Electrical and Computer Engineering, University of California, Riverside, Riverside, CA 92521 USA, E-mail: (jegarcia@ece. ucr.edu; [email protected]; [email protected]); M. T. Ouimet is with the SPAWAR Systems Center, Pacific, San Diego, CA 92152 USA, E-mail: ([email protected]). (Corresponding author: Zaher M. Kassas.) 0018-9251 C 2019 IEEE strategy’s performance in environments containing one and two UTs. The proposed OMLP strategy is able to localize UTs to within 4 m of their true locations. Additionally, increasing the planning horizon is demonstrated to yield better UT localization at the cost of increased computational burden. I. INTRODUCTION Underwater vehicles require knowledge of their po- sition to perform their mission [1]. Global navigation satellite system (GNSS) signals become severely attenuated underwater, rendering them unusable at depths below a few feet. Resurfacing to re-establish GNSS connection [2], [3] is not desirable in situations where stealth and covertness are required. The underwater vehicle may instead rely on signals from underwater transponders (UTs) with known locations to determine its position estimate. This paper fo- cuses on localizing a set of UTs arbitrarily predeployed at unknown locations. Once the UT locations are known, they could serve as reference beacons for underwater navigation. This paper considers the following problem. An au- tonomous surface vehicle (ASV) makes acoustic range measurements to estimate the positions (i.e., localize) sev- eral predeployed UTs that are rigidly attached to the sea floor. Where should the ASV place itself to use range measurements to optimally localize the UTs? This problem has similarities to optimal sensor deploy- ment, to which many optimization criteria have been de- veloped [4]–[8]. Among the most common optimization criteria is the D-optimality criterion, namely maximization of the logarithm of the determinant of the Fisher informa- tion matrix [9]. The D-optimality criterion yields the max- imum reduction in target location uncertainty as measured by the volume of the uncertainty ellipsoid [10], [11]. An alternative criterion to D-optimality was proposed in [12], referred to as maximum innovative logarithm-determinant (MILD), which seeks to maximize the volume of the inno- vation matrix. MILD was shown to be both computationally efficient and identical to the D-optimality criterion under linear Gaussian assumptions and was demonstrated to yield comparable performance with nonlinear pseudorange mea- surements. Another computationally efficient criterion with a geometric interpretation that approximately minimizes the dilution of precision was proposed in [13]. This criterion was shown to yield a family of convex programs that can be solved in parallel. A collaborative sensor placement strategy was developed in [14], wherein a network of coordinated ASVs attempt to optimally place themselves to localize a single UT. The dual of this problem, wherein landmarks are optimally placed to aid in vehicle positioning, was explored in [15]. While classical sensor placement problems consider planar environments, a three-dimensional (3-D) sensor placement strategy is needed for certain underwater and aerial applications. In underwater applications, an addi- tional complexity arises from the fact that, once submerged, an autonomous underwater vehicle (AUV) is deprived of 2836 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Upload: others

Post on 07-Aug-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

Autonomous Surface VehicleMultistep Look-AheadMeasurement LocationPlanning for OptimalLocalization of UnderwaterAcoustic Transponders

JESSE R. GARCIA , Student Member, IEEEJAY A. FARRELL , Fellow, IEEEZAHER M. KASSAS , Senior Member, IEEEUniversity of California, Riverside, CA, USA

MICHAEL T. OUIMETSPAWAR Systems Center Pacific, San Diego, CA, USA

An underwater vehicle may utilize underwater transponders(UTs) for navigation in the absence of global navigation satellite sys-tem signals. However, the position of each UT must be known by theunderwater vehicle. The problem of an autonomous surface vehicle(ASV) optimally planning measurement locations to localize a set ofarbitrarily predeployed acoustic UTs is considered. The ASV is as-sumed to make noisy range measurements to the UTs. A maximum aposteriori estimator is derived to localize the UTs. In addition, a multi-step look-ahead (MSLA) ASV optimal measurement location planning(OMLP) strategy is developed. This planning strategy prescribes fu-ture multistep measurement locations. A physical interpretation of theproposed planner in the single-step, single transponder case is pro-vided. Simulation results are presented demonstrating the tradeoffbetween expected localization performance and computational timeassociated with various look-ahead horizons and travel distances. Ex-perimental results are given illustrating the proposed MSLA OMLP

Manuscript received August 3, 2018; revised December 26, 2018; releasedfor publication December 28, 2018. Date of publication April 4, 2019; dateof current version December 5, 2019.

DOI. No. 10.1109/TAES.2019.2909253

Refereeing of this contribution was handled by Z. Davis.

This work was supported in part by the Office of Naval Research underGrant N00014-16-1-2768.

Authors’ addresses: J. R. Garcia, J. A. Farrell, and Z. M. Kassas are withthe Department of Electrical and Computer Engineering, University ofCalifornia, Riverside, Riverside, CA 92521 USA, E-mail: ([email protected]; [email protected]; [email protected]); M. T. Ouimet is with theSPAWAR Systems Center, Pacific, San Diego, CA 92152 USA, E-mail:([email protected]). (Corresponding author: Zaher M. Kassas.)

0018-9251 C© 2019 IEEE

strategy’s performance in environments containing one and two UTs.The proposed OMLP strategy is able to localize UTs to within 4 m oftheir true locations. Additionally, increasing the planning horizon isdemonstrated to yield better UT localization at the cost of increasedcomputational burden.

I. INTRODUCTION

Underwater vehicles require knowledge of their po-sition to perform their mission [1]. Global navigationsatellite system (GNSS) signals become severely attenuatedunderwater, rendering them unusable at depths below a fewfeet. Resurfacing to re-establish GNSS connection [2], [3]is not desirable in situations where stealth and covertnessare required. The underwater vehicle may instead rely onsignals from underwater transponders (UTs) with knownlocations to determine its position estimate. This paper fo-cuses on localizing a set of UTs arbitrarily predeployedat unknown locations. Once the UT locations are known,they could serve as reference beacons for underwaternavigation.

This paper considers the following problem. An au-tonomous surface vehicle (ASV) makes acoustic rangemeasurements to estimate the positions (i.e., localize) sev-eral predeployed UTs that are rigidly attached to the seafloor. Where should the ASV place itself to use rangemeasurements to optimally localize the UTs?

This problem has similarities to optimal sensor deploy-ment, to which many optimization criteria have been de-veloped [4]–[8]. Among the most common optimizationcriteria is the D-optimality criterion, namely maximizationof the logarithm of the determinant of the Fisher informa-tion matrix [9]. The D-optimality criterion yields the max-imum reduction in target location uncertainty as measuredby the volume of the uncertainty ellipsoid [10], [11]. Analternative criterion to D-optimality was proposed in [12],referred to as maximum innovative logarithm-determinant(MILD), which seeks to maximize the volume of the inno-vation matrix. MILD was shown to be both computationallyefficient and identical to the D-optimality criterion underlinear Gaussian assumptions and was demonstrated to yieldcomparable performance with nonlinear pseudorange mea-surements. Another computationally efficient criterion witha geometric interpretation that approximately minimizes thedilution of precision was proposed in [13]. This criterionwas shown to yield a family of convex programs that can besolved in parallel. A collaborative sensor placement strategywas developed in [14], wherein a network of coordinatedASVs attempt to optimally place themselves to localize asingle UT. The dual of this problem, wherein landmarks areoptimally placed to aid in vehicle positioning, was exploredin [15].

While classical sensor placement problems considerplanar environments, a three-dimensional (3-D) sensorplacement strategy is needed for certain underwater andaerial applications. In underwater applications, an addi-tional complexity arises from the fact that, once submerged,an autonomous underwater vehicle (AUV) is deprived of

2836 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 2: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

GNSS signals; hence, GNSS-derived positioning1 informa-tion becomes unavailable. This forces a reliance on onboardinertial sensor suites that provide positioning estimates thatdeteriorate over time in the absence of an aiding source (e.g.,GNSS or signals of opportunity [16], [17]). Underwater si-multaneous localization and mapping (SLAM) techniqueshave been developed to mitigate such deterioration [18].

Inertial sensors may be made robust to integral effectsand drift via real-time corrections. This requires sensors(e.g., acoustic, magnetic, and pressure) to periodically cor-rect for inertial navigation system (INS) drift [19]. Thestandard long baseline (LBL) acoustic approach to aidedinertial navigation uses an extended Kalman filter to fuseranges to multiple UTs with other AUV sensors [20]. Analternative approach to LBL was developed in [21], whichuses an imaging sonar. Several magnetic/geomagnetic field-aided approaches have been developed. The inversion ofthe gradient of the magnetic field vector associated with amagnetic dipole at a fixed, known location was proposedin [22]. Simulated results emphasize the improvement innavigation quality (i.e., the decrease in vehicle position er-ror) by utilizing tensor Euler deconvolution in tandem witheigenanalysis of the magnetic gradient tensor. This paperis motivated by the computational issues associated withanalytical magnetic field inversion (i.e., when the magneticfield gradient tensor is singular). Another robust, computa-tionally efficient algorithm for magnetic-aided navigationwas developed in [23]. An adaptive unscented Kalman fil-ter for magnetic navigation in a SLAM framework (e.g.,the magnetic dipole’s position is unknown) was formulatedin [24].

Environmental features may also be used to mitigatedecay of quality of position estimates associated with in-ertial navigation. For example, a sonar-based terrain-aidednavigation (TAN) framework was developed in [25] to aidan AUV. This is paired with the work in [22] to develop ahybrid TAN-magnetic aided INS in [26]. This system wasshown to yield lower AUV position error than either TANor magnetic-aided navigation methods alone. An adaptiveparticle filter approach using sonar is developed in [27]. In[28], criteria were developed to ensure observability of thenonlinear system when ranging to a single acoustic beacon,while [29] and [30] derive such criteria when measuringpseudoranges to multiple terrestrial signal transmitters.

Multistep look-ahead (MSLA) planning, also knownas model-predictive control or receding horizon control,has seen several applications in SLAM-type problems.This scheme is adopted in [31] for AUV fleet formationcontrol. In [32], MSLA trajectory planning was numer-ically evaluated for its prospective benefit in a roboticSLAM environment. A similar numerical evaluation waspresented in [29] in both radio SLAM and mapping-only

1In this paper, the term “positioning” refers to the vehicle estimating itsown states. The term “localization” will be used in reference to a vehicleestimating the states of some other entity.

scenarios of several terrestrial transmitters using pseu-dorange measurements. A hybrid receding horizon pathplanner/model predictive AUV controller for dynamic nav-igation based on environmental information was developedin [33]. An adaptive path planning framework was devel-oped in [34] to react to hazardous environments while ma-neuvering at high speed. A receding horizon planner forantisubmarine warfare AUV applications that is feasiblein real time was presented in [35]. The controller is able todrive the AUV in such a way as to minimize the localizationerror of a mobile undersea target based on measurementsfrom a bistatic sonar array.

A single-step look-ahead optimal measurement locationplanning (OMLP) strategy was developed in [36], whichspecifies the next optimal location at which the ASV shouldplace itself to make acoustic range measurements to theUTs. This paper extends [36] and makes three contribu-tions. First, a maximum a posteriori (MAP) frameworkis developed to localize UTs from noisy acoustic rangemeasurements. Second, an MSLA OMLP strategy is de-veloped that plans future D-optimal measurement locationsby specifying the ASV’s yaw rate. Third, simulation andexperimental results are presented demonstrating the supe-riority of the MSLA OMLP strategy over the single-steplook-ahead as well as random ASV motion.

The remainder of this paper is organized as follows.Section II describes the ASV’s dynamics and observa-tion model. Section III formulates the UT localization andMSLA OMLP problems for an arbitrary number of UTsin the environment. Section IV presents a MAP approachto solve 1) the UT localization problem and 2) a compu-tationally efficient approach to solve the OMLP problem.Section V presents simulation results demonstrating bothsolutions. Section VI presents experimental results showingOMLP for one and two UTs with a localization accuracy ofa few meters. Concluding remarks are given in Section VII.

II. MODEL DESCRIPTION

The following nomenclature and conventions will beused throughout this paper unless stated otherwise. Vectorswill be column and represented by lower-case, italicized,and bold characters (e.g., x). Matrices will be representedby upper-case bold characters (e.g., X).

The symbol n will denote a specific measurement epoch,while N denotes the total number of measurements madeby the ASV. The symbol M denotes the total number ofUTs in the environment, while m = 1, . . . , M serves toindex each UT. The symbol k corresponds to the number ofsteps planned by the MSLA OMLP. When k = 0, the ASVrandomly plans the next measurement location.

A. Vehicle State and Model

Let pV (n) ∈ R3 represent the 3-D position of the ASV.

The velocity of the ASV at measurement epoch n is vV (n) ∈R

3 and is represented in the vehicle’s coordinate frame. Theoptimal next location from which the ASV should make a

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2837

Page 3: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

measurement is p∗V (n + 1). This location is determined by

solving an optimization problem that is constrained by akinematic model with a maximum distance constraint.

To adopt an MSLA structure, the following model isassumed to propagate the ASV’s position pV and yaw angleφV

�V :

{pV (n + 1) = pV (n) + Rz [φV (n + 1)] vV (n)T

φV (n + 1) = φV (n) + ωV (n)T , n = 1, . . . , N

(1)

where T is the sampling period, ωV is the yaw rate, and Rz

is a rotation matrix about the z-axis of the local frame byangle φV , which has the form

Rz [φV (n)] =

⎡⎢⎣

cos(φV (n)) − sin(φV (n)) 0

sin(φV (n)) cos(φV (n)) 0

0 0 1

⎤⎥⎦. (2)

The model in (1) assumes that T is sufficiently small suchthat φV and vV over nT ≤ t < (n + 1)T are constant.

The ASV’s state is defined as xV � [ pTV , φV ]T ∈ R

4.

B. UT State and Model

It is assumed that M UTs exist in the environment atfixed, unknown positions. Let pm

T ∈ R3 be the position of

the mth UT, which propagates according to pT (n + 1) =pT (n), n = 1, 2, . . . , N .

C. Observation Model

The acoustic range observation made by the ASV atthe nth measurement epoch is related to the mth UT’sposition by

zm(n) = rm(n) + w(n), (3)

where rm(n) � ‖ pV (n) − pmT ‖2 is the true range between

the ASV and UT and ‖ · ‖2 is the Euclidean norm, andw is the measurement noise, which is modeled as w(n) ∼N (0, σ 2) and assumed to be temporally independent aswell as independent and identically distributed across all N

measurement epochs. The vector of N range measurementsto the mth UT is denoted by

zm = rm + wm ∈ RN, (4)

where rm � [rm(1), . . . , rm(N)]T and wm � [wm(1), . . . ,wm(N)]T. The distribution of the measurement vector iszm ∼N (rm, R), where R = σ 2 IN×N .

III. UT LOCALIZATION AND MSLA OMLP PROBLEMFORMULATION

This section formulates the UT localization and MSLAOMLP problems The set of all ASV locations is denotedby NPV . This set is constructed as

NPV = { pV (1), . . . , pV (N)}. (5)

The ASV positions are assumed to be known (e.g., fromGNSS measurements). The estimate of the location of the

mth UT ( pmT ) based on N range measurements is denoted

by pmT (N). The range can be predicted as

rm(n) � ‖ pmT (N) − pV (n)‖2. (6)

The vector of estimated ranges is

rm �[rm(1), . . . , rm(N)

]T ∈ RN. (7)

The Jacobian vector hm(n) � ∂rm(n)∂ pm

T

| pmT (n) is

hm(n) =[

pmT (N) − pV (n)

]T

rm(n)∈ R

1×3. (8)

The vectors in (8) can be stacked to form the matrix

Hm = [(hm(1))T, . . . , (hm(N))T]T ∈ R

N×3. (9)

PROBLEM 1 UT LOCALIZATION: Given zm and NPV ,estimate pm

T .

PROBLEM 2 MSLA OMLP: Given the set of UTestimates using N range measurements { pm

T (N)}Mm=1,determine the next k values of the yaw rate[ωV (N), . . . , ωV (N + k − 1)] that will drive the ASV tothe locations { pV (N + 1), . . . , pV (N + k)} in order to min-imize the aggregate uncertainty of the set of future estimates{ pm

T (N + k)}Mm=1, as measured by the logarithm of the de-terminant of the corresponding estimation error covariance.

IV. UT LOCALIZATION AND MSLA OMLP SOLUTION

This section presents solutions to the UT localizationand OMLP problems, which were formulated in Section III.

A. Solution to the UT Localization Problem

To reduce the linearization error on UT localization,an iterative Gauss–Newton approach is adopted [37] alongwith a MAP estimation formulation. The iterative estima-tion algorithm is initialized with an estimate and corre-sponding covariance, denoted by 0 pm

T and 0PmT , respectively,

with the assumption 0 pmT ∼ N ( pm

T ,0PmT ). In the case when

one has little information about pmT , 0Pm

T is set to be verylarge.

In what follows, the presubscript j denotes the iterationnumber. Specifically, the estimate of pm

T at the j th iterationusing N measurements is j pm

T (N). Estimated ranges andJacobians at the j th iteration, j r

m(n) and j hm(n), are com-puted by substituting j pm

T (N) into (6) and (8). Similarly,j rm and j Hm are constructed by substituting j r

m(n) andj hm(n), ∀n ≤ N , into (7) and (9).

A MAP framework for UT localization is adopted [37],[38]. The basic optimization problem is

pmT (N) � argmax

pmT

[p(zm | pm

T ) p( pmT )

]. (10)

Define the composite vectors

y �[(zm)T, (0 pm

T )T]T(11)

f ( pmT ) �

[(rm)T, ( pm

T )T]T. (12)

2838 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 4: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

The covariance matrix for y is

C = blkdiag[R, (0Pm

T )], (13)

where blkdiag(·) denotes a block-diagonal matrix. Maxi-mizing the objective function of (10) is equivalent to mini-mizing

JMAP( pmT ) � 1

2

[y − f ( pm

T )]T

C−1 [y − f ( pm

T )]. (14)

This allows the maximization problem in (10) to be rewrit-ten as

pmT (N) � argmin

pmT

JMAP( pmT ). (15)

The MAP estimate of pmT is computed by solving the batch

nonlinear least-squares optimization problem in (15) itera-tively according to the following steps. Start the iterationcounter with j = 0. Let the j th estimate of pm

T be j pmT .

The estimated range vector is j rm = rm| pmT =j pm

T. The vector

f ( pmT ) is linearized around j pm

T at each iteration, yielding

f ( pmT ) ≈ f (j pm

T ) + Vj δ pj , (16)

where

Vj �[

Hm(j pmT )

I3×3

]∈ R

(N+3)×3 (17)

δ pj � pmT −j pm

T . (18)

Define δ yj � y − f (j pmT ) to simplify notation. The MAP

objective function in (15) can be re-expressed as

JMAP(δ pj ) = (δ yj − Vj δ pj )TC−1(δ yj − Vj δ pj ). (19)

The linearized objective function (19) is minimized when

δ pj ≡ (VT

j C−1Vj

)−1VT

j C−1δ yj . (20)

The estimate at the next iteration is

j+1 pmT =j pm

T + δ pj , (21)

where j pmT is the j th estimate of the position of the mth

UT, and δ pj is defined in (20). The iterations continue until‖j+1 pm

T −j pmT ‖2 ≤ δmin, or as long as j ≤ Jmax, for user-

defined Jmax and δmin. After convergence, the covarianceis

NPmT ≡ (

VTj C−1Vj

)−1. (22)

B. Solution to the MSLA OMLP Problem

The D-optimality criterion [39, p. 387] is used to deter-mine the future k-step trajectory { pV (N + n)}kn=1 to maxi-mize the information gain. For this section, the subscript N

corresponds to the number of previous ASV locations. Thesymbol Ym

N ∈ R3×3 denotes the information matrix after N

ASV measurements to the mth UT (i.e., YmN = (NPm

T )−1).The single-step look-ahead (i.e., k = 1), commonly re-

ferred to as a greedy strategy, was formulated in [36] as

x∗ = argmaxx

J (x)

subject to g(x) ≤ dmax,(23)

where

J (x) � log det [YN+1 (x)] (24)

YN+1 � blkdiag[Y1

N+1, . . . , YMN+1

](25)

g(x) � ‖x − pV (N)‖2 (26)

x � pV (N + 1). (27)

The distance constraint dmax represents the maximum dis-tance the ASV may travel between epochs N and N + 1.

In what follows, the single-step OMLP is generalizedto the MSLA OMLP. Define

DmN+k (ωk) �

(hm(N + k)

)Thm(N + k)

where the symbol ωk � ωV (N + k − 1) is used to simplifynotation. Recall from (8) that hm(N + k) depends on ωk

because ωV (·) is integrated through (1) to determine N+kPV .The information about transponder m at epoch N + k

may be expressed as

YmN+k(ωk) � σ−2 (

HmN+k

)THm

N+k

= σ−2[(

HmN+k−1

)THm

N+k−1 + DmN+k (ωk)

]= Ym

N+k−1 + σ−2 DmN+k (ωk) . (28)

The information YmN+k may be expressed as a function

of YmN by recursively using (28)

YmN+k(ωV ) = Ym

N + σ−2k∑

n=1

DmN+n (ωn), (29)

where ωV = {ωn}kn=1.The MSLA D-optimality problem is

ωV∗ = argmax

ωV

J (ωV )

subject to �V

{g(ωV (n)) ≤ ωmax}kn=1 ,

(30)

where

J (ωV ) � log det [YN+k (ωV )] (31)

YN+k � blkdiag[Y1

N+k, . . . , YMN+k

](32)

g(ωV (n)) � |ωn|. (33)

The distance constraint in (23) is subsumed into vV in (1).The angular rate constraint g(ωV (n)) restricts each ωn to anenvelope of ± ωmax. Notice that Ym

N+k(ωV ) is determinedvia linearization about estimate pm

T (N). The first elementof ω∗

V , namely ω∗1, is applied to maneuver the ASV to

obtain xV (N + 1) [cf. (1)], while the remaining elementsω∗

2, . . . , ω∗k are discarded. Note that these elements are dis-

carded since at the next epoch, the ASV will receive newacoustic measurements which will be used to obtain a “re-fined” solution by solving the MSLA again (i.e., recede theplanning horizon). Also note that while these elements arediscarded, they impact the obtained optimal solution ω∗

1,which is computed as part of the multistep vector ω∗

V .Next, the MSLA D-optimality problem (30) is simpli-

fied by exploiting the block-diagonal structure of YN+k(ωV )

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2839

Page 5: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

to write

ω∗V = argmax

ωV

log f (ωV )

subject to �V

{g(ωV (n)) ≤ ωmax}kn=1

(34)

where f (ωV ) = ∏Mm=1det[Ym

N+k(ωV )]. Using the proper-ties of the logarithm function

log f (ωV ) =M∑

m=1

log det[Ym

N+k(ωV )]. (35)

Further simplifications may be made if σ−2 ∑kn=1 Dm

N+n

(xn) is linearly separable, i.e., if

YmN+k(ωV ) = Ym

N + d1(ωV )dT2(ωV ),

where the vectors di(ωV ) ∈ R3×1, i = 1, 2. While this de-

composition may not hold in general, it is clear from theform of (29) that this is true when k = 1. In this case,di(ωV ) = (hm(N + 1))T. When this decomposition holds,the matrix determinant properties developed in AppendixA, namely (59), could be applied to (35) to give

det[Ym

N+k(ωV )] = det

[Ym

N

].

det[1 + dT

2(ωV )(YmN )−1 d1(ωV )

].

(36)

Define

αm(ωV ) � dT2(ωV )(Ym

N )−1 d1(ωV ), (37)

which is a positive scalar. Using (36) and (37), the opti-mization function in (35) can be simplified to

log f (ωV ) =M∑

m=1

log[ [

1 + αm(ωV )]det

[Ym

N

] ]. (38)

Properties of the logarithm function allow further simplifi-cation to

log f (ωV ) =M∑

m=1

log[1 + αm(ωV )

] +M∑

m=1

log det[Ym

N

].

(39)The term

∑Mm=1 log det[Ym

N ] is constant with respect to ωV ,so it can be dropped from the optimization function.

Letting J (ωV ) �∑M

m=1 log [1 + αm(ωV )], the opti-mization problem of (34) can be written as

ω∗V = argmax

ωV

J (ωV )

subject to �V

{g(ωV (n)) ≤ ωmax}kn=1 .

(40)

Note that for M = 1, (40) is the same as

ω∗V = argmax

ωV

J ′(ωV )

subject to �V

{g(ωV (n)) ≤ ωmax}kn=1

(41)

where J ′(ωV ) � α(ωV ).

TABLE IASV Locations Used in Monte

Carlo Analysis

V. SIMULATION RESULTS

This section presents simulation results for the UT local-ization and MSLA OMLP problems. The following quanti-tative metrics will be used in this section. Define the positionerror vector of the mth UT at the nth epoch as

pmT (n) � pm

T − pmT (n) ∈ R

3. (42)

The augmented position error vector for all M UTs is

pT (n) �[(

p1T (n)

)T, . . . ,

(pM

T (n))T

]T∈ R

3M. (43)

The magnitude of the position error vector is ‖ pmT (n)‖2.

The normalized estimation error squared (NEES) at epochn is

ε(n) � pTT (n)Yn pT (n). (44)

The localization performance is evaluated using MonteCarlo tests. Let c ∈ {1, 2, . . . , C} be the index of a spe-cific Monte Carlo trial. The localization root-mean-squareestimation error (RMSEE) of the mth UT at the nth epochis defined as

pmT (n) �

√√√√ 1

C

C∑c=1

‖ pmT (n)‖2

2. (45)

The D-optimal value at the nth measurement epoch is de-noted by log det [Yn]. The average computational time willbe denoted by tc.

A. Gauss–Newton MAP Estimator for UT Localization

The Gauss–Newton MAP estimator developed inSection IV-A assumes the measurement noise to be inde-pendent across different UT range measurements. There-fore, each UT may be estimated independently. This sec-tion evaluates the Gauss–Newton MAP estimator on a sin-gle UT for a predescribed ASV trajectory. The superscriptm = 1 is dropped in this section. A Monte Carlo analysis isperformed, simulating 500 runs of the MAP estimation al-gorithm. For each run, the ASV made N = 4 measurementsfrom the locations listed in Table I to localize pT .

The UT was fixed at pT = [17, 15, −6]T. The value0 pT was drawn from a Gaussian distribution with covari-ance 0PT = blkdiag [100, 100, 4] and mean pT . The USBLmeasurement noise, defined in (4), has σ = 0.1 m based onthe SeaTrac x150 USBL product sheet [40].

The actual 95% confidence ellipse was calculated inthe Easting–Northing (E–N) plane by substituting the

2840 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 6: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

Fig. 1. Scatter plot of pT (4) in the E − N plane.

Fig. 2. Histogram of (44) at measurement epoch n = N = 4 for allMonte Carlo runs. The orange curve displays a χ2

3 pdf.

upper-left 2 × 2 block of (22) and associated elements of(21) into the equations of Appendix B. Also, the estimated95% confidence ellipse fitted to 500 samples of pT (4)was calculated in the E–N plane using the equations inAppendix B. Fig. 1 compares these ellipses and shows themto be comparable.

The NEES of the MAP estimator will be evaluated next.It is expected that ε(n) ∼ χ2

3M (i.e., has a chi-squared proba-bility density function (pdf) with 3M degrees of freedom).Fig. 2 displays a normalized histogram of (44) at epochn = N = 4 across all Monte Carlo runs. A curve corre-sponding to a χ2

3 distribution is overlayed onto this his-togram. The curve fits the values of (44) closely, furthervalidating that the estimator is performing correctly.

B. MSLA OMLP Evaluation for M = 1

This section evaluates the MSLA OMLP for the sin-gle UT environment (the superscript m is dropped fromthis section). Again, a Monte Carlo analysis is performed.For this analysis, 1000 runs consisting of 50 (i.e., N = 50)acoustic measurements were simulated. Table II summa-rizes the simulation settings. The initial UT location 0 pT

is drawn from a Gaussian distribution with mean pT andcovariance 0PT = blkdiag

[104, 104, 4

]for all Monte Carlo

runs. Three look-ahead lengths were simulated: k = 1, 2,

TABLE IISingle UT Localization Simulation

Settings

Fig. 3. Evolution of (a) D-optimal value and (b) localization RMSEE(45) as functions of measurement number n. These values are shown for

various values of k (k = 0 denotes a random ASV trajectory).

and 3. Additionally, a random ASV trajectory (k = 0) wassimulated, subject to the constraints in (34) to demonstratethe benefit gained from planning the measurement locationsto improve UT localization.

It is expected that the D-optimal value obtained by us-ing larger values of k will be larger than those obtainedfor smaller k at the cost of longer average computationaltime. However, due to the linearizations made in (29), theremay be discrepancies in this trend at earlier measurementepochs. This trend in D-optimal value should in turn yieldsmaller values of (45). Additionally, it is intuitive to ex-pect (45) and the D-optimal value to approach lower andupper bounds, respectively, determined by the number ofmeasurements N , limiting the performance enhancementsachievable by increasing k.

Fig. 3 displays the D-optimal value and (45) for variousvalues of k as functions of n. These values are averagedover all Monte Carlo runs. Table III presents results forthis simulation at specific epochs indicated by n. Rows of

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2841

Page 7: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

TABLE IIISingle UT Localization Simulation Results

Fig. 4. Histogram of (44) at measurement epoch n = N = 50 for allMonte Carlo runs in a single UT setup. The orange curve displays a

χ23 pdf.

this table organize data for different look-ahead durationindicated by k.

It is clear from Table III that the expected data trendholds in general. However, there are discrepancies in certainvalues at certain epochs. Nevertheless, these discrepanciesare insignificant considering the magnitude of the standarddeviation associated with these values at these epochs (thelast column of Table III). The value tc (in seconds) increasesfrom 0.18 s when k = 1, to 0.43 s when k = 2, to 0.88 swhen k = 3.

It is worthwhile to verify the consistency of the estima-tor developed in Section IV-A when applied to the MSLAOMLP problem. The normalized position error (44) maybe examined as in Section V-A. It is expected that (44) willbe distributed according to a chi-squared pdf with 3M = 3degrees of freedom. Fig. 4 presents the histogram of thisvalue at measurement epoch n = N over all Monte Carloruns for the k = 2 planner. It is clear that (44) follows theexpected distribution, suggesting the estimator is consis-tent. This trend was observed for the values of k = 1 andk = 3 as well.

It was empirically demonstrated in [36] that the solutionto the greedy OMLP problem is achieved when h(N + 1)is collinear with the eigenvector associated with the largesteigenvalue of YN . This solution could be achieved by in-creasing the maximum allowable distance the ASV couldmove between measurements. In the new framework,however, this solution requires varying the ASV’s forward

Fig. 5. Optimally planned ASV measurement locations determined atepoch n = 1 for k = 1. The green sequence of diamonds indicates theoptimal solution p∗

V (2) as a function of vx . The red asterisk denotespV (1). The eigenvector corresponding to the largest eigenvalue of Y1 is

denoted by the dashed purple arrow pointing in the direction of the majoraxis of the purple ellipse.

velocity, denoted by vx , while fixing ωmax = 180◦T

. Thisdemonstration focuses on the planned ASV location afterthe first measurement has been made (i.e., N = 1). Fig. 52

displays the resulting p∗V (2) as vx is varied from 1 to 80 m/s

in increments of 1 m/s. To avoid clutter, this figure dis-plays p∗

V (2) determined using a subset of vx ∈ [1, 80] m/s.It is clear that this result is consistent with that presentedin [36].

C. MSLA OMLP Evaluation for M = 4

This section evaluates the MSLA OMLP for a multi-UT environment with M = 4. A Monte Carlo analysiswith 1000 runs is performed in an identical fashion tothat in Section V-B with the simulation settings tabulatedin Table IV.

Fig. 6 displays the D-optimal value and (45) for variousvalues of k as functions of n. These values are averaged overall Monte Carlo runs. These values are shown for specificepochs in Table V. Columns of this table present (45) of UTm = 2, the D-optimal value, and standard deviation in D-optimal value. Rows of this table organize data for differentlook-ahead duration k at specific epochs.

Again, it can be seen that the expected localizationtrends hold in general. Discrepancies in the D-optimal valuetrend may again be ignored considering the magnitude ofthe standard deviation associated with these values at these

2Two optimal measuring locations p∗V (N + 1) exist. Fig. 5 focuses on

those solutions that lie along the axis pointing northeast.

2842 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 8: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

TABLE IVMultiple UT (M = 4) Localization

Simulation Settings

Fig. 6. Evolution of (a) D-optimal value and (b) localization RMSEE(45) as functions of measurement number n for the multiple UT

environment. The D-optimal value is averaged over Monte Carlo runs.(b) Results for UT m = 2, as those for other UTs follow similar trends.

epochs (the last column in Table V). The average compu-tational time tc increases from 0.24 s when k = 1, to 0.72 swhen k = 2, to 1.62 s when k = 3.

As with the single UT case, the estimator is consistentwhen applied in an environment with multiple UTs. Fig. 7presents the histogram of (44). It is clear that this histogramfollows a χ2

3M distribution, indicating that the estimator isconsistent.

Fig. 8 displays the measurement paths for each valueof k during a single Monte Carlo run. The ASV’s ini-tial position was pV (1) = [−3, 9, 0]T m with yaw angleφV (1) = −30◦. Note that all optimally planned trajecto-ries tend to orbit around the set of UTs. It is interesting tonote the similarity between the trajectories planned using

TABLE VMultiple UT (M = 4) Localization Simulation Results

Fig. 7. Histogram of (44) at measurement epoch n = N = 50 for allMonte Carlo runs in a multiple (M = 4) UT setup. The orange curve

displays a χ212 pdf.

k = 1 and k = 2, and how they differ from that plannedwith k = 3. These trajectories begin to diverge at n = 30.From the data in Table V, it can be seen that both (45) andD-optimal value are very similar across all k ≥ 1 (i.e., forall optimally planned ASV trajectories) at this epoch.

VI. EXPERIMENTAL RESULTS

This section presents experimental results for the UTlocalization and MSLA OMLP problems. The data collec-tion scheme as well as analysis and description are pro-vided. This section will use the same quantitative metricsof localization RMSEE (45) and D-optimal value.

A. Data Collection

Data collection for both single and multiple transpon-der environments occurred on January 4, 2018 along pier169 at SPAWAR SSC Pacific, San Diego, California, USA.Fig. 9 illustrates the testing environment. Two SeaTrac x010acoustic beacons acted as UTs and were fixed to the pier ata depth of 1 m.

A manned surface craft equipped with a SeaTrac x150USBL beacon and Hemisphere V104TM satellite-based aug-mentation system GPS Compass maneuvered in the oceannear the pier while ranging to each UT. The GPS re-

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2843

Page 9: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

Fig. 8. Simulated ASV trajectories for various values of k. The initialASV location is displayed as a green dot with a black outline.

Fig. 9. Top-down view of the testing area and acoustic ranginglocations as measured by the Hemisphere V104TM GPS Compass. Theorange dots are locations where USBL range measurements were made

from the ASV to the UTs. The green push-pin markers display theconfiguration of UTs during data collection.

ceiver computed differential GPS (DGPS) position esti-mates, which were accurate to 1 m. Range data and GPSfixes were acquired at 0.67 Hz and 1 Hz, respectively, andwere written to two separate files during data collection.All data was time-stamped with UTC time, which was usedto align data in postprocessing. Ground truth positions ofthese UTs were determined by averaging GPS fixes at eachUT mounting point over periods of 3 min.

B. Noise Analysis

The range measurement standard deviation was esti-mated using sets of approximately 250 range measurementsaccording to the procedure described in [36]. These valueswere estimated as σ1 = 62 mm and σ2 = 113 mm. For the

following, the maximum of both standard deviation esti-mates is used as the standard deviation associated with theUSBL range sensor (i.e., σUSBL ≡ max{σi} = 113 mm). Itwas also noted that r1 is biased with respect to ground truthby approximately 0.2 m, which may be explained by theuncertainty associated with the GPS-derived estimates.

To compensate for these errors, one must first return tothe models. First, let pV (n) represent the estimate of theASV location at measurement epoch n as determined viaGPS. Additionally, evaluate the Jacobian vector hm(n) of(8) at the measurement location pV (n)

hm(n) =[

pmT (N) − pV (n)

]T

rm(n)∈ R

1×3. (46)

Recall the model for the USBL range measurement in (3).The range can be approximated using a first-order Taylorseries expansion around the current estimated UT loca-tion and ASV position at the nth measurement epoch (i.e.,pm

T (N) and pV (n)). Define rm(n) [cf. (6)] as

rm(n) � ‖ pmT (N) − pV (n)‖2. (47)

The approximated range is

zm(n) ≈ rm(n) + hm(n)[δ pm

T − δ pV (n)] + w(n), (48)

where

δ pmT � pm

T (N) − pmT (N) (49)

δ pV (n) � pV (n) − pV (n). (50)

The residual is defined as

δzm(n) � zm(n) − rm(n) (51)

= hm(n)δ pmT + w(n) − hm(n)δ pV (n). (52)

Define q(n) � hm(n)δ pV (n) as the error due to uncertaintyin the GPS solution. The total error in the range measure-ment is ε(n) � w(n) + q(n). It is reasonable to assume thatw(n) and q(n) are white and are not correlated with eachother.

With this in mind, the measurement noise may now bemodeled as ε(n) ∼ N (0, σ 2

ε ), where σ 2ε = σ 2

USBL + σ 2GPS.

The value σ 2GPS is the variance of q(n). Note that, by con-

struction of q(n), σ 2GPS is the GPS position error projected

onto hm(n). For purposes of analysis, σGPS is set to be aconstant (i.e., σ 2

GPS = 1 m). With this assumption, the totalmeasurement standard deviation is σε = 1.006 m.

It is worth noting that in some practical applications,the noise covariance matrices may be poorly known andthe “luxury” of characterizing them in this manner could beinfeasible. To circumvent this issue, adaptive filters couldbe used [41].

C. Processed Results: Evaluation of MSLA OMLP(M = 1)

This section evaluates the MSLA OMLP approach pre-sented in Section IV-B for a single UT environment. Allpositions are represented in a local East, North, Up co-ordinate frame centered at the earliest collected GPS lo-

2844 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 10: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

TABLE VISingle UT (M = 1) Localization

Experimental Settings

Fig. 10. Evolution of (a) D-optimal value and (b) localization RMSEE(45) as functions of measurement number n for the single UT

experimental environment. These values are shown forvarious values of k.

cation of the ASV. This point, represented in the geode-tic frame, is g pV = [32.705◦N, 117.236◦W, −1.478 m].Again, a Monte Carlo analysis is performed. This analy-sis consists of 100 Monte Carlo runs. Each run has randominitial estimates of transponder locations. An ASV trajec-tory comprising 50 measurement locations was producedfor each run. The initial ASV position is held constant overall runs. Instead of simulating range measurements as wasdone in Section V-B, this section uses the range measure-ments collected at SPAWAR. The experimental settings aretabulated in Table VI.

Fig. 10 presents the D-optimal value and localizationRMSEE (45). These values are averaged over all MonteCarlo runs. Localization results at specific epochs are givenin Table VII. This table is organized identically to Table III.Note that, similar to the simulation analysis, one expects

TABLE VIISingle UT Localization Experimental Results

TABLE VIIIMultiple UT (M = 2) Localization

Experimental Settings

that a larger value of k corresponds to better localizationaccuracy [i.e., smaller values of (45)]. This holds in general;however, discrepancies at all epochs for k = 2 can be seen.Similarly, values of (45) are significantly larger than thoseobserved in Section V-B. The most likely cause is that theGPS measurement of the UT position is not correct eitherdue to GPS errors or the challenge of placing the GPSreceiver directly above the UT. The average convergencetime tc increases from 0.19 s when k = 1, to 0.39 s whenk = 2, to 0.72 s when k = 3.

D. Processed Results: Evaluation of MSLA OMLP(M = 2)

This section evaluates the MSLA OMLP approach pre-sented in Section IV-B for a multiple UT environment withM = 2. A Monte Carlo analysis is performed with the samesetup as in Section VI-C. The experimental settings aretabulated in Table VIII. As with in the single UT case,σε = 1 m.

Fig. 11 presents (45) and averaged D-optimal value asfunctions of measurement epoch. Values at selected epochsare provided in Table IX. As in previous cases, unexpectedtrends in the D-optimal value are insignificant consideringthe magnitude of the standard deviation associated withthese values at these epochs. The values of (45) are largerthan those in Section V-C. This increase is likely associated

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2845

Page 11: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

Fig. 11. Evolution of (a) averaged D-optimal value and (b) localizationRMSEE (45) as functions of measurement number n for the multiple UT(M = 2) experimental environment. These values are shown for various

values of k. Note that (b) presents (45) for UT m = 1 only, as that for UTm = 2 is similar.

TABLE IXMultiple UT (M = 2) Localization Experimental Results

with errors in the GPS measurement of the UT locations.The convergence time tc increases from 0.21 s, to 0.48 s, to0.99 s for the one-step, two-step, and three-step look-aheadplanners, respectively.

Fig. 12 demonstrates the measurement paths determinedby varying the value of k, along with the randomly plannedtrajectory, during a single Monte Carlo run.

The ASV’s initial position is pV (1) = [8.62,

−6.77, 0]T m with yaw angle φV (1) = 13◦. As inSection V-C, it can be seen that all optimally planned ASVtrajectories orbit around the set of UTs.

Fig. 12. ASV trajectories for various values of k using experimentaldata. UT locations are denoted by red symbols outlined in black. Theinitial ASV location is displayed as a green dot with a black outline.

VII. CONCLUSION AND DISCUSSIONS

This paper provided a MAP estimation algorithm for lo-calizing UTs as well as a strategy for the ASV to determinethe best future locations at which acoustic ranges must bemade to localize the UTs. The simulation results demon-strated the performance of the proposed MAP estimator andthe OMLP strategy. The experimental results based on datacollected at SPAWAR SSC Pacific demonstrated the per-formance of the MSLA OMLP strategy in an environmentcontaining two UTs. Both simulation and experimental re-sults demonstrated the localization benefit gained by theASV when planning measurement locations as opposedto randomly choosing measurement locations. Addition-ally, localization accuracy is shown to increase for largerlook-ahead distances at the cost of larger computationalcost. It is interesting to note, however, that there are dimin-ishing returns to larger look-ahead distances when consid-ering longer measurement paths. In general, as the numberof measurements increases, the marginal improvement inlocalization accuracy obtained from larger look-ahead dis-tances is outweighed by the additional computational cost.Future research endeavors seek to adapt this problem tocases where an omnidirectional sensor may not be avail-able. How might the use of a directional sensor change theplanning method?

APPENDIX APROPERTIES OF MATRIX DETERMINANTS

Consider the block matrix Q ∈ RK×K , constructed as

Q =[

A B

C D

](53)

2846 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 12: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

where A ∈ RM×M , B ∈ R

M×N , C ∈ RN×M , D ∈ R

N×N ,and N = K − M . When A and D are invertible

det [Q] = det [A] det[D − CA−1B

](54)

= det [D] det[A − BD−1C

]. (55)

Now, consider the matrix Q′, constructed as

Q′ =[

I B

−CT I

]. (56)

Using (54) and (55)

det [Q]′ = det [I] det[I + CTI−1B

](57)

= det [I] det[I + BI−1CT] . (58)

From (57) and (58), it is clear that

det[I + CTI−1B

] = det[I + BI−1CT] . (59)

APPENDIX BCALCULATION OF CONFIDENCE ELLIPSES FROM CO-VARIANCE MATRICES

One may calculate the confidence ellipse correspond-ing to the estimate p ∈ R

2 and associated covariance ma-trix P ∈ R

2×2 as follows [42]. First, determine the largesteigenvalue and associated eigenvector (denoted λmax andvmax = [xmax, ymax]T, respectively) of P. Define the con-stants a � c

√λmax and b � c

√λmin, where c is a constant

scaling factor corresponding to the desired confidence levelof our ellipse. For a confidence level of 95%, c = 2.4477.Points along the confidence ellipse are now calculated as[

x

y

]= p + R(φ)

[a cos(θ)

b sin(θ)

], ∀ θ ∈ [0, 2π), (60)

where

φ � atan2(ymax, xmax) (61)

R(φ) �[

cos(φ) − sin(φ)

sin(φ) cos(φ)

]. (62)

ACKNOWLEDGMENT

The authors would like to thank J. Khalife and E. Aghapourfor helpful discussions and also the unmanned marine vehi-cles (UMV) lab and the members of HAMMER at SPAWARSSC Pacific for help with data collection.

REFERENCES

[1] L. Paull, S. Saeedi, M. Seto, and H. LiAUV navigation and localization: A reviewIEEE J. Ocean. Eng., vol. 39, no. 1, pp. 131–149, Jan. 2014.

[2] S. Kim, B. Ku, W. Hong, and H. KoPerformance comparison of target localization for active sonarsystemsIEEE Trans. Aerosp. Electron. Syst., vol. 44, no. 4, pp. 1371–1380, Oct. 2008.

[3] Y. Huang, Y. Zhang, B. Xu, Z. Wu, and J. ChambersA new adaptive extended Kalman filter for cooperative local-izationIEEE Trans. Aerosp. Electron. Syst., vol. 54, no. 1, pp. 353–368,Feb. 2018.

[4] P. Zhan, D. Casbeer, and A. SwindlehurstAdaptive mobile sensor positioning for multi-static target track-ingIEEE Trans. Aerosp. Electron. Syst., vol. 46, no. 1, pp. 120–132,Jan. 2010.

[5] J. Morales and Z. KassasOptimal receiver placement for collaborative mapping of sig-nals of opportunityIn Proc. ION GNSS Conf., Sep. 2015, pp. 2362–2368.

[6] D. Moreno-Salinas, A. Pascoal, and J. ArandaOptimal sensor placement for acoustic underwater target posi-tioning with range-only measurementsIEEE J. Ocean. Eng., vol. 41, no. 3, pp. 620–643, Jul.2016.

[7] X. Fang, W. Yan, and W. ChenSensor placement for underwater source localization with fixeddistancesIEEE Geosci. Remote Sens. Lett., vol. 13, no. 9, pp. 1379–1383,Sep. 2016.

[8] J. Khalife and Z. KassasOptimal sensor placement for dilution of precision minimiza-tion via quadratically constrained fractional programmingIEEE Trans. Aerosp. Electron. Syst., to be published.

[9] D. UcinskiOptimal Meas. Methods for Distributed Parameter System Iden-tification. Boca Raton, FL, USA: CRC Press, 2005.

[10] D. Moreno-Salinas, N. Crasta, M. Ribiero, B. Bayat, A. Pascoal,and J. ArandaIntegrated motion planning, control, and estimation for range-based marine vehicle positioning and target localizationIn Proc. IFAC Conf. Control Appl. Marine Syst., Sep. 2016,vol. 49, no. 23, pp. 34–40.

[11] Z. Kassas and T. HumphreysMotion planning for optimal information gathering in oppor-tunistic navigation systemsIn Proc. AIAA Guid., Navig., Control Conf., Aug. 2013,pp. 4551–4565.

[12] Z. Kassas, T. Humphreys, and A. ArapostathisGreedy motion planning for simultaneous signal landscapemapping and receiver localizationIEEE J. Sel. Topics Signal Process., vol. 9, no. 2, pp. 246–258,Mar. 2015.

[13] J. Morales and Z. KassasOptimal collaborative mapping of terrestrial transmitters: Re-ceiver placement and performance characterizationIEEE Trans. Aerosp. Electron. Syst., vol. 54, no. 2, pp. 992–1007, Apr. 2018.

[14] B. Ferriera, A. Matos, H. Campos, and N. CruzLocalization of a sound source: Optimal positioning of sensorscarried on autonomous surface vehiclesIn Proc. IEEE OCEANS Conf., Sep. 2013, pp. 1–8.

[15] J. Perez-Ramirez, D. Borah, and D. VoelzOptimal 3-D landmark placement for vehicle localization usingheterogeneous sensorsIEEE Trans. Veh. Technol., vol. 62, no. 7, pp. 2987–2999, Sep.2013.

[16] J. Morales, P. Roysdon, and Z. KassasSignals of opportunity aided inertial navigationIn Proc. ION GNSS Conf., Sep. 2016, pp. 1492–1501.

[17] J. Morales, J. Khalife, and Z. KassasCollaborative autonomous vehicles with signals of opportunityaided inertial navigation systemsIn Proc. ION Int. Tech. Meet. Conf., Jan. 2017, 805–818.

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2847

Page 13: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

[18] F. Hidalgo and T. BrunlReview of underwater SLAM techniquesIn Proc. Int. Conf. Autom., Robot. Appl., Feb. 2015, pp. 306–311.

[19] P. Miller, J. Farrell, Y. Zhao, and V. DjapicAutonomous underwater vehicle navigationIEEE J. Ocean. Eng., vol. 35, no. 3, pp. 663–678, Jul. 2010.

[20] J. Zhang, C. Shi, K. Tang, and Y. HanResearch and implementation on multi-beacon aided AUV in-tegrated navigation algorithmIn Proc. IEEE OCEANS Conf., Jun. 2017, pp. 1–5.

[21] V. Djapic et al.Novel method for underwater navigation aiding using a com-panion underwater robot as a guiding platformsIn Proc. IEEE OCEANS Conf., Jun. 2013, pp. 1–10.

[22] F. Teixeira and A. PascoalMagnetic navigation and tracking of underwater vehiclesIFAC Proc. Volumes, vol. 46, no. 33, pp. 239–244, 2013.

[23] M. Ejaz, J. Iqbal, N. Ahsan, and A. NawazRobust geomagnetic aided inertial navigation of underwatervehicles using the ICP algorithmIn Proc. Asia Pac. Conf. Comput. Intell. Ind. Appl., Nov. 2009,vol. 1, pp. 257–262.

[24] M. Wu and J. YaoAdaptive UKF-SLAM based on magnetic gradient inversionmethod for underwater navigationIn Proc. Int. Conf. Unmanned Aircraft Syst., Jun. 2015, pp. 839–843.

[25] F. Teixeira, J. Quintas, and A. PascoalAUV terrain-aided Doppler navigation using complementaryfilteringIn Proc. IFAC Conf. Maneuvering Control Marine Craft,Sep. 2012, vol. 45, pp. 313–318.

[26] J. Quintas, F. Teixeira, and A. PascoalMagnetic signal processing methods with application to geo-physical navigation of marine robotic vehiclesIn Proc. IEEE OCEANS Conf., Sep. 2016, pp. 1–8.

[27] T. Zhou, D. Peng, C. Xu, W. Zhang, and J. ShenAdaptive particle filter based on Kullback-Leibler distance forunderwater terrain aided navigation with multi-beam sonarIET Radar, Sonar Navigat., vol. 12, no. 4, pp. 433–441, 2018.

[28] P. Batista, C. Silvestre, and P. OliveiraSingle range aided navigation and source localization: Observ-ability and filter designSyst. Control Lett., vol. 60, no. 8, pp. 665–673, Jun. 2011.

[29] Z. Kassas and T. HumphreysReceding horizon trajectory optimization in opportunistic nav-igation environmentsIEEE Trans. Aerosp. Electron. Syst., vol. 51, no. 2, pp. 866–877,Apr. 2015.

[30] Z. Kassas and T. HumphreysObservability analysis of collaborative opportunistic navigationwith pseudorange measurementsIEEE Trans. Intell. Transp. Syst., vol. 15, no. 1, pp. 260–273,Feb. 2014.

[31] H. Li, P. Xie, and W. YanReceding horizon formation tracking control of constrainedunderactuated autonomous underwater vehiclesIEEE Trans. Ind. Electron., vol. 64, no. 6, pp. 5004–5013, Jun.2017.

[32] S. Huang, N. Kwok, G. Dissanayake, Q. Ha, and G. FangMulti-step look-ahead trajectory planning in SLAM: Possibilityand necessityIn Proc. IEEE Int. Conf. Robot. Autom., Apr. 2005, pp. 1091–1096.

[33] C. Shen, Y. Shi, and B. BuckhamIntegrated path planning and tracking control of an AUV: Aunified receding horizon optimization approachIEEE/ASME Trans. Mechatron., vol. 22, no. 3, pp. 1163–1173,Jun. 2017.

[34] M. Dunbabin and R. LamentAdaptive receding horizon control for a high-speed autonomoussurface vehicle in narrow waterwaysIn Proc. IEEE Conf. Control Technol. Appl., Aug. 2017,pp. 235–240.

[35] G. Ferri, A. Munafo, R. Goldhahn, and K. LePageA non-myopic, receding horizon control strategy for an AUVto track an underwater target in a bistatic sonar scenarioIn Proc. IEEE Conf. Decis. Control, Dec. 2014, pp. 5352–5358.

[36] J. Garcia, J. Farrell, and Z. KassasOptimal measurement location planning for localizing under-water transpondersIn Proc. IEEE/ION Position Location Navigat. Symp.,Apr. 2018, pp. 480–486.

[37] Y. Bar-Shalom, T. Kirubarajan, and X. LiEstimation with Applications to Tracking and Navigation.Hoboken, NJ, USA: Wiley, 2002.

[38] S. KayFundamentals of Statistical Signal Processing: Estimation The-ory. Englewood Cliffs, NJ, USA: Prentice-Hall, 1993.

[39] S. Boyd and L. VanderbuergheConvex Optimization. Cambridge, U.K.: Cambridge Univ.Press, 2004.

[40] SeaTrac Micro-USBL Tracking and Data Modems, rev. 5,Blueprint Design Engineering Ltd, Mar. 2016. [Online]. Avail-able: https://www.blueprintsubsea.com/seatrac/support.php

[41] Y. Huang, Y. Zhang, B. Xu, Z. Wu, and J. A. ChambersA new adaptive extended Kalman filter for cooperative local-izationIEEE Trans. Aerosp. Electron. Syst., vol. 54, no. 1, pp. 353–368,Feb. 2018.

[42] W. Hoover and M. RockvilleAlgorithms for confidence circles and ellipsesU.S. Dept. Commerce, Nat. Oceanic Atmospheric Admin.,Washington, DC, USA, Tech. Rep. National Ocean Service(NOS) 107 Coast & Geodetic Survey (CGS) 3, Sep. 1984.

Jesse R. Garcia (S’16) received the B.S.(Hons.) degree in electrical and computer engi-neering from the University of California, Riverside, Riverside, CA, USA, in 2016, wherehe is currently working toward the M.S. degree in electrical engineering.

He is a member of the Autonomous Systems Perception, Intelligence, and Naviga-tion (ASPIN) Laboratory. His research interests include underwater robotics, estimation,autonomous vehicles, and navigation.

2848 IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 55, NO. 6 DECEMBER 2019

Page 14: Autonomous Surface Vehicle Multistep Look-Ahead Measurement …aspin.eng.uci.edu/papers/Autonomous_surface_vehicle... · 2019-12-18 · Autonomous Surface Vehicle Multistep Look-Ahead

Jay A. Farrell (F’08) received the B.S. degree in physics and electrical engineeringfrom Iowa State University, Ames, IA, USA, in 1986, and the M.S. and Ph.D. degrees inelectrical engineering from the University of Notre Dame, Notre Dame, IN, USA, in 1998and 1989, respectively.

He is currently a Professor with the Department of Electrical and Computer Engineer-ing, University of California, Riverside, where he has served three terms as DepartmentChair and is currently a Associate Dean. He has authored more than 250 technical publi-cations and three books.

Dr. Farrell was the recipient of the Engineering Vice President’s Best Technical Publi-cation Award at Charles Stark Draper Lab (1989–1994) in 1990, and Recognition Awardsfor Outstanding Performance and Achievement in 1991 and 1993. He was named a GNSSLeader to Watch for 2009–2010 by GPS World Magazine and a winner of the ConnectedVehicle Technology Challenge by the U.S. Department of Transportation’s (DOT’s) Re-search and Innovative Technology Administration in July 2011. He is a Fellow of AAAS,a Distinguished Member of IEEE CSS. He was the IEEE Control Systems Society asGeneral Chair of IEEE CDC 2012, and as President in 2014. He is currently an AmericanAutomatic Control Council Vice President in 2018.

Zaher M. Kassas (S’98–M’08–SM’11) received the B.E.(Hons.) degree in electricalengineering from the Lebanese American University, Beirut, Lebanon, in 2001, the M.S.degree in electrical and computer engineering from The Ohio State University, Columbus,OH, USA, in 2003, and the M.S.E. degree in aerospace engineering and the Ph.D. degreein electrical and computer engineering from The University of Texas at Austin, Austin,TX, USA, in 2010 and 2014, respectively.

He is currently an Assistant Professor with the University of California, Riverside,Riverside, Ca, USA, and the Director of the Autonomous Systems Perception, Intelligence,and Navigation (ASPIN) Laboratory. His research interests include cyber-physical systems,estimation theory, navigation systems, autonomous vehicles, and intelligent transportationsystems.

Dr. Kassas was the recipient of the National Science Foundation (NSF) Faculty EarlyCareer Development Program (CAREER) award in 2018, the Office of Naval Research(ONR) Young Investigator Program (YIP) award in 2019, and the 2018 IEEE Walter FriedAward and the 2018 ION Samuel Burka Award.

Michael T. Ouimet received the B.S degree in mechanical engineering from the Universityof Notre Dame, Notre Dame, IN, USA, and the M.S. and Ph.D. degrees in mechanical andaerospace engineering from University of California, San Diego, La Jolla, CA, USA.

He is currently with the Unmanned Maritime Vehicles lab, SSC Pacific as a Re-searcher. His main research interests include autonomy for multiagent unmanned systems,especially in the intersection of control theory, artificial intelligence, and machine learningfor robotics.

GARCIA ET AL.: ASV MSLA MEASUREMENT LOCATION PLANNING 2849