tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7....

29
Tightly Coupled Ultrashort Baseline and Inertial Navigation System for Underwater Vehicles: An Experimental Validation M. Morgado Department of Electrical and Computer Engineering, Instituto Superior T´ ecnico, Universidade T´ ecnica de Lisboa, Lisbon, Portugal e-mail: [email protected] P. Oliveira Department of Mechanical Engineering, Instituto Superior T´ ecnico, Universidade T´ ecnica de Lisboa, Lisbon, Portugal e-mail: [email protected] C. Silvestre Department of Electrical and Computer Engineering, Instituto Superior T´ ecnico, Universidade T´ ecnica de Lisboa, Lisbon, Portugal Department of Electrical and Computer Engineering, Faculty of Science and Technology, University of Macau, Taipa, Macau, China e-mail: [email protected] Received 14 November 2011; accepted 4 October 2012 This paper presents a new ultrashort baseline (USBL) tightly coupled integration technique to enhance error estimation in low-cost strapdown inertial navigation systems (INSs), with application to underwater vehicles. In the proposed strategy, the acoustic array spatial information is directly exploited in an extended Kalman filter (EKF) implemented in a direct feedback structure. Instead of using the USBL position fixes or computed range and elevation/bearing angles to correct the INS error drifts, as in classical loosely coupled strategies, the novel tightly coupled strategy directly embeds in the EKF the round-trip-time and time-difference-of-arrival of the acoustic signals arriving at the onboard receivers. The enhanced performance of the proposed filtering technique is evidenced both through extensive numerical simulations and with experimental data obtained in field tests at sea. The tightly coupled filter is also shown to be able to operate closer to theoretical performance lower bounds, such as the posterior Cram´ er-Rao lower bound, using Monte-Carlo simulations. This paper details the design and description of an USBL/INS prototype to be used as a low-cost navigation system, including the acoustic processing and positioning system, fully developed in-house. The developed system validates the usage of the proposed technique with real data in real world operation scenarios, and its enhanced performance compared to classical strategies is evaluated experimentally (median improvement level of 15% in typical operating conditions). Improved and faster convergence to nominal trajectories from multiple initial conditions, as well as enhanced accelerometer and rate gyros estimation capabilities, are also demonstrated experimentally for the new tightly coupled filter. C 2012 Wiley Periodicals, Inc. 1. INTRODUCTION Worldwide, there has been an increasing interest in the use of underwater vehicles to expand the ability to accurately survey large ocean areas. Routine operations such as envi- ronmental monitoring, surveillance, underwater inspection of estuaries, harbors and pipelines, and geological and bi- ological surveys—see Pascoal et al. (2000)—are tasks com- monly performed at present either by remotely operated vehicles (ROVs) or by autonomous underwater vehicles (AUVs). The use of these robotic platforms requires low- cost, compact, high-performance, robust navigation sys- tems that can accurately estimate the vehicle’s position and attitude. In fact, the design and implementation of Direct correspondence to: M. Morgado, e-mail: marcomorgado@ isr.ist.utl.pt navigation systems stands out as one of the most critical steps toward the successful operation of autonomous ve- hicles and marine robotic vehicles. The ability to perform the aforementioned procedures at increasing depths, often life-threatening or impossible for humans, makes marine robotic vehicles stand out as one of the strongest areas of investigation and efforts by the robotics scientific commu- nity (Bowen et al., 2009; Jalving et al., 2003; Lurton and Millard, 1994; Napolitano et al., 2005). For other interesting and detailed surveys on underwater vehicle navigation and its relevance, see Whitcomb (2000) and more recently Kin- sey et al. (2006). This paper presents a new tightly coupled integration technique to enhance error estimation in strap- down inertial navigation systems (INSs) for underwater ve- hicles, in which the acoustic array spatial information is di- rectly exploited in the navigation system algorithm, while Journal of Field Robotics 30(1), 142–170 (2013) C 2012 Wiley Periodicals, Inc. View this article online at wileyonlinelibrary.com DOI: 10.1002/rob.21442

Upload: others

Post on 28-Feb-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Tightly Coupled Ultrashort Baseline and InertialNavigation System for Underwater Vehicles: AnExperimental Validation

• • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • •

M. MorgadoDepartment of Electrical and Computer Engineering, Instituto Superior Tecnico, Universidade Tecnica de Lisboa, Lisbon, Portugale-mail: [email protected]. OliveiraDepartment of Mechanical Engineering, Instituto Superior Tecnico, Universidade Tecnica de Lisboa, Lisbon, Portugale-mail: [email protected]. SilvestreDepartment of Electrical and Computer Engineering, Instituto Superior Tecnico, Universidade Tecnica de Lisboa, Lisbon, PortugalDepartment of Electrical and Computer Engineering, Faculty of Science and Technology, University of Macau, Taipa, Macau, Chinae-mail: [email protected]

Received 14 November 2011; accepted 4 October 2012

This paper presents a new ultrashort baseline (USBL) tightly coupled integration technique to enhance errorestimation in low-cost strapdown inertial navigation systems (INSs), with application to underwater vehicles.In the proposed strategy, the acoustic array spatial information is directly exploited in an extended Kalmanfilter (EKF) implemented in a direct feedback structure. Instead of using the USBL position fixes or computedrange and elevation/bearing angles to correct the INS error drifts, as in classical loosely coupled strategies, thenovel tightly coupled strategy directly embeds in the EKF the round-trip-time and time-difference-of-arrivalof the acoustic signals arriving at the onboard receivers. The enhanced performance of the proposed filteringtechnique is evidenced both through extensive numerical simulations and with experimental data obtained infield tests at sea. The tightly coupled filter is also shown to be able to operate closer to theoretical performancelower bounds, such as the posterior Cramer-Rao lower bound, using Monte-Carlo simulations. This paperdetails the design and description of an USBL/INS prototype to be used as a low-cost navigation system,including the acoustic processing and positioning system, fully developed in-house. The developed systemvalidates the usage of the proposed technique with real data in real world operation scenarios, and its enhancedperformance compared to classical strategies is evaluated experimentally (median improvement level of 15%in typical operating conditions). Improved and faster convergence to nominal trajectories from multiple initialconditions, as well as enhanced accelerometer and rate gyros estimation capabilities, are also demonstratedexperimentally for the new tightly coupled filter. C© 2012 Wiley Periodicals, Inc.

1. INTRODUCTION

Worldwide, there has been an increasing interest in the useof underwater vehicles to expand the ability to accuratelysurvey large ocean areas. Routine operations such as envi-ronmental monitoring, surveillance, underwater inspectionof estuaries, harbors and pipelines, and geological and bi-ological surveys—see Pascoal et al. (2000)—are tasks com-monly performed at present either by remotely operatedvehicles (ROVs) or by autonomous underwater vehicles(AUVs). The use of these robotic platforms requires low-cost, compact, high-performance, robust navigation sys-tems that can accurately estimate the vehicle’s positionand attitude. In fact, the design and implementation ofDirect correspondence to: M. Morgado, e-mail: [email protected]

navigation systems stands out as one of the most criticalsteps toward the successful operation of autonomous ve-hicles and marine robotic vehicles. The ability to performthe aforementioned procedures at increasing depths, oftenlife-threatening or impossible for humans, makes marinerobotic vehicles stand out as one of the strongest areas ofinvestigation and efforts by the robotics scientific commu-nity (Bowen et al., 2009; Jalving et al., 2003; Lurton andMillard, 1994; Napolitano et al., 2005). For other interestingand detailed surveys on underwater vehicle navigation andits relevance, see Whitcomb (2000) and more recently Kin-sey et al. (2006). This paper presents a new tightly coupledintegration technique to enhance error estimation in strap-down inertial navigation systems (INSs) for underwater ve-hicles, in which the acoustic array spatial information is di-rectly exploited in the navigation system algorithm, while

Journal of Field Robotics 30(1), 142–170 (2013) C© 2012 Wiley Periodicals, Inc.View this article online at wileyonlinelibrary.com • DOI: 10.1002/rob.21442

Page 2: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 143

Figure 1. Typical mission scenario—an underwater vehicle, equipped with an INS and an inverted USBL array, interrogates anearby transponder and listens for its acoustic responses to obtain transponder relative position measurements.

focusing on low-cost and affordable high-performance nav-igation platforms.

INSs provide a self-contained passive means for three-dimensional positioning in the open ocean with excel-lent short-term accuracy. However, unbounded position-ing errors induced by the uncompensated rate gyro andaccelerometer errors degrade INS accuracy over time. Thisperformance degradation and the limitations inherent inlow-cost INSs, attributed to open-loop unbounded esti-mation errors, uncompensated sensor noise, and bias ef-fects, are often tackled by merging additional informationsources with nonlinear filtering techniques. Among a di-verse set of techniques (Bar-Shalom et al., 2001; Crassidis,2006), an extended Kalman filter (EKF) in a direct-feedbackconfiguration (Brown and Hwang, 1997; Vasconcelos et al.,2011) is commonly adopted to estimate and compensate forthe accumulation of the INS integration errors. Within themultitude of available aiding devices, such as inclinome-ters, magnetic compasses, Doppler velocity loggers (DVLs),depth pressure sensors, laser range finders, etc., the globalpositioning system (GPS) is a very popular choice and acommonly adopted solution in aerial and land-based ap-plications (Grewal et al., 2007; Sukkarieh et al., 1999). Theopacity (i.e., high attenuation) of the ocean environmentto most electromagnetic signals makes acoustic propaga-tion the preferable method to obtain practical range mea-surements. Other practical navigation methodologies thatdo not use acoustics typically involve surfacing the vehicleregularly to obtain intermittent GPS corrections (Yun et al.,1999). Available underwater acoustic positioning systems(Milne, 1983; Vickery, 1998)—such as long baseline (LBL)systems, which entail cumbersome and time-consuming in-stallation and calibration procedures; hull-mounted shortbaseline (SBL) systems, which have to be rigidly mountedto a vessel hull and are affected by the natural bendingof the hull; and finally ultrashort baseline (USBL) systems,which provide factory-calibrated and fast deployable sys-tems that are suited for low-cost navigation systems—standoften as the primary choice for underwater positioning(Jaffre et al., 2005; Kinsey and Whitcomb, 2004; Lee et al.,2004; Miller et al., 2010; Smith and Kronen, 1997).

This paper addresses the synthesis and design of mod-ern navigation systems with application to underwater ve-hicles, focusing on small arrays of acoustic receivers as themain sensor suite installed onboard the underwater vehi-cle, in particular with what is known as an USBL acous-tic positioning system (Milne, 1983). The considered mis-sion scenarios are illustrated in Figure 1, which displaysan underwater vehicle that is equipped with an INS andan USBL array in an inverted USBL configuration (Vick-ery, 1998) that interrogates a single nearby transponder lo-cated in a known position of the vehicle’s mission area, en-gaging in interrogations over considerable distances rang-ing typically from a few meters to several kilometers. Thisinterrogation scheme to obtain the round-trip-time (RTT)of travel of the acoustic waves allows for the use of low-cost clocks onboard the vehicle to obtain driftless rangemeasurements. Recent advances in underwater navigationrelying on synchronous one-way travel times of acous-tic waves, available from costly higher precision oscilla-tors, were presented by Eustice et al. (2011). In addition topaving the way for future fully autonomous systems thatdo not require surface mission support vessels, invertedUSBL configurations allow for the sound velocity to be con-sidered constant while operating in the same underwaterlayer as the transponders (for instance, bottom operationwhile interrogating a bottom placed transponder). Thesekinds of inverted configurations also allow for other un-derwater autonomous routine operations, such as under-water interventions as well as homing and docking to un-derwater stations (Jaffre et al., 2005; Sanz et al., 2010). Anoverview of USBL/INS tightly coupled technology entrypoints into underwater navigation is presented in Figure 2,where the solution presented herein is built upon state-of-the-art inverted USBL loosely coupled solutions and intro-duces enhanced performance through full tightly coupledINS/USBL configurations. The claimed performance en-hancement of the novel tightly coupled solution is demon-strated in comparison with the state-of-the-art loosely cou-pled USBL, which also estimates the INS inertial sensorsbiases, whereas classical and substandard solutions are notconsidered.

Journal of Field Robotics DOI 10.1002/rob

Page 3: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

144 • Journal of Field Robotics—2013

Figure 2. Overview of coupling technologies on USBL/INS sensor fusion—classical structures, in which the USBL array tracks un-derwater vehicles while mounted on surface vessels, need to relay the positioning information through cables or acoustic modems.Classical and substandard INS solutions typically do not estimate the inertial sensor biases resulting in very poor performance,while state-of-the-art technology offers improved navigation capabilities by moving the USBL array onboard the underwater vehi-cle (inverted USBL) and estimating rate gyros and accelerometer biases. The solution presented herein builds upon state-of-the-artinverted USBL loosely coupled solutions and introduces enhanced performance through full tightly coupled USBL and INS con-figurations.

1.1. Motivation

Typical USBL/INS integration techniques, usually referredto as loosely coupled (Grewal et al., 2007), rely on solv-ing positioning and sensor fusion problems separately,not taking into account the acoustic array geometry inthe navigation system. The new proposed tightly coupledUSBL/INS integration strategy directly exploits the acous-tic array spatial information, resorting to an EKF in a direct-feedback configuration. A loosely coupled system is com-monly known in the literature (Grewal et al., 2007) as amodular system in which each module is able to operateon its own and can be easily decoupled from the others.

A typical example of a loosely coupled system is aGPS positioning device providing world coordinate posi-tion fixes, whereas an INS provides open-loop integrationof the inertial sensors, and the information fusion is per-formed a posteriori. In this framework, the INS does nothave any prior knowledge on what kind of position fix al-gorithm is being applied to the pseudoranges measured toeach orbiting satellite, nor is the GPS aware of to which en-tity it is providing information. In a tightly coupled con-figuration, both entities are aware of each other’s existenceand cooperate, in some sense, to provide enhanced per-formance to the end-user. In a tightly coupled system, theGPS system directly provides the measured satellite pseudoranges to the INS algorithm, whereas the INS algorithm es-timates, among other inertial system errors, the user clockbias and drift, propagation delays, errors derived from at-mospheric effects, and other associated GPS errors (Grewalet al., 2007). Tightly coupled approaches for the GPS/INSnavigation problem for aerial and land vehicles have been

addressed previously in the literature (Knight, 1997; Yi andGrejner-Brzezinska, 2006). Nevertheless, to the best of ourknowledge, this paper and the work presented herein rep-resents the first time that a tightly coupled strategy hasbeen applied to underwater navigation using small arraysof acoustic receivers such as an USBL positioning system.

1.2. Paper Organization

The paper is organized as follows: The core USBL acous-tic positioning sensor is first described in Section 2, and themain aspects of the navigation system and the proposedarchitecture are presented in Section 3. The EKF-based in-ertial error model is introduced in Section 3.1, and boththe new tightly coupled and classical loosely coupled in-tegration strategies are detailed in Section 3.2 while imple-mentation and discretization details are briefly outlined inSection 3.3. Section 4 provides an analysis based on nu-merical simulation results and a comparison to theoreticalperformance lower bounds, namely the posterior Cramer-Rao lower bound (PCRLB). The prototype system designis presented in Section 5 and the experimental evaluationand validation of the proposed technique is reported inSection 6. Finally, Section 7 presents some concluding re-marks and comments on future directions of research.

2. ULTRASHORT BASELINE POSITIONING SYSTEM

This section introduces the main sensor suite adoptedin this work. The USBL sensor consists of a small andcompact array of acoustic transducers that allows for the

Journal of Field Robotics DOI 10.1002/rob

Page 4: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 145

computation of a transponder position in the vehicle coor-dinate frame, based on the travel time of acoustic signalsemitted by the transponder (Milne, 1983). This travel time isobtained from the RTT of travel of acoustic signals from thepinger installed on the vehicle to the transponder placed ata known position and back to the receivers on the USBL ar-ray. Taking into account the quantization performed by theacoustic system, and assuming that the transponder per-forms a similar acoustic processing, the RTT to each of thereceivers on the USBL array is given by

tRTT i = [tp + εt ]Ts+ td + [tr i + εc + εd i]Ts

,

where tp is the nominal travel time from the pinger on thevehicle to the transponder, tr i is the nominal travel timefrom the transponder to the ith acoustic receiver, [·]Ts

rep-resents the acoustic sampling quantization

[x]Ts= Ts round(x/Ts), x ∈ R,

where Ts is the acoustic sampling period (and consequentlythe maximal available time-resolution), and round(·) isthe standard mathematical round-to-nearest-integer opera-tor. The terms εt and εc represent, respectively, the noiseat the transponder and at the receivers (common to allreceivers—it includes transponder-receiver relative motiontime-scaling effects and errors in sound propagation veloc-ity), and εd i captures additional differential error sources,much smaller than the common mode errors. The responsedelay time td is considered to be known, so it can be re-moved upon reception of the signals.

The measurements of transponder-receivers traveltimes are commonly obtained by dividing the RTT by 2,as suggested in Milne (1983). It is then reasonable to con-sider, under the vehicle stationary assumption during theinterrogation/reply cycle (valid for short interrogation dis-tances and slow vehicle speed such as those considered inthis work) and neglecting the small time difference inducedby the position of the onboard acoustic trigger/pinger rela-tive to the receiving array, that the travel times between thetransponder and the receivers can be computed by remov-ing the known reply delay on the transponder and half ofthe average measured RTT,

tr i = tRTT i − tD − 〈tRTT〉/2,

where 〈tRTT〉 is the average of the RTT given by 〈tRTT〉 =∑nr

i=1 tRTT i . Thus, for the sake of simplicity, the range mea-surements between the transponder and the receivers in-stalled onboard the vehicle (as measured by the USBL de-vice) ρi are assumed to be

ρir = vptr i + ηc + ηd i,

where vp is the underwater sound speed, assumed to beknown and constant for confined mission scenarios, ηc rep-resents the measurement noise induced by the common er-ror to all receivers, and the term ηd i represents the differ-ential noise induced by the additional error sources and the

Figure 3. USBL system reference frames—the body-fixed co-ordinate frame is rigidly attached to the vehicle, while theEarth-fixed reference frame is attached to a fixed point on themission area. The centroid of the onboard receivers serves asthe reference point for the USBL reference frame. The north,east, and down axes in the Earth-fixed reference frame {E} arerepresented, respectively, by the letters n, e, and d.

acoustic quantization performed by the USBL system. Asdepicted in Figure 3, the position of the transponder in thevehicle coordinate frame is given by

r = RT (s − p), (1)

where the matrix R ∈ SO(3) is the shorthand notation forthe body {B} to Earth {E} coordinate frames rotation ma-trix E

BR, the operator (·)T represents the matrix transpose[thus RT ∈ SO(3) represents the inverse rotation matrixfrom {E} to {B}], r ∈ R

3 is the position of the transponderin {B} , s ∈ R

3 is the position of the transponder in Earth-fixed coordinates, and p ∈ R

3 is the position of the vehiclein Earth-fixed coordinates. Let {U} denote the coordinateframe attached to the USBL receiving array, which is cen-tered at the centroid of the receivers such that

nr∑i=1

U bi = 0,

where U bi ∈ R3 denotes the position of the receiver in the

USBL coordinate frame {U} , and nr is the number of in-stalled receivers on the array. The distances between thetransponder and the receivers installed onboard the vehi-cle (as measured by the USBL device) can be written as

ρi = ‖U bi − U r‖, (2)

where U r ∈ R3 is the position of the transponder in {U} .

The installation of the USBL array on the vehicle can bedescribed by a transformation defined in the special Eu-clidean group SE(3) that relates vectors in {U} to vectorsin {B} . Let U x be a vector in {U} , B

U R be the installationrotation matrix between {U} and {B} , and BpU be the in-stallation position offset between {U} and {B} . The repre-sentation of U x in {B} is given by

Bx = BpU + BU RU x,

Journal of Field Robotics DOI 10.1002/rob

Page 5: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

146 • Journal of Field Robotics—2013

Figure 4. Navigation system block diagram—a direct-feedback loop in which an EKF dynamically estimates the INS errors andinertial sensors biases, with the aid of external sensors. The work presented in this paper focuses mostly on the measurementresidual computation block and on the EKF observation models, to provide software-based improvements from the additionalTDOA information, available from the tightly coupled USBL positioning system.

which allows us to write

U x = BU RT (Bx − BpU ). (3)

Applying the frame transformation from Eq. (3), thedistances between the transponder and the receivers inEq. (2) can be simply written in the {B} reference frame as

ρi = ∥∥BU RT (bi − BpU ) − B

U RT (r − BpU )∥∥

= ∥∥BU RT (bi − r)

∥∥ = ‖bi − r‖, (4)

where bi ∈ R3 denotes the position of the receiver in {B} .

Finally, using Eq. (1) in Eq. (4) yields

ρi = ‖s − p − Rbi‖. (5)

3. USBL-AIDED INERTIAL NAVIGATION SYSTEM

This section details the USBL-aided inertial navigation ar-chitecture adopted in this work and presents the noveltightly coupled USBL sensor fusion technique in Sec-tion 3.2.2. The specific USBL sensor-based INS-aiding tech-niques are presented in Section 3.2, and implementation de-tails are provided in Section 3.3. The overall architecture isbriefly outlined here before introducing the EKF modeledinertial error dynamics in Section 3.1. Appendix D providesadditional details on the internal structure of the INS algo-rithm used herein.

The INS is the backbone algorithm that performs atti-tude, velocity, and position numerical integration from rategyro and accelerometer triads data, rigidly mounted on thevehicle structure (strap-down configuration). The nonideal

inertial sensor effects, due to noise and bias, are dynami-cally compensated by the EKF to enhance the navigationsystem’s performance and robustness. Position, velocity, at-titude, and bias compensation errors are estimated by in-troducing the aiding sensors data in the EKF, and are thuscompensated in the INS according to the direct-feedback(Brown and Hwang, 1997) configuration shown in Figure 4.The INS numerical integration algorithms adopted in thiswork are based on the work detailed in Savage (1998a) andSavage (1998b). Applications within the scope of this workare characterized by confined mission scenarios and limitedoperational time, allowing for a simplification of the frameset to Earth and body frames and the use of an invariantgravity model without loss of precision.

3.1. EKF Modeled Inertial Error Dynamics

In a stand-alone INS, bias and inertial sensor error com-pensation is usually performed based on extensive offlinecalibration procedures and data. The usage of filteringtechniques in navigation systems allows for the dynamicestimation of inertial sensor nonidealities, bounding theINS errors. From the myriad of existing filtering techniques,such as particle filters and the unscented Kalman filter(UKF), among others, the EKF is used in this work to es-timate and compensate for the INS errors. The inertial er-ror dynamics, based on perturbational rigid body kinemat-ics, were brought to full detail by Britting (1971) and areapplied to local navigation by modeling the position, ve-locity, attitude, and bias compensation errors dynamics,

Journal of Field Robotics DOI 10.1002/rob

Page 6: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 147

Figure 5. Overview of the sensor information being provided by the USBL to the INS in each of the two alternative sensor-fusionfiltering techniques—the loosely coupled version of the filter provides a position fix computed from the measured range anddirection of the transponder (computed as described in Appendix A), whereas the new tightly coupled technique directly exploitsthe spatial information from the array, and provides the measured ranges from all receivers and all possible combinations of RDOAmeasurements.

respectively,

δp = δv,

δv = −Rδba − (Rar )×δλ + Rna,

δλ = −Rδbω + Rnω,

˙δba = −nba,

˙δbω = −nbω, (6)

where the position and velocity linear errors are defined,respectively, by

δp = p − p, δv = v − v, (7)

the matrix R ∈ SO(3) is the shorthand notation for the body{B} to Earth {E} coordinate frame rotation matrix E

BR, andthe attitude error rotation vector δλ is defined by R(δλ) �RRT and bears a first-order approximation,

R(δλ) � I3 + [δλ×] ⇒ [δλ×] � RRT − I3, (8)

of the direction cosine matrix (DCM) form [see AppendixD for details on the usage of the DCM formulation in theINS, and in particular Eq. (D1)]. In particular, the pro-posed filter underlying the error model (6) includes thesensor’s noise characteristics directly in the covariance ma-trices of the EKF and allows for attitude estimation usingan unconstrained, locally linear, and nonsingular attitudeparametrization. Once computed, the EKF error estimatesare fed into the INS error correction routines as depicted inFigure 4. The attitude estimate, R−

k , is compensated usingthe rotation error matrix R(δλ) definition, which yields

R+k = RT

k (δλk)R−k ,

where RTk (δλk) is parametrized by the rotation vector δλk

[according to Eq. (D1) in Appendix D]. The remaining statevariables are linearly compensated using

p+k = p−

k − δpk, v+k = v−

k − δvk,

b+a k = b−

a k − δba k, b+ω k = b−

ω k − δbω k.

After the error correction procedure is completed, the EKFerror estimates are reset. Therefore, linearization assump-tions are kept valid and the attitude error rotation vectoris stored in the R+

k matrix, preventing attitude error esti-mates to fall in singular configurations. At the start of thenext computation cycle (t = tk+1), the INS attitude and ve-locity/position updates are performed on the corrected es-timates (λ+

k , v+k , p+

k ).

3.2. USBL Sensor-based INS Aiding

To tackle INS error buildup, the EKF relies on observa-tions from external aiding sensors to accurately estimatethe INS errors and correct them by relying on the directfeedback mechanism presented herein. This section intro-duces an external aiding technique based on the ranges andrange-difference-of-arrival (RDOA) measured by a USBL,installed in an inverted configuration onboard the AUV(Vickery, 1998). A more detailed overview of the informa-tion flow, of both the state-of-the-art loosely coupled aidingtechnique and the novel tightly coupled filter, can be seenin the sequel in Figure 5.

Journal of Field Robotics DOI 10.1002/rob

Page 7: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

148 • Journal of Field Robotics—2013

3.2.1. Loosely Coupled USBL/INS

The transponder position fix, as measured by the USBL anddescribed in Appendix A, can be described in body-fixedcoordinates by

Brr = RT (s − p) + nr, (9)

where s is the transponder’s position in an Earth-fixed co-ordinate frame, p is the position of the body-fixed frameorigin in the Earth-fixed frame, and nr represents the rel-ative position measurement noise, characterized by takinginto account the acoustic sensor noises and the USBL posi-tioning system. The estimate of the relative position of thetransponder in the body-fixed frame can be computed us-ing the INS a priori estimates R and p as follows:

B r = RT (s − p).

Using the position error definition (7) and replacing the ro-tation matrix R by the attitude error δλ approximation (8),manipulation of Eq. (9) yields

Brr = B r + RT δp + (RT δλ)×B r + RT (δλ)×δp + nr. (10)

Thus, ignoring the second-order error term δλ × δp and us-ing the properties of the cross product and skew-symmetricmatrices in Eq. (10) yields

Brr = B r + RT δp − (B r)×RT δλ + nr.

The measurement residual used as observation in theEKF is given by the comparison between the measuredtransponder position fix and the estimated transponder po-sition, leading to

δzr = Brr − B r = RT δp − (B r)×RT δλ + nr. (11)

Finally, in order to correctly describe nr, a stochasticlinearization is performed on Eq. (11) (see Appendix B foradditional details on the stochastic linearization performedon the residual position measurement).

3.2.2. Tightly Coupled USBL/INS

Classical loosely coupled strategies rely on position fixescomputed prior to the filtering state, based on the sameset of nonlinear range and RDOA measurements from theUSBL subsystem. In such traditional approaches, the po-sitions of the receivers onboard are not explicitly knownby the filtering architecture. This section describes theproposed tightly coupled technique used to aid the INSwith the USBL sensor information. The tightly coupledUSBL/INS integration strategy directly exploits the acous-tic array spatial information to calculate the distances fromthe transponders to each receiver on the USBL array, and itfeeds this information directly into the EKF.

Using the position and attitude error definitions (7)and (8), respectively, yields for the range measurement ofreceiver i in Eq. (5)

ρi = ‖s − p + δp − Rbi + (δλ)×Rbi‖. (12)

Using the properties of the cross product and skew-symmetric matrices in Eq. (12) yields

ρi = ‖s − p + δp − Rbi − (Rbi)×δλ‖. (13)

To improve performance, the EKF is directly fed with rangemeasurements between the transponder and all receiversonboard, and also the RDOA between all receivers. Alter-natively, the filter may be driven by one range observationand a set of independent RDOA measurements. The sameset of observations that are used by the USBL subsystemto compute transponder position fixes are instead directlyprovided to the tightly coupled filter. Thus, the filter hasdirect knowledge of the receivers’ positions on the localarray and direct access to the raw range and RDOA mea-surements. Ultimately, this direct connection allows the fil-ter to extract better raw and unmodified information fromthe acoustic measurements instead of relying on modifiedor transformed data from the USBL positioning schemes.

3.2.3. Additional Vector Observation

This section introduces an additional vector observation toimprove the overall observability properties of the naviga-tion system. The physical coupling between attitude andvelocity errors, evidenced in Eq. (6), also enables the useof the USBL position fixes to partially estimate attitude er-rors. However, as this physical attachment is invariant inthe body-fixed coordinate frame, the attitude error is notfully observable solely from the rate gyros, accelerometers,and USBL measurements.

As convincingly argued in Goshen-Meskin and Bar-Itzhack (1992) for observability analysis purposes, a GPS-only aided INS with bias estimation can be approximatedby a concatenation of piecewise time-invariant systems,and, under that assumption, full observability is met byperforming specific manoeuvres along the desired trajec-tory. Based on the observability theorem (Rugh, 1996), andas discussed in Morgado et al. (2006), a local weak observ-ability analysis of the system reveals that either stopped oralong a straight line path, full observability is only achievedusing at least three transponders (on a nonsingular geom-etry) or two transponders and a magnetometer. Moreover,along curves, two transponders or one transponder and amagnetometer are sufficient to achieve full observability.Interestingly enough, specific in-flight alignment manoeu-vres, such as transitions between straight paths to curves,excite the nonobservable directions of the system, turn-ing the filter to full observability, as discussed in Goshen-Meskin and Bar-Itzhack (1992). Practical observability inreal-world mission scenarios is nonetheless often achiev-able given that external environmental disturbances (forinstance, underwater currents and the vehicle’s own con-trol action to counteract disturbances) excite some unob-servable directions. However, each specific manoeuvre of-ten includes a subset of unobservable states, and hence

Journal of Field Robotics DOI 10.1002/rob

Page 8: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 149

additional aiding sensors are of interest to improve fullstate estimation. Recent work by Koifman and Bar-Itzhack(1999) and Vasconcelos et al. (2011) has been directedtoward the inclusion of vehicle dynamic information tostrengthen the system observability.

Thus, the required extra attitude measurement can bedrawn from observations of Earth-fixed reference vectors,with onboard sensors that measure the same quantity, inbody-fixed coordinates, as follows:

Bxr = RT Ex + nx,

where Ex is the nominal reference vector in Earth-fixed co-ordinates, assumed known and locally constant, and nx isthe vector measurement noise. The vector-aiding measure-ment residual is computed by comparing this vector obser-vation to the same estimated quantity given the INS a prioriattitude estimate R as

Bzc = Bxr − RT Ex = RT (RRT − I)Ex + nx (14)

in body-fixed coordinates, or in Earth-fixed coordinates asEzc = RBxr − Ex = (RRT − I)Ex + Rnx . (15)

Using the attitude error approximation (8) and the proper-ties of the cross product in Eqs. (14) and (15) yields

Bzc = −RT (Ex)×δλ + nx (16)

in body-fixed coordinates, or in Earth-fixed coordinates asEzc = −(Ex)×δλ + Rnx . (17)

The vector observation model in Eq. (16) or in Eq. (17)can be particularized to any suitable vector sensor, suchas gyroscopic inclinometers and magnetometers. Other ex-amples of vector observations can be found in the low-frequency content of bias-compensated triads of a fiber op-tic rate gyroscope (FOG) that are able to measure Earth’srotation vector, and in image- or sonar-based algorithmsthat extract vector features from the environment. Magne-tometers are typically used in the literature as heading sen-sors providing only one angle measurement, whereas us-ing a triad of magnetometers allows for the exploitation ofall information available from Earth’s magnetic field—i.e.two angles when the local magnetic model includes boththe magnetic declination and inclination. Thus, a triad ofmagnetometers is adopted in this work as a vector obser-vation sensor providing measurements of the Earth’s mag-netic field in body-fixed coordinates,

Bmr = RT Em + nx,

where Em is the nominal local magnetic field vector inEarth-fixed coordinates, assumed known and locally con-stant, and nm is the magnetometer measurement noise. Themagnetometers are considered to be calibrated and com-pensated for bias, scale factors, and nonorthogonality ofthe input axis prior to each mission, using, for instance,attitude-independent batch processing methods (Alonso

and Shuster, 2002). Applying the observation models inEqs. (16) and (17) to the magnetometers triad yields

Bzc = −RT (Em)×δλ + nm (18)

in body-fixed coordinates, or in Earth-fixed coordinates as

Ezc = −(Em)×δλ + Rnm. (19)

3.3. Implementation

This section presents the implementation details of theproposed filters in a stochastic filtering setup, particularlysuited to be implemented in an EKF in a direct-feedbackstructure as presented herein. A comparison of the informa-tion flow of both the state-of-the-art loosely coupled aidingtechnique and the novel tightly coupled filter can be seenin Figure 5.

Without loss of generality, let any continuous-timestate space model be described as{

xc = Fc(xc)xc + Gc(xc)nxc+ uc,

zc = Hc(xc)xc + nzc,

where xc is the state vector, Fc is the state dynamics matrix,nxc

is the state noise, Gc is the matrix that links the statenoise to the state evolution, uc is a vector that representsthe known deterministic inputs, and zc represents the vec-tor measurements that relate to the state vector through thematrix Hc and is disturbed by the observation noise vectornzc

. The state and measurement noises are considered tobe zero-mean, uncorrelated, additive white Gaussian noise(AWGN) processes, and with covariance matrices Qc andRc such that

E{nxc

(t)nxc(τ )T

} = Qc(t)δ(t − τ ),

E{nzc

(t)nzc(τ )T

} = Rc(t)δ(t − τ ). (20)

Specifying for the navigation system at hand, the state vec-tor to be estimated consists of the inertial errors in position,velocity, and attitude, and it also includes the inertial sensorbiases offsets as described in Section 3.1,

xc = δxins = [δpT δvT δλT δba

T δbωT]

T ∈ R15,

and the state model of the first-order INS error Eqs. (6)given by

xc = δxins = Fins(x)δxins + Gins(x)nins,

where nins = [np

T naT nω

T nbaT nbω

T]

T ∈ R15, np ∼ N

(0, �p) is a fictitious white-noise process associated withthe position integration error. This fictitious noise serves toaccount for unmodeled dynamics in the filters, and it alsoacts as a tuning knob on the filter, allowing for the tuningof the frequency response of the filter from the inputsto the position estimate. The matrices Fins ∈ R

15×15 and

Journal of Field Robotics DOI 10.1002/rob

Page 9: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

150 • Journal of Field Robotics—2013

Gins ∈ R15×15 are given by

Fins(x) =

⎡⎢⎢⎢⎢⎢⎢⎣

0 I3 0 0 0

0 0 −(Rar)× −R 0

0 0 0 0 −R0 0 0 0 0

0 0 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎦

,

Gins(x) = diag(I3, R, R,−I3, −I3),

where diag(·, . . . , ·) represents a block diagonal matrix. Thestate noise covariance matrix is given by

Qins = diag(�p, �a, �ω, �ba , �bω) ∈ R

15×15.

3.3.1. Magnetometer Vector Aiding

Although the measurement residuals (18) and (19) describethe same attitude information, the linearized measurementmatrix for Eq. (19) is constant and the components of δλ canbe related directly with those of Em. The observation ma-trix for the magnetometer vector observation is then givenby

Hc mag = [0 0 −(Em)× 03×9

] ∈ R3×15.

The measurement noise covariance matrix comes as

Rc mag = R�mRT ∈ R3×3,

where �m is the covariance matrix of the white Gaussianmagnetometer noise nm ∼ N (0,�m).

3.3.2. Loosely Coupled USBL

The loosely coupled filter uses the USBL position fixes,computed using the planar-wave approximation as de-scribed in Appendix A, which are fed to the EKF to correctthe inertial errors in the INS and estimate the inertial sen-sor biases. The EKF computes the measurement residualsusing Eq. (11), where a stochastic linearization is performedto correctly describe the positioning error nr. Thus, the setof measurements provided to the loosely coupled filter isgiven by

zc USBL = δzr = Brr − B r.

Taking into account Eq. (11), the measurement JacobianHc USBL ∈ R

3×15 is given by

Hc USBL = [RT 03×3 −(B r)×RT 03×3 03×3

].

Based on a stochastic linearization of the USBL position fix,described in Appendix B, the covariance matrix of the ob-servation noise is given by

Rc USBL =⎡⎣Vxx Vxy Vxz

Vyx Vyy Vyz

Vzx Vzy Vzz

⎤⎦ ,

where the matrix elements Vkj with k = x, y, z and j =x, y, z are given by

Vkj = E{(g(ρ, dk) − E{g(ρ, dk)})}{(g(ρ, dj ) − E{g(ρ, dj )})}= dk djE{(ρ − ρ)2} + dkρE{(ρ − ρ)(dj − dj )}

+ dj ρE{(ρ − ρ)(dk − dk)}+ ρ2E{(dk − dk)(dj − dj )}.

3.3.3. Tightly Coupled USBL

The tightly coupled USBL provides measurements of theranges between the transponder and the onboard receivers,which are described using the INS estimates by Eq. (13).The EKF is also fed with the set of RDOA between the re-ceivers with a higher precision. Thus, the set of measure-ments provided to the filter is given by the residuals

zc USBL = [ρ1 · · · ρnm

ρ1 − ρ2 ρ2 − ρ3 · · · ρnr−1 − ρnr

]T ,

where nm is the number of stand-alone measured ranges,either one or the full set of nr measurements (nm ≤ nr ), andthe observation Jacobian is given by

Hc USBL

= [Hu,1

T · · · Hu,nm

T Hu,1T − Hu,2

T · · · Hu,nr−1T − Hu,nr

T]

T ,

where each component Hu,i ∈ R1×15 is given by

Hu,i

=[

(s − p − Rbi)T

‖s − p − Rbi‖01×3

−(s − p − Rbi)T R(bi)×RT

‖s − p − Rbi‖01×3 01×3

],

which is only well-defined for ρi = ‖s − p − Rbi‖ > 0. Themeasurement covariance matrix is defined as

Rc USBL =[σ 2

c 1nm×nm + σ 2d Inm RT

cross

Rcross 2σ 2d Inc

,

],

where nc is the number of combinations used for the setof RDOA measurements, with nc ≤ C

nr

2 = nr (nr−1)/2. Thehigher precision at which the RDOA are measured com-pared to the actual ranges is expressed in the fact thatσ 2

c σ 2d . The covariance matrix Rcross is given by

Rcross =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎣

σ 2d −σ 2

d 0 0 · · · 0

σ 2d 0 −σ 2

d 0 · · · 0...

0 · · · 0 σ 2d 0 −σ 2

d

0 · · · 0 0 σ 2d −σ 2

d

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎦

.

3.3.4. Discretization

The implementation of the state-space model in a discrete-time setting is easily obtained resorting to the classical zero-order hold discretization method described in Gelb (1974),

Journal of Field Robotics DOI 10.1002/rob

Page 10: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 151

in which the inputs are regarded as constant during sam-ple times of the discrete filter. The system state is given bythe INS error model xc = δxins, and the remaining quan-tities are set by nxc = nins, Fc(x) = Fins(x), Gc(x) = Gins(x).The state covariance in Eq. (20) is given by Qc = Qins. Thediscrete state transition matrix is approximated by

�k(tk+1, tk) ≈ eFk (tk+1−tk ),

where Fk = Fc|t=tk (x+k ) is evaluated at the updated state es-

timate x+k , and e(·) represents the matrix exponential. The

measurements are provided to the filter by stacking theobservations from the magnetometer and the USBL wheneach is available. The processing of the EKF allows for ei-ther a batch-processing of the observations from severalasynchronous sensors or from stacking the observationsin the same measurement vector and adjusting the corre-sponding measurement covariance matrices. The state esti-mate is updated using the standard EKF state update equa-tion

x+k = x−

k + Kk

(zk − h

(x−k

)),

where Kk is the Kalman matrix gain, x+k is the updated state

estimate, x−k is the prior state estimate, zk are the aiding sen-

sor observations, and h(x−k

)is the set of predicted obser-

vations, evaluated at the prior state estimate. The Kalmangain and remaining quantities are calculated using the stan-dard EKF update and propagation equations (Gelb, 1974),

Kk = (P−

k HT

k + Ck

)(HkP−

k HTk + Rk + HkCk + Ck

T HT

k

)−1,

P+k = (I − KHk)P−

k − KkCT

k ,

P−k+1 = φkP+

k φTk + Qk,

where P−k is the prior state estimation error covariance ma-

trix, P+k is the posterior corrected state estimation error co-

variance matrix, and Hk is the Jacobian matrix of the non-linear observation equation h

(x−k

)evaluated at the prior

state estimate as presented in the previous sections. Thematrices Rk and Qk are the discrete equivalents of, respec-tively, the observations and state noise covariance matricesfrom sample time tk to tk+1. Under the zero-order hold ap-proximation, Rk and Qk are given by (Gelb, 1974)

Qk �[GkQc|t=tk GT

k

]Ts,

Rk � Rc|t=tk /Ts,

where Gk = Gc|t=tk , and Ts = (tk+1 − tk) is the filter sampletime. As described in Section 3.1, the estimated INS errorsare passed on to the error-correction routines in the INS al-gorithm and reset in the EKF maintaining valid all previouslinearisation assumptions.

4. NUMERICAL SIMULATION RESULTS ANDPERFORMANCE EVALUATION

The overall navigation system performance was assessedin simulation using extensive Monte-Carlo simulations inwhich the filtering setup and all the sensors are exposed todifferent initial conditions and noise sequences. Prior to de-tailing the numerical simulation tests and characteristics, abrief overview of the theoretical bounds used in this workto assess the performance of the proposed techniques ispresented below.

Theoretical performance bounds have long been pur-sued as an important design tool that helps gauge the at-tainable performance by any estimator based on presetconditions of process observations and noise. This kind ofbound also provides an assessment of whether imposedperformance specifications are feasible. A commonly usedlower bound for time-invariant statistical models is theCramer-Rao lower bound, which provides a lower boundon the estimation error of any estimator of an unknownconstant parameter of that particular statistical model. Ananalogous bound for random parameters for nonlinear,nonstationary system models, referred to as the BayesianCramer-Rao lower bound (BCRB), was first derived in VanTrees (1966) and carefully reviewed in Van Trees (1968) andVan Trees and Bell (2007). A discrete-time version, knownas posterior Cramer-Rao lower bound (PCRLB), was intro-duced in Van Trees (1968) and has proven to be a valuableanalysis tool to assess the performance of discrete-time dy-namical estimators. It is also suitable for nonlinear, non-stationary dynamical systems, as is the case of the workpresented herein, and it was recently used for point andextended target tracking (Zhong et al., 2010). The solutionproposed in Van Trees (1968) did not allow, however, for anefficient computation of the bound, and a recursive methodfor an efficient computation of the PCRLB for the discrete-time case was presented in Tichavsky et al. (1998). Readersnot familiar with the theory beyond the PCRB should fol-low the review presented in Tichavsky et al. (1998) and ref-erences therein. An overview of the PCRLB applied to thiswork is provided in Appendix C.

The USBL receiving array has a baseline of approxi-mately 30 cm and is composed of four receivers that are in-stalled in the positions given by b1 = [ 0.2 −0.15 0 ]T m, b2 =[ 0.2 0.15 0 ]T m, b3 = [ 0.4 0 0.15 ]T m, and b4 = [ 0.4 0 −0.15 ]T

m, with respect to the body-fixed coordinate frame {B} ,where the inertial measurement unit (IMU) is also in-stalled in a strap-down configuration. A schematic rep-resentation of the USBL array can be found in Figure 5,in Section 3, and the actual array can be seen in the se-quel in Figure 11(b), in Section 5. The INS provides open-loop integrated estimates of the platform position, veloc-ity, and attitude with a frequency of 50 Hz, and the triadof accelerometers is inspired by a realistic, commerciallyavailable sensor package, the Crossbow® CXL02TG3 tri-axial accelerometer, considered to provide specific force

Journal of Field Robotics DOI 10.1002/rob

Page 11: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

152 • Journal of Field Robotics—2013

Figure 6. Vehicle trajectory.

measurements corrupted by additive uncorrelated, biasedwhite Gaussian noise, with a standard deviation of 0.6 mg(that is, 5.886 × 10−3 m/s2 for a gravity constant of g ≈9.81 m/s2). The rate gyros are also inspired by a realis-tic sensor package, the Silicon Sensing CRS03 triaxial rategyro, and are thus considered to be disturbed by additive,uncorrelated, biased white Gaussian noise, with a standarddeviation of 0.05 deg/s. A magnetometer is also used inthe proposed solution, as described in Section 3.2.3, and in-spired by the commercially available Crossbow® CXM113triaxial magnetometer, which is assumed to be calibratedfor bias, scale factors, and nonorthogonality of the inputaxis, but is disturbed by AWGN with zero-mean and stan-dard deviation of 60 μG. The range measurements betweenthe transponder and the reference receiver (receiver 1) areconsidered to be disturbed by additive, zero-mean whiteGaussian noise, with 0.3 m standard deviation, while theRDOA between receiver 1 and the other three receiversis considered to be measured with an accuracy of 6 mm.The transponder is located in local inertial coordinates ats = [ 0 100 0 ]T m. The vehicle describes the trajectory de-picted in Figure 6, and the inertial sensor biases were mod-eled in simulation as unknown constants different fromzero, although in the filters the corresponding state vari-ables are tuned to be slowly time varying. In practice, theinertial sensor biases are actually slowly time varying withunknown time constants, as will be seen in the experimen-tal results analysis in Section 6.

To correctly compute the PCRLB for the specific nav-igation systems in the analysis herein, the numerical inte-gration algorithms are executed in a direct-feedback settingwithout the EKF in the loop for several Monte-Carlo real-izations of the inertial sensors AWGN disturbances. In eachintegration step, the errors that arise from the inertial algo-

rithms are recorded for posterior evaluation of the PCRLBand reset in the correction routines. Using this setup, theevaluated PCRLB assesses the attainable performance ofany estimator that is placed in the direct-feedback loop. Thebound was computed with several points, in particular for

M={10, 20, 100, 200, 300, 400, 500, 600, 700, 800, 900, 1,000},(21)

with n = 12 elements and for each value i < n in the set,compared against the bound computed with 1,000 points(the last value of M). Thus, the plot in Figure 7 comparesthe norm of the error of computing the bound with the firstn − 1 points of the set M in (21) compared to the boundcomputed with 1,000 points (the last value of M). As illus-trated in Figure 7, the bound computation error changessignificantly when computed with 100 points relatively to10 or 20 points. The error to the bound computed with1,000 points does not change significantly when computedwith either 800 or 900 points, leading to the conclusionthat computing the bound with 900 points is similar tocomputing with 800, which indicates that a tight boundhas been found. The lower bound computed herein, i.e.,the PCRLB, represents the best-case scenario of the attain-able performance by any estimator given the trajectory ofthe vehicle and the available sensors. This relation is wellestablished by the theory behind the PCRLB, briefly de-scribed in Appendix C. Instead of having two separatebounds for each USBL observation model—the loosely cou-pled and the tightly coupled—a single bound that uses thefull nonlinear range equations between the vehicle and thetransponder is considered in the sequel. The reasoning be-hind the usage of this single bound to gauge the attain-able performance of both strategies lies in the fact that theloosely coupled filter uses the same acoustic hydrophonesas the tightly coupled solution to compute the position ofthe transponder. Ultimately, the loosely coupled transpon-der positioning computation applies a transformation tothe underlying raw acoustic data, which is in part respon-sible for the difference of performance between the twostrategies, which is what this study seeks to assess.

The loosely coupled and the tightly coupled fusiontechniques were compared to the PCRLB resorting to N =100 Monte-Carlo runs in which all the sensors are subjectto different noise sequences and the filtering setup is ex-posed to different initial conditions drawn from a normalGaussian distribution with zero mean and standard devi-ations given by 5 m in position error, 0.5 m/s in velocityerror, 1 deg in attitude error; 0.0785 m/s2 in accelerometerbias misalignment, and 10/3 deg/s in rate gyro bias mis-alignment. The mean position error from the Monte-Carloevaluation of both strategies is compared in Figure 8(a)where it can be confirmed that both strategies are unbiasedestimators for position. The performance enhancement ofthe tightly coupled strategy is evident from the root-mean-square (RMS) position estimation error comparison in

Journal of Field Robotics DOI 10.1002/rob

Page 12: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 153

Figure 7. Posterior Cramer-Rao bound Monte-Carlo evaluation with several evaluation points. The plot compares the norm ofthe error of computing the bound with the first n − 1 points of the set M in Eq. (21) compared to the bound computed with 1,000points (the last point of M). The bound computation error changes significantly when computed with 100 points relative to 10 or20 points. The error to the bound computed with 1,000 points does not change significantly when computed with either 800 or 900points, leading to the conclusion that computing the bound with 900 points is similar to computing with 800, which indicates thata tight bound has been found.

Figure 8. Position mean and root-mean-square (RMS) estimation error for tightly coupled vs. loosely coupled comparison withN = 100 Monte-Carlo runs. The performance enhancement in the position estimation accuracy of the tightly coupled strategy isevident from the comparison with the loosely coupled one, where the tightly coupled EKF is shown to operate near the PCRLB asopposed to the loosely coupled EKF.

Figure 8(b). The tightly coupled EKF is shown to clearlyoperate near the PCRLB as opposed to the loosely coupledsolution, with improved estimation accuracy.

The enhancement in position estimation performanceis further evidenced through the analysis of the norm ofthe position error in Figure 9. The remaining quantities (ve-

locity, attitude and inertial sensor biases) are also shownto bear clear improvements with the novel tightly coupledfusion technique when compared to the loosely coupled fu-sion classical strategy. Assuming the ergodicity hypothesis,the steady-state estimation error for the remaining quanti-ties is summarized in Table I for the velocity and attitude

Journal of Field Robotics DOI 10.1002/rob

Page 13: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

154 • Journal of Field Robotics—2013

Figure 9. Root-mean-square (RMS) of the norm of the positionestimation error for tightly coupled vs. loosely coupled com-parison with N = 100 Monte-Carlo runs. Notice the sawtoothpattern that is caused by the difference in update rates betweenthe acoustic corrections at 1 Hz and the open-loop INS numer-ical integration routines at 50 Hz.

Table I. Steady-state velocity and attitude estimation error fortightly coupled vs. loosely coupled comparison with N = 100Monte-Carlo runs.

Velocity error Attitude error‖δv‖ (ms−1) ‖δλ‖ (rad)

Filter coupling Average RMS Average RMS

Loosely coupled 6.6e − 02 0.180 1.4e − 03 5.7e − 05Tightly coupled 3.4e − 02 0.045 7.5e − 04 4.3e − 05

Table II. Steady-state inertial sensors bias estimation error fortightly coupled vs. loosely coupled comparison with N = 100Monte-Carlo runs.

R.G. Bias Accel. Bias‖δbω‖ (rad s−1) ‖δba‖ (ms−2)

Filter coupling Average RMS Average RMS

Loosely coupled 4.5e − 05 2.2e − 08 1.2e − 03 1.3e − 04Tightly coupled 2.7e − 05 1.9e − 08 9.1e − 04 1.4e − 04

estimation errors, and in Table II for the inertial sensor bi-ases estimation errors.

The performance enhancement of the novel tightlycoupled technique to directly merge the range and RDOAobservations of an USBL acoustic positioning system intothe direct-feedback EKF (see Figure 4 in Section 3) isevident from the numerical simulation results presentedherein.

5. EXPERIMENTAL SYSTEM DESIGN

The marine habitat naturally poses a huge challenge forsystems development mainly due to its harsh environ-ment, in which marine robotic vehicles have to withstandhigh pressures. Several capable and high-performancenavigation systems are readily available on the mar-ket, such as the PHINSTM underwater INS, the com-

bined USBL+INS+GPS surface tracking system GAPSTM,and the long-range USBL tracking device POSIDONIATM,all from IXSEA®indexcommercially available products.LinkQuest® also provides lower-performance and lower-cost USBL systems that can also be submerged. Commer-cially available solutions, however, do not often allow di-rect access to the travel times of the acoustic waves on thearray receivers. Moreover, commercial inertial navigationsystems often have their outputs downgraded due to ex-port regulations. The combination of these facts, and theprohibitively high cost of current commercial solutions, sti-fle the design of low-cost marine robotic vehicles.

This section is directed toward the development ofa high-performance, low-cost navigation research systemthat meets the need of providing direct access to thetime-of-arrival (TOA) of the acoustic waves on the arrayreceivers, bearing the valuable knowledge inherent to theassembly and design of such a system. An architecture forthe open prototype is proposed, the acoustic array designis discussed, the inertial sensor package is presented, sup-porting acoustic signals to be used are briefly enumerated,and implementation issues are detailed. As a by-product,the system also includes the design of a transponder thatreplies to the interrogations sent out by the integratedUSBL/INS system. An interesting and similar invertedUSBL design was presented in Jaffre et al. (2005); how-ever, this design does not include the tightly coupled fea-tures presented herein as it requires external attitude mea-surements for global navigation capabilities. An alternativenavigation system design can be found in Yun et al. (1999),in which the vehicle navigates underwater using an INSand surfaces sporadically to get GPS position fixes and cor-rect the errors of the INS. A summary of the design pre-sented herein was previously presented by Morgado et al.(2010).

The proposed USBL/INS hardware and software ar-chitecture consists mainly of two major stand-alone sys-tems: the first is the ensemble between the acoustic arrayand the inertial unit, providing acoustic signal acquisitionand processing. The latter is a transponder that scouts forsignals sent by the USBL array and replies after a previ-ously stipulated elapsed time in order for the array to com-pute RTT to the transponder. In this section, all proposedsystems and signal processing techniques are presented.

5.1. USBL/INS System Overview

The integrated USBL/INS hardware is housed in an alu-minum pressure tube capable of withstanding pressuresup to 600 m (tested in a water pressure chamber). TheUSBL array is built using Bosch-Rexroth® aluminum rodsand connections, which allow for a highly configurable ar-ray structure for optimal design during the evaluation andtesting phases. The array is composed of four preampli-fied HTI® -96-MIN (4) hydrophones placed in a nonplanar

Journal of Field Robotics DOI 10.1002/rob

Page 14: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 155

Figure 10. Integrated USBL/INS system diagram.

configuration (which allows for a three-dimensionaltransponder localization) and is coupled to the aluminumpressure tube using a specially designed coupling devicecarved from Delrin highly resistant polymer plastic. Thepinger that interrogates the transponder is an ITC® 1042and is also attached to the array coupling device.

Installed inside the system tube, a D.SignT® digitalsignal processor (DSP) (D.SignT, 2010) package providesthe main processing power of the system that i) performsacoustic signal detection using high-speed fast Fouriertransforms (FFTs), ii) generates interrogation signals to thetransponder using pulse width modulation (PWM), iii) pro-vides system data logging, and iv) includes an Ethernetinterface to a console computer (which is used only forsystem configuration, system status checks, and data up-load/download). The power is provided by a 3,700 mAh11.1 V lithium polymer (LiPo) battery and a bank of direct-current to direct-current (DC-DC) converters, allowing foran estimated over 4 h system autonomy if used as a stand-alone system. When coupled to an underwater robotic ve-hicle, power can be supplied from the vehicle’s own powerand the Ethernet interface becomes available for data com-munications with the vehicle’s control systems. The mainsystem blocks are depicted in Figure 10. The processor isa Texas Instruments® C6713 floating point DSP and theacoustic signal acquisition is performed by a D.SignT®

ADDA16 card which provides four 16-bit resolution syn-chronous acquisition channels, each with a sampling rateof 250 kHz. This acquisition card is connected to four au-tomatic gain controlled (AGC) signal amplifiers, whosegain can either be set in automatic mode or overridden byan analogue voltage control, from a digital-to-analog con-verter (DAC) also available on the DSP module. The receiv-ing amplifiers are fine-tuned to operate on the band of 20–30 kHz.

To interrogate the transponder, the DSP card gen-erates PWM (with an update rate of 250 kHz and aresolution of 1/200 steps of PWM) through a complex pro-grammable logic device (CPLD) that drives the transmis-

sion power amplifier and an underwater acoustic trans-ducer. The power amplifier and acoustic transducer sys-tem are also fine-tuned to transmit maximum energy on theband of 20–30 kHz. A 16-bit card (also designed in-house)with a Phillips® XAS3 microcontroller and a bank of 12 syn-chronous 24-bit high-performance Sigma-Delta analog-to-digital converters (ADCs) provide the sampling capabili-ties of the IMU. From this ADCs bank, nine of the channelssample the triads of accelerometers and rate gyros that con-stitute the IMU and the magnetometers, whereas two otherchannels provide supply voltage and accelerometer casingtemperature sampling for best performance achievement.Sampling rates of up to 150 Hz can be selected, without lossof performance. A RS-232 serial link, with a baud rate of115,200 bps, is the interface between the DSP module andthe microcontroller.

The IMU and magnetometer are pictured in Figure11, where the triad of single-axis rate gyros can be seenon the left of the aluminum frame, and the accelerome-ter’s triad is housed inside the black casing. This aluminumframe also supports the flux-gate magnetometer triad cardabove the accelerometer casing, as depicted in Figure 11.This sensor suite was previously tested and validated inother operational systems. The calibration of the IMU isalso performed in-house using a high-performance inertialcalibration table. The unit depicted in Figure 11(a) uses aCrossbow® CXL02TG3 triaxial accelerometer, three SiliconSensing® CRS03 rate gyros, and the Crossbow® CXM113triaxial flux-gate magnetometer.

At a high level, the system works as follows: the mi-crocontroller collects data from the IMU and sends them tothe DSP via the RS-232 serial link. The RS-232 interface onthe DSP receives the data and stores them in memory usingdirect memory access (DMA) controllers without interrupt-ing the core processor, which is doing time-critical acous-tic signal processing. The DSP processes these data whenpossible at noncritical time-points. At prespecified instantsof time (e.g., once a second), the DSP sends out a ping tothe transponder and turns on the receiving subsystem to

Journal of Field Robotics DOI 10.1002/rob

Page 15: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

156 • Journal of Field Robotics—2013

Figure 11. Prototype system: IMU, acoustic array and core processing systems.

listen for the responses. After several pinging events and re-sponses, and based on several factors such as vehicle max-imum speed underwater sound speed, and others, the DSPis able to start closing the listening time-windows to im-prove multipath rejection. If responses from the transpon-der eventually get lost, the DSP rolls back to a fully opentime search window until it gets a lock again. The proto-type under development is depicted in Figure 11(b), wherethe acoustic array and IMU core processing systems can beseen attached to one of the covers of the pressure-housingaluminum tube.

The signal detection subsystem operates using a two-level scheme: the first, called raw-detection, is time-criticaland uses fast computations on the input signal to com-ply with the speed at which the signal is updated atthe inputs. This phase completes the processing of onepredefined signal channel while storing all the data forall channels in memory. This raw-detection scheme oper-ates using matched-filters of the expected signal based FFTsand overlap-add convolution mechanisms, as illustrated inFigure 12. Upon a correct detection of the signal in onechannel, the signal is guaranteed by the detection schemeto be fully available on all channels; then, the listening sub-system is turned off and a fine-detection scheme is per-formed on all channels to obtain the TOA of the signal ateach hydrophone.

5.2. Transponder System Overview

The transponder system is simply a subset of the previouslydescribed integrated USBL/INS system in Section 5.1. This

system has to listen for ping requests sent by the USBL/INSsystem and respond to them with a predefined signal after apredefined interval of time. For this purpose, the transpon-der only requires a receiving channel and does not need theIMU and the microcontroller to interface with it.

Thus, the transponder system inherits from the previ-ously described system the following blocks: the DSP withthe acoustic acquisition card and the PWM generator, oneAGC amplifier, the emission power amplifier, the batteryand bank of DC-DC converters, and one acoustic trans-ducer that serves as a receiving hydrophone and as a trans-mitting transducer. Additional electronics are also addedfor coupling the transmitting and receiving circuits to thesame acoustic transducer in order to avoid the appearanceof high transmission voltages at the receiving ADCs whenresponding to the ping requests.

5.3. Acoustic Signaling Techniques Overview

The design of any underwater acoustic ranging system re-quires a measurement of the time-of-flight (TOF) of a signal.As for any coherent detection problem, an accurate TOF es-timate may be obtained by passing the input signal througha matched-filter whose impulse response is a time-reversedreplica of the expected signal (Tolstoy, 1993). Under idealconditions, the filter output is related to the autocorrela-tion function of the received signal. Acoustic signals havebeen used for precise underwater range measurement byTOF of acoustic waves in recent decades (Milne, 1983). His-torically, due to the simplicity of the hardware involved,sinusoidal pulses were the primary choice for underwater

Figure 12. Raw detection scheme.

Journal of Field Robotics DOI 10.1002/rob

Page 16: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 157

Figure 13. Experimental system mounted on a support vessel—The USBL/INS system tube and acoustic array are submergedunder approximately 2 m of water. The platform is allowed to swivel, attached to the end of an aluminum rod to avoid structuralstress, damages, and unwanted vibrations, even though a series of cables are restricting the movement of the submerged system.

range measurements. Recent advances and the availabil-ity of low-cost, high-speed DSP hardware and software,amplifiers, and wide band acoustic transducers have al-lowed for the use of more advanced signaling techniques,such as chirp tone bursts and spread-spectrum signals(Austin, 1994; Bingham et al., 2007).

Specially designed spread-spectrum modulated sig-nals have interesting autocorrelation properties (Sarwateand Pursley, 1980) allowing for a narrower output peak ofthe matched-filter and improving the performance of thedetector. In practice, the output peak of the matched-filtersspreads in time, mainly due to unequalized distortions andnonideal conditions, thus degrading the performance ofsuch detectors. In general, spread-spectrum signals haveseveral advantages when compared to conventional signal-ing for underwater range estimation: they present a bet-ter signal-to-noise ratio (SNR), robustness to ambient andjamming noise, multiuser capabilities, improved detectionjitter, and the ability to better resolve multipath, which isone of the biggest problems in underwater channel acous-tic propagation. These specially designed signals are typi-cally generated using either a frequency hopping spread-spectrum sequence (FHSS) or direct spread-spectrum se-quence (DSSS) codes. In the scope of the work presentedherein, the attention is focused on DSSS modulated signals.Closely related work can be found in Austin (1994) andBingham et al. (2007).

The acoustic signaling techniques employed in the ex-perimental setup were directly inherited from the contri-butions presented by Morgado et al. (2011), where a novelmethodology for the design and implementation of trans-mission pulse-shaping filters was proposed. The new de-sign concept is based on closed-loop control strategies withpreview information, and it was shown to be able to im-prove the detection of the transmitted signals by directlymodifying the transmission pulse-shaping filters. The per-formance enhancement was evident from the presented

simulation results and was validated using real data fromexperimental sea trials.

6. EXPERIMENTAL VALIDATION

Experimental trials were conducted at sea in Sesimbra,Portugal, to assess the feasibility of the tightly coupledUSBL/INS system under development. In these tests, thealuminum pressure tube with the attached array was se-curely suspended on the side of a surface vessel, as de-picted in Figure 13, at a depth of approximately 2 m, whilethe transponder was installed on a kayak with the trans-ducer, also submerged approximately 2 m deep. To ob-tain valid evaluation datasets, both the acquisition andthe transmission of the acoustic signals were synchronizedwith the GPS one-pulse-per-second (1-PPS) clock signal. Atotal of five deployments of the navigation system, labeleddeployments #1, #2, #3, #4, and #5, are reported in this sec-tion. All the deployments were recorded using real-timekinematics (RTK) GPS receivers with a precision better than4 cm.

To assess the navigation system capabilities on typicalvehicle manoeuvres, the support vessel with the USBL/INSattached follows a lawn mower trajectory on deploy-ment #1, as depicted in Figure 14(a), moving toward thetransponder, which was installed on a kayak and mooredto two buoys maintaining a steady position throughoutthe entire trajectory. Several key parameters, such as thenominal magnetic field vector and the local gravity vec-tor, were evaluated prior to the mission using the coordi-nates of a reference point near the operation area. Fromthis point forward, this reference point also serves as theorigin of the north-east-down (NED) coordinates posi-tion representation on the local tangent plane (LTP) refer-ence frame. Thus, using magnetic charts (Macmillan et al.,2004), the local magnetic field vector is given by {E} m =[ 0.2645 −0.0149 0.3464 ]T G, which means a declination angle of

Journal of Field Robotics DOI 10.1002/rob

Page 17: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

158 • Journal of Field Robotics—2013

Figure 14. Support vessel trajectory on deployment #1 withthe USBL/INS filter estimates, USBL fixes, and transponderposition. The transponder position was calibrated with theonboard RTK GPS prior to the beginning of the trajectory.The loosely coupled and tightly coupled filter estimates arealso plotted together with the resulting USBL position fixesin north-east-down (NED) coordinates represented on the lo-cal tangent plane (LTP) reference frame. The performance en-hancement is not immediately discernible in this figure—seeFigure 18(b) for a better illustration of the performance en-hancement on this deployment.

approximately −3.22 deg and a dip angle of about 52.6 deg.Using local gravity charts (Imagery and Agency, 2000), thelocal gravity constant is given by g = 9.8002 m/s2. TheUSBL receivers were placed on the array at the locationsgiven by

{C}b1 = [0.1790 −0.15 0]T m,

{C}b2 = [0.3885 0 −0.15]T m,

{C}b3 = [0.1790 0.15 0]T m,

{C}b4 = [0.3885 0 0.15]T m,

relative to a reference frame {C} located on the cover of thealuminum housing cylinder. The origin of the {C} referenceframe is located 0.3 m from the body-fixed reference frame

along the x axis, that is, BpC = [ 0.3 0 0 ]T m. None of thesequantities was actually calibrated using specific calibrationprocedures, but rather they were measured mechanically.The underwater sound velocity was measured at the oper-ation depth using a sound velocity profiler (SVP) and set ata constant value of 1515 m/s.

The inertial sensor experimental noise characteristicswere evaluated using benchmark datasets obtained withthe platform stationary after the internal temperature of theaccelerometers was stabilized. The accelerometer and rategyro noise variances were found to be very close to thoseprovided in each respective technical data sheet, whereasthe magnetometer noise needed to be adjusted from themagnitude of 60 μG indicated on the data sheet to a moreexperimentally suitable 0.6 mG (evaluated with the plat-form stationary for several minutes before starting the tra-jectory). In fact, the temperature of the accelerometers is acritical factor in the stability of the overall system estimatesas the accelerometer biases are known to vary with tem-perature. Bearing in mind the low-cost and complexity ofthese types of units, no steps toward temperature compen-sation of the accelerometers were taken whatsoever. The fil-ter structure must tackle this issue by modeling the biasesas slowly time-varying to accommodate such variations.The magnetometers were calibrated prior to the mission us-ing a batch-processing method known as the TWOSTEP al-gorithm (Alonso and Shuster, 2002).

The acquired USBL data for the trajectory in deploy-ment #1 are represented in the USBL reference frame {U}in Figure 15. Immediately after the system start-up, acousticoutliers are evident, suggesting the need to have an efficientcausal outlier detection scheme, as the EKF is highly sensi-tive to outliers present in the input data. Although other so-lutions could be devised in place of the EKF, such as outlier-robust Kalman filters (Gandhi and Mili, 2010), the solutionimplemented in this work was based on prefiltering theacoustic data with an explicit outlier detection causal fil-ter prior to feeding the ranging and RDOA information tothe EKF. The filter used in this work is the median abso-lute deviation (MAD) outlier detection and rejection filter(Menold et al., 1999), a causal filter based on the medianof a running window with nine (9) acoustic samples. Thisfilter does not change the input data nor does it introducephase, only acting as an active valid/outlier data classifier.The behavior of the outlier classifier is shown to correctlyidentify the outliers present in Figure 15.

To experimentally evaluate the time-difference-of-arrival (TDOA) noise on the USBL acoustic system, themeasured TDOA raw data were smoothed using a sym-metrical running average window with a width equiva-lent to five (5) acoustic samples, and they were then com-pared to the actual raw data. The difference of arrival interms of taps or samples, where one tap is equivalent toan acoustic sample time of 4 μs = (250 kHz)−1, is pre-sented in Figure 16. Ignoring the obvious outliers in the

Journal of Field Robotics DOI 10.1002/rob

Page 18: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 159

Figure 15. USBL processed position fixes on deployment #1—440 fixes represented from a total of 451 detected receptions(97.56% validated). The remaining detected acoustic receptionsviolated physical constraints of the receiving array such as themaximum delay taps between any two hydrophones. A total of14 outliers (3.2%) were correctly classified as outliers. The vali-dation data are obtained from a smoothed estimate of the IMUtilt angles and calibrated magnetometer heading, i.e., withoutremoval of accelerometer biases, thus the validation data arebiased by a tilt error from the accelerometer biases and serveonly to evaluate the alignment of the sensor frames and dy-namic behavior of the sensor data.

beginning of the comparison, the standard deviation be-tween the smoothed and raw data is actually less thanone tap, which was considered in the numerical simulationresults presented in this work. Although not representedhere, the range between the two stations was found to beless than 0.2 m during the entire trajectory.

To avoid the initial gap in acoustic data around the100 s mark, the navigation phase is set to start at t = 130s, allowing the filter to properly align and converge with-out acoustic data outages. The behavior of the system un-der USBL data outage will be discussed and analyzed next.The filters are initialized with the mean values of the GPSrecorded position from the five seconds prior to the begin-ning of the navigation phase. The velocity estimate and allthe inertial sensor bias estimates are set to zero. The filtersestimates and USBL position fixes in NED coordinates inthe local tangent plane (LTP) reference frame are also rep-resented in Figure 14, where the performance enhancementis not immediately discerned. A thorough analysis is re-quired to further assess the performance and compare bothstrategies experimentally. In fact, as the platform was swiv-elling below the supporting vessel, there were no groundtruth data readily available in the LTP reference frame. Toevaluate the overall performance of the navigation filters, aperformance comparison will be provided using the USBL

smoothed raw data as ground truth in the reference frameof the USBL by converting the estimated position in theLTP reference frame to the USBL reference frame, and alsoby using the integrated platform attitude estimates. Theattitude estimates from both the loosely coupled and thetightly coupled filters are shown to be very similar in Fig-ure 17.

The performance of both filters is thus evaluated inUSBL coordinates, which reflects an overall assessmentfrom both the estimated position and the attitude. This al-lows the results to be further compared with smoothed datafrom the USBL raw position fixes, which are unaffectedby the platform tilt estimation errors (potentially inheritedfrom estimation errors of the accelerometer biases). Bothfilter position estimates converted to the USBL referenceframe are plotted in Figure 18(a). Smoothing out the rawUSBL position fixes with symmetric running average filters(with a width equivalent to five acoustic samples) allowsfor the error comparison in Figure 18(b). The same evalua-tion methodology is applied to deployment #2 reported inFigure 18(d) and to deployments #3, #4, and #5 reportedin Figure 19. Although the performance improvement indeployments #1 and #4 is not as clearly evidenced by thesimulation results as the other deployments, the tightlycoupled filter is shown to perform slightly better than theloosely coupled version. Nonetheless, the performance im-provement is more evident from the comparison for de-ployment #2 in Figure 18(d), for deployment #3 in Figure19(b), and for deployment #5 in Figure 19(f), and also fromthe overall position RMS estimation error comparison sum-marized in Table III. Computing the improvement level ofthe tightly coupled filter in the five deployments reportedin Table III yields the median improvement level of approx-imately 15% of the tightly coupled filter against the looselycoupled version.

Ultimately, the performance of the overall system isclosely related to the capabilities of such a system to es-timate the inertial sensor biases. In fact, one of the waysof providing evidence of the performance enhancementof the proposed tightly coupled strategy is by showingthe overall improvement of its bias estimation by forcingan outage of the USBL measurements and observing howboth filters cope with the absence of data for long peri-ods. Thus, a USBL outage was forced for 30 s, in the in-terval [340, 370] s, and the resulting filter output in posi-tion is plotted in Figure 20, where the tightly coupled filteris shown to diverge significantly less than the loosely cou-pled filter. The difference in accelerometer bias estimationbetween both strategies is not as easily discernible by directcomparison as when comparing the behavior during acous-tic data outages. Nonetheless, the estimates of the rate gy-ros and accelerometer biases are shown in Figure 21, whichdemonstrates their slowly time varying behavior.

Improved rate gyros bias estimates are also demon-strated through an open-loop INS numerical integration

Journal of Field Robotics DOI 10.1002/rob

Page 19: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

160 • Journal of Field Robotics—2013

Figure 16. Experimental evaluation of USBL noise on deployment #1 in number of taps in TDOA computation—The evaluationdata represent the error in the TDOA computation in number of taps (where one tap corresponds to an acoustic sample time) whencomparing the raw TDOA data to a 5-s window running-average filter. The computed 1σ standard deviation for each TDOA iscalculated ignoring outliers.

Figure 17. Estimated attitude on deployment #1—as the plat-form was swivelling below the support vessel, there were noground truth data readily available for attitude evaluation pur-poses. Nonetheless, an overall performance comparison is per-formed in the sequel.

assessment of the bias-compensated rate gyros measure-ments in Figure 22. The estimated rate gyros biases fromboth strategies are used to correct the rate gyros measure-ments, which are then fed to the INS attitude numericalintegration schemes in an open loop for approximately6.5 min, starting at 50 s into deployment #4. These open-loop integration results are compared in Figure 22(b), re-vealing a smaller drift for the novel tightly coupled tech-nique, which ultimately demonstrates improved rate gyrosbias estimates and enhanced overall attitude estimation.

Another case that demonstrates enhanced perfor-mance in practice is the faster and improved convergenceto the nominal trajectory from multiple initial conditionson deployment #5, reported in Figure 23. An interesting ad-vantage of the tightly coupled filtering technique can alsobe found in the handling of outliers. Acoustic outliers aretypically caused by incorrect detection in specific combi-nations of acoustic receivers, not necessarily in all possiblepairs of receivers. In such situations, the tightly coupled fil-ter has the advantage of being able to use only a subset ofthe TDOA measurements, as opposed to the loosely cou-pled solution, in which an error in a single TDOA measure-ment causes immediate outliers in the position fix. Such an

Journal of Field Robotics DOI 10.1002/rob

Page 20: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 161

Figure 18. INS estimates on deployments #1 and #2 converted to the USBL reference frame—the position estimates of theUSBL/INS system are converted to the USBL reference frame using the estimated attitude. This allows for an overall perfor-mance assessment by directly comparing the filter outputs to the acoustic positioning data. The error is obtained from comparingthe INS estimates converted to the USBL reference frame, with a 5-s window running-average filter smoothed USBL trajectory.

example from deployment #3 is illustrated in Figure 24, inwhich the pair δ2−4 in the first two outliers, and pairs δ2−3and δ3−4 in the first outlier, can still be used by the tightlycoupled filter to provide corrections to the INS, whereas theloosely coupled filter is not able to use this information.

7. CONCLUSIONS

This paper presented a new USBL tightly coupled inte-gration technique to enhance error estimation in low-cost

strap-down INSs with application to underwater vehicles.In the proposed tightly coupled strategy, the acoustic ar-ray spatial information is directly exploited, resorting tothe extended Kalman filter implemented in a direct feed-back configuration. Classical loosely coupled techniquesuse the position fix of transponders in known positions ofthe mission area, computed from the range and bearing an-gles of the waves arriving at the onboard acoustic array.The planar wave approximation of the arriving acousticsignals is often employed to compute the position fix. The

Journal of Field Robotics DOI 10.1002/rob

Page 21: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

162 • Journal of Field Robotics—2013

Figure 19. INS estimates on deployments #3 to #5 converted to the USBL reference frame—the position estimates of the USBL/INSsystem are converted to the USBL reference frame using the estimated attitude. This allows for an overall performance assessmentby directly comparing the filter outputs to the acoustic positioning data.

Journal of Field Robotics DOI 10.1002/rob

Page 22: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 163

Table III. Position RMS estimation error performance im-provement overview on all deployments of the navigation fil-ters. The median improvement level is around 15% for thetightly coupled strategy.

Position RMS estimation error (m)

Deployment #1 #2 #3 #4 #5

Loosely coupled 1.3405 1.3175 2.3970 2.7052 2.0625Tightly coupled 1.3314 1.0359 1.8871 2.4812 1.7482Improvement level 0.7% 21.4% 21.3% 8.3% 15.3%

Figure 20. Inertial error evaluation during enforced USBLsensor outage on deployment #1—in order to evaluate the be-havior of the system during USBL sensor outages, the USBLwas shutdown for 30 s in the final straight path of the trajec-tory.

proposed enhancement technique makes use of the full in-formation available from the USBL device, such as the TOAand TDOA or RDOA of the signals arriving at the differenthydrophones. The physical coupling between attitude andvelocity errors, evidenced in Eq. (6), also enables the useof the USBL position fixes to partially estimate attitude er-rors. However, as this physical attachment is invariant inthe body-fixed coordinate frame, the attitude error is not

fully observable solely from the rate gyros, accelerometers,and USBL measurements. Thus, an additional source of at-titude information was introduced, drawn from the obser-vations of the Earth’s magnetic field, provided by an on-board magnetometer.

The overall navigation system performance was as-sessed in simulation using extensive Monte-Carlo simula-tions in which the filtering setup was exposed to differ-ent initial conditions and all the sensors to different noisesequences. The performance enhancement of this noveltightly coupled technique is evident from the numericalsimulation results presented in Section 4, which comparethe estimation performance of the new strategy to classicalmerging strategies based on loosely-coupled techniques.Moreover, the tightly coupled EKF was shown to be able tooperate near the PCRLB performance lower bound, whichwas not attainable by the loosely coupled EKF. The pro-posed tightly coupled technique was further validated inan experimental setup with data acquired at sea in a to-tal of five deployments. The inertial platform and acous-tic positioning system that compose the USBL/INS systemwere entirely designed and built in-house, allowing the fil-tering architecture to have direct access to the RTT andTDOA of the acoustic waves arriving at the hydrophoneson the USBL array, which is not typically provided bycommercially available systems. In addition to validatingthe proposed technique, the experimental data obtained atsea allowed for the performance enhancement to be evi-denced through real data. Although the improvement intwo of the deployments was not as clear as was indicatedby the simulation results and the remaining deployments,the improvement was consistent in the majority of the de-ployments. The median improvement level in position wasdemonstrated to be approximately 15%. Nonetheless, byforcing USBL sensor outages during the operation of thefilters, the new tightly coupled technique was revealed tobe more effective in estimating the accelerometers biases,which was reflected by the smaller position error drift dur-ing the enforced acoustic data outage. The rate gyros biaseswere also shown to be better estimated using the tightlycoupled strategy, by comparing the attitude estimates drifterror from open-loop integration of the bias-compensated

Figure 21. Inertial sensor biases estimates on deployment #1—the inertial sensor estimates are shown to converge to slowly time-varying quantities. The tightly coupled (TC) estimates are overlaid over the loosely coupled (LC) biases estimates with the sameline pattern and half line thickness.

Journal of Field Robotics DOI 10.1002/rob

Page 23: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

164 • Journal of Field Robotics—2013

Figure 22. Open-loop attitude integration assessment—the bias estimates from both filtering strategies are used to correct the rategyros measurements and numerically integrate using the INS attitude algorithms in an open loop for approximately 6.5 min, start-ing at 50 s into deployment #4. The smaller attitude open-loop integration drift of the tightly coupled compensated measurementsdemonstrates improved rate gyros bias estimates and overall enhanced attitude estimation performance.

Figure 23. Convergence from multiple initial position es-timates of both strategies on deployment #5 in Earth-fixedcoordinates—the convergence of the tightly coupled filter iscompared to the loosely coupled filter from multiple initial po-sitions, where it can be seen that the newly proposed techniqueexhibits faster and improved convergence to the nominal tra-jectory.

angular velocity measurements. Other improvements in-clude enhanced convergence to the nominal trajectory frommultiple initial conditions, and the ability to use a subset ofTDOA measurements in the tightly coupled sensor space,

whereas the loosely coupled sensor space is unable to cor-rect the INS due to detected position measurement outliers.

ACKNOWLEDGMENTS

This work was supported by FCT (ISR/IST plurianualfunding) through the PIDDAC Program funds, and par-tially funded by the project FCT PTDC/EEA-CRO/111197/2009 - MAST/AM of the FCT and by the EU ProjectTRIDENT (Contract No. 248497). The work of M. Mor-gado was supported by a Ph.D. Student Scholarship withreference SFRH/BD/25368/2005, from the Fundacao paraa Ciencia e a Tecnologia. The authors also gratefullyacknowledge the contribution of Manuel Rufino, BrunoCardeira, Luis Sebastiao, Joao Botelho, Pedro Serra, AndreOliveira, Joao Oliveira, Pedro Batista, and Tiago Gasparthroughout the development of the prototype and sea trials.

APPENDIX A: USBL POSITION FIX COMPUTATION

The USBL system computes the range and direction ofthe transponders using the planar approximation of theacoustic waves as in the classical approach presented inYli-Hietanen et al. (1996).

Using the planar-wave approximation, illustrated inFigure 25, it can be written for any RDOA,

ρi − ρj = −(bi − bj )T d.

In matrix form, for all possible RDOA measurements comes

Ad = −,

Journal of Field Robotics DOI 10.1002/rob

Page 24: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 165

Figure 24. Measured time-difference-of-arrival (TDOA) on deployment #3—three incorrect measurements were detected by theoutlier detection algorithm designed in the Cartesian sensor-space of the loosely coupled solution. These outliers are caused byincorrect detection in specific combinations of acoustic receivers, not necessarily in all possible pairs of receivers, as is the case ofpair δ2−4 in the first two outliers, and pairs δ2−3 and δ3−4 in the first outlier. In such situations, the tightly coupled filter has theadvantage of being able to use only a subset of the TDOA measurements, as opposed to the loosely coupled solution, in which anerror in a single TDOA measurement causes outliers in the position fix.

where

A =

⎡⎢⎢⎢⎢⎢⎣

(b1 − b2)T

(b1 − b3)T

...(bnr −2 − bnr )T

(bnr −1 − bnr )T

⎤⎥⎥⎥⎥⎥⎦ , =

⎡⎢⎢⎢⎢⎢⎣

ρ1 − ρ2

ρ1 − ρ3

...ρnr −2 − ρnr

ρnr −1 − ρnr

⎤⎥⎥⎥⎥⎥⎦ .

To minimize the total estimation error

J = (Ad + )T W(Ad + ),

where W is a weighting matrix, the unique solution for d isgiven by the well known weighted least-squares solution,

d = −A#W, (A1)

where A#W is the weighted pseudoinverse matrix of A given

by

A#W = (AT WA)−1AT W,

and W is given by the inverse of the covariance matrixof .

Figure 25. Planar wave approximation.

If, as assumed in this work, there is only access to in-dependent measurements of a subset of , the directiond is computed using only the subset of equations corre-sponding to those measurements. Given that the originof the USBL frame {U} is located at the centroid of the

Journal of Field Robotics DOI 10.1002/rob

Page 25: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

166 • Journal of Field Robotics—2013

receivers, the range of the transponder can be computedby averaging the measured ranges from all receivers. Therelative position of a transponder expressed in {U} is thencomputed by

rr = ρrdr , (A2)

and it can be described in body-fixed coordinates by

Brr = RT (s − p) + nr,

where s is the transponder’s position in the Earth-fixed co-ordinate frame, p is the position of the body-fixed frameorigin in the Earth-fixed frame, and nr represents the rel-ative position measurement noise, characterized by takinginto account the acoustic sensor noises and the USBL posi-tioning system (A2).

APPENDIX B: USBL POSITION FIX STOCHASTICLINEARIZATION

The range measurements provided by the USBL system, asdescribed in Section 3.2, are given by

ρir = vptr i + ηc + ηd i ,

where vp is the underwater sound velocity, ηc representsthe measurement noise induced by the common error to allreceivers, and the term ηd i represents the differential noiseinduced by the additional error sources and the acousticquantization performed by the USBL system, or equiva-lently in matrix form,

ρr = vptr + ηc1nr ×1 + ηd,

where ηc is a scalar random variable considered to bedrawn from a zero mean Gaussian distribution, that is,ηc ∼ N (0, σ 2

c ), and ηd is an nr × 1 vector random variabledrawn from a zero mean Gaussian distribution with ηd ∼

N (0nr ×1, σ2d Inr ). The random variables ηc and ηd are also as-

sumed to be uncorrelated, that is,

E{ηdηc} = 0,

where E{·} represents the expected value operator.Thus, the covariance matrix for the measured ranges

ρr is easily given by

E{(ρr − E{ρr })(ρr − E{ρr })T } = σ 2c 1nr ×nr + σ 2

d Inr .

The relationship between the RDOA observation vec-tor and the measured distances ρr is established as

= Cρr ,

where C is a combination matrix that encodes the combi-nations between the receivers to generate the RDOA vector

and is given by

C =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

1 −1 0 0 · · · 0

1 0 −1 0 · · · 0...

0 · · · 0 1 0 −1

0 · · · 0 0 1 −1

⎤⎥⎥⎥⎥⎥⎥⎥⎦

.

Taking into account the least-squares solution for thetransponder direction (A1), the transponder direction co-variance is given by

E{(d − E{d})(d − E{d})T }= A#

WC(σ 2

c 1nr ×nr + σ 2d Inr

)CT A# T

W .

The range from the transponder to the centroid of theUSBL array is computed from averaging the range mea-surements of all receivers, and its covariance is given by

E{(ρ − E{ρ})2} = 1n2

r

11×nr

(σ 2

c 1nr ×nr + σ 2d Inr

)1T

1×nr.

The covariance between the measured range ρr and thedirection dr becomes

E{(ρ − E{ρ})(d − E{d})T }

= − 1nr

11×nr

(σ 2

c 1nr ×nr + σ 2d Inr

)CT A# T

W .

The position measurement (A2) can be separated intoits three Cartesian components as

rr = [ρdx ρdy ρdz

]T. (B1)

Let the scalar function g :{R

+; [0, 1]} → R be given by

g(ρ, dk) = ρdk, k = {x, y, z} ,

which allows Eq. (B1) to be written as

rr = [g(ρ, dx ) g(ρ, dy ) g(ρ, dz)

]T. (B2)

Using a Taylor series expansion of Eq. (B2) around thenominal values of ρ and dk yields

g(ρ, dk) = g(ρ, dk) + ∇g(ρ, dk)[

ρ − ρ

dk − dk

]

+12

[ρ − ρ dk − dk

]∇2g(ρ, dk)[

ρ − ρ

dk − dk

]

+ h.o.t. (B3)

The gradient ∇g(ρ, dk) in Eq. (B3) is given by

∇g(ρ, dk)∣∣∣ ρ=ρ

dk=dk

=[

∂g(ρ, dk)∂ρ

∂g(ρ, dk)∂dk

]= [

dk ρ].

Journal of Field Robotics DOI 10.1002/rob

Page 26: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 167

The second-order derivative ∇2g(ρ, dk) is given by

∇2g(ρ, dk)∣∣∣ ρ=ρ

dk=dk

=

⎡⎢⎢⎢⎣

∂2g(ρ, dk)∂2ρ

∂2g(ρ, dk)∂ρ∂dk

∂2g(ρ, dk)∂dk∂ρ

∂2g(ρ, dk)∂2dk

⎤⎥⎥⎥⎦ =

[0 11 0

].

The higher-order terms (h.o.t.) in Eq. (B3) are all nullbecause all derivatives of g(ρ, dk) over the second order arealso null. Thus, the expected value of g(ρ, dk) is given by(Papoulis, 1984)

E{g(ρ, dk)} =∞∫

−∞

∞∫−∞

g(ρ, dk)f (ρ, dk)∂ρ∂dk, (B4)

where f (ρ, dk) is the joint probability density function of ρ

and dk , which verifies (Papoulis, 1984)

∞∫−∞

∞∫−∞

f (ρ, dk)∂ρ∂dk = 1.

Combining Eqs. (B3) and (B4) yields

E{g(ρ, dk)} = ρdk + E{(ρ − ρ)(dk − dk)}.Thus, the expected value of the transponder relative

position is given by

E{r} =

⎡⎢⎣

E{g(ρ, dx )}E{g(ρ, dy )}E{g(ρ, dz)}

⎤⎥⎦ ,

and its covariance can be written as

E{(r − E{r})(r − E{r})T } =

⎡⎢⎣

Vxx Vxy Vxz

Vyx Vyy Vyz

Vzx Vzy Vzz

⎤⎥⎦ ,

where the matrix elements Vkj with k = x, y, z and j =x, y, z are given by

Vkj = E{(g(ρ, dk) − E{g(ρ, dk)})}{(g(ρ, dj ) − E{g(ρ, dj )})}= dk djE{(ρ − ρ)2} + dk ρE{(ρ − ρ)(dj − dj )}

+ dj ρE{(ρ − ρ)(dk − dk)}+ ρ2E{(dk − dk)(dj − dj )}.

APPENDIX C: PERFORMANCE BOUNDS

Consider the general form for the process and observationmodels,

xk+1 = fk(xk, wk),

zk = hk(xk, vk), (C1)

where xk is the system state at sample time k, {zk} is theset of available measurements, {wk} and {vk} are indepen-

dent white processes, and fk and hk are possibly nonlinearnonstationary functions.

Let xk be any estimate of the true state vector of theprocess given by Eqs. (C1). Since we are interested in theclass of unbiased estimators for the state vector, we havethe following inequality for the covariance of the estimationerror:

Pk = E{[xk − xk][xk − xk]T } ≥ J−1k ,

where Jk is the posterior Fisher information matrix (FIM)defined as

Jk = E{ − ∇xk

∇Txk

log p(xk, zk)},

where p(xk, zk) is the joint probability density function ofthe state vector and observations throughout the full ex-tent of the trajectory of the underlying dynamical system.Tichavsky et al. (1998) showed that the FIM Jk can be effi-ciently computed using the following recursion:

Jk+1 = D22k − D21

k

(Jk + D11

k

)−1D12k ,

where

D11k = Ep(xk+1|zk+1)

{−∇xk

∇Txk

log p(xk+1|xk)}

,

D12k = Ep(xk+1|zk+1)

{−∇xk

∇Txk+1

log p(xk+1|xk)}

=[D21

k

]T,

D22k = Ep(xk+1|zk+1)

{−∇xk+1∇T

xk+1log p(xk+1|xk)

}+ Ep(xk+1|zk+1)

{−∇xk+1∇T

xk+1log p(zk+1|xk+1)

}, (C2)

and the recursion is initialized with

J0 = E{−∇x0∇T

x0log p(x0)

}.

To compute the terms in Eqs. (C2), the expected valueoperator Ep(xk+1|zk+1){· · · } needs to be evaluated. The com-putational complexity of these expectations depends en-tirely on the structure of the underlying process and ob-servation models, and involves solving integral terms that,in general, do not have closed-form solutions in the partic-ular case of the estimators presented herein. As suggestedin Simandl et al. (2001), the terms can be estimated usingMonte-Carlo simulations by replacing the expected valueswith the sample mean of the realizations, e.g., the term D11

k

in Eqs. (C2) is computed as

D11k = 1

M

M∑j=1

−∇xk∇T

xklog p(xk+1|xk)

∣∣xk=xk (j ) ,

where {xk(j )}Nk=0 is the j th realization of the state trajectory,j = 1, 2, . . . , M , and M is the number of Monte-Carlo real-izations.

As expected, as M increases, the quality of theMonte-Carlo estimates improves, however there is norule of thumb on selecting an M that guarantees thatthe estimates will be satisfactory. One possible rule is to

Journal of Field Robotics DOI 10.1002/rob

Page 27: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

168 • Journal of Field Robotics—2013

try several values for M and test if the new bound withM2 realizations has changed significantly compared to thebound with M1 < M2 realizations.

In the scope of the work presented herein, we restrictthe PCRLB evaluation to the discrete-time AWGN case forwhich the process and observation models are given by

xk+1 = fk (xk) + wk,

zk = hk (xk) + vk,

where xk now represents the discretized errors of the navi-gation systems at sample time k, {wk} is the process equiv-alent AWGN, {vk} is the measurement equivalent AWGN,{zk} is the set of available measurements that are related tothe state vector by the nonlinear nonstationary observationfunction hk , and fk models the discretized navigation sys-tem error model.

In this framework, the logarithmic terms of the PCRLBrecursion can be written as

log p(xk+1, xk) = c1 − 12

[xk+1 − fk (xk)]T Q−1k

× [xk+1 − fk(xk)],

log p(zk+1, xk+1) = c2 − 12

[zk+1 − hk+1 (xk+1)]T R−1k+1

× [zk+1 − hk+1(xk+1)],

where c1 and c2 are constants, Qk is the discrete equivalentprocess AWGN covariance matrix, and Rk is the discreteequivalent observation AWGN covariance matrix.

Thus, it can be easily derived for the terms (C2) that

D11k = E

{FT

k (xk)Q−1k Fk(xk)

},

D12k = −E

{FT

k (xk)}

Q−1k ,

D22k = Q−1

k + E{

HTk+1(xk+1)R−1

k+1Hk+1(xk+1)}

,

where

Fk(xk) = [∇xkfTk

]T,

Hk+1(xk+1) = [∇xk+1 hTk+1

]Tare the Jacobian matrices of fk and hk+1, respectively, eval-uated at their true values.

APPENDIX D: INTERNAL INS DETAILS

For highly manoeuvrable vehicles, the INS numerical in-tegration must properly address the angular, velocity, andposition high-frequency motions, referred to as coning,sculling, and scrolling, respectively, to avoid the buildupof estimation errors. The INS multirate approach, basedon the work detailed in Savage (1998a; 1998b), computesthe dynamic angular rate/acceleration effects using high-speed, low-order algorithms, whose output is periodically

fed to a moderate-speed algorithm that computes atti-tude/velocity resorting to exact, closed-form equations.Applications within the scope of this work are character-ized by confined mission scenarios and limited operationaltime, allowing for a simplification of the frame set to Earthand body frames and the use of an invariant gravity modelwithout loss of precision.

The inputs provided to the inertial algorithms are theintegrated inertial sensor output increments

υ(τ ) =τ∫

tk−1

ardt, α(τ ) =τ∫

tk−1

ωrdt,

where ωr represents the rate gyro triad readings and ar rep-resents the accelerometer readings, also known as the mea-sured body specific force,

ar � BaSF = B v + ω × Bv − Bg,

where Bg is the nominal local gravity vector in body coor-dinates Bg = RT Eg, and Eg is the locally constant gravityvector in Earth-fixed coordinates.

The inertial sensor readings are corrupted by zeromean white noise n and random-walk bias, b = nb, yield-ing

ar = B v + ω × Bv − Bg − δba + na,

ωr = ω − δbω + nω,

where δb = b − b denotes bias compensation error, b is thenominal bias, b is the estimated bias, and the subscripts a

and ω identify accelerometer and rate gyro quantities, re-spectively.

The attitude moderate-speed algorithm (Savage,1998b) computes body attitude in DCM form,

Bk−1Bk

R(λk) = I3 + sin ‖λk‖‖λk‖ (λk)× + 1 − cos ‖λk‖

‖λk‖2 (λk)2×, (D1)

where the operator ‖ · ‖ is the usual l2-norm for vectorssuch that ‖x‖ = (xT x)

12 , {Bk} is the body frame at time k,

and Bk−1Bk

R(λk) is the rotation matrix from {Bk} to {Bk−1} co-ordinate frames, parametrized by the rotation vector λk .The rotation vector updates are based on the Bortz equa-tion (Bortz, 1971), and are formulated as

λk = αk + βk

in order to denote angular integration and coning attitudeterms αk and βk , respectively. The attitude high-speed algo-rithm computes βk as a summation of the high-frequencyangular rate vector changes using simple, recursive com-putations (Savage, 1998a), providing high-accuracy results.

Using the equivalence between strap-down attitudeand velocity/position algorithms (Roscoe, 2001), the samemultirate approach is applied (Savage, 1998b) to computeexact velocity updates at moderate speed,

vk = vk−1 + EBk−1

RBk−1 vSF k + vG/Cor k,

Journal of Field Robotics DOI 10.1002/rob

Page 28: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

Morgado et al.: An Experimental Validation • 169

where Bk−1 vSF k is the velocity increment related to thespecific force, and vG/Cor k represents the velocity incre-ment due to gravity and the Coriolis effect (Savage, 1998b).The term Bk−1 vSF k also accounts for high-speed velocityrotation and high-frequency dynamic variations due to an-gular rate vector rotation, yielding

Bk−1 vSF k = υk + vrot k + vscul k,

where vrot k and vscul k are the rotation and scullingvelocity increments, respectively, computed by the high-frequency algorithms. Interestingly enough, a standardlow-power consumption DSP-based hardware architectureis sufficient for running the INS algorithms using maximalcomputational accuracy at high execution rates. This allowsusing maximal precision so that the computational accu-racy of the INS output is only diminished by the inertialsensor noise and biases effects.

REFERENCES

Alonso, R., & Shuster, M. (2002). Complete linear attitude-independent magnetometer calibration. Journal of the As-tronautical Sciences, 50(4), 477–490.

Austin, T. (1994). The application of spread spectrum signalingtechniques to underwater acoustic navigation. In Proceed-ings of the 1994 Symposium on Autonomous UnderwaterVehicle Technology, 1994, AUV ’94.

Bar-Shalom, Y., Li, X., Kirubarajan, T., & Wiley, J. (2001). Esti-mation with applications to tracking and navigation. Wi-ley Online Library.

Bingham, B., Blair, B., & Mindell, D. (2007). On the design ofdirect sequence spread-spectrum signaling for range es-timation. In Proceedings of the OCEANS ’07 MTS/IEEE,Vancouver, Canada.

Bortz, J. (1971). A new mathematical formulation for strap-down inertial navigation. IEEE Transactions on Aerospaceand Electronic Systems, 7(1), 61–66.

Bowen, A., Yoerger, D., Taylor, C., McCabe, R., Howland, J.,Gomez-Ibanez, D., Kinsey, J., Heintz, M., McDonald, G.,Peters, D., et al. (2009). The Nereus hybrid underwa-ter robotic vehicle. Underwater Technology: The Interna-tional Journal of the Society for Underwater, 28(3), 79–89.

Britting, K. (1971). Inertial navigation systems analysis. JohnWiley & Sons, Inc.

Brown, R. G., & Hwang, P. Y. C. (1997). Introduction to randomsignals and applied Kalman filtering. 3rd ed. John Wiley& Sons.

Crassidis, J. (2006). Sigma-point Kalman filtering for inte-grated gps and inertial navigation. IEEE Transactions onAerospace and Electronic Systems, 42(2), 750–756.

D.SignT (2010). D.Module.C6713 technical data sheet. http://www.dsignt.de/download/tdd6713.pdf. Rev. 1.4.

Eustice, R., Singh, H., & Whitcomb, L. (2011). Synchronous-clock, one-way-travel-time acoustic navigation for under-water vehicles. Journal of Field Robotics, 28(1), 121–136.

Gandhi, M., & Mili, L. (2010). Robust Kalman filter based ona generalized maximum-likelihood-type estimator. IEEETransactions on Signal Processing, 58(5), 2509–2520.

Gelb, A. (1974). Applied optimal estimation. Cambridge, MA:MIT Press.

Goshen-Meskin, D., & Bar-Itzhack, I. (1992). Observabilityanalysis of piece-wise constant systems - part I: Theory.IEEE Transactions on Aerospace and Electronic Systems,28(4), 1056–1067.

Grewal, M., Weill, L., Andrews, A., & Wiley, J. (2007). Globalpositioning systems, inertial navigation, and integration.Wiley Online Library.

Imagery, N., & Agency, M. (2000). World geodetic system1984—its definition and relationships with local geode-tic systems. Technical report, Department of Defense.TR8350.2 Amendment 1.

Jaffre, F., Austin, T., Allen, B., Stokey, R., & Von Alt, C. (2005).Ultra short baseline acoustic receiver/processor. In Pro-ceedings of the OCEANS ’05 MTS/IEEE, vol. 2, pp. 1382–1385, Washington, D.C.

Jalving, B., Gade, K., Hagen, O., & Vestgard, K. (2003). A tool-box of aiding techniques for the HUGIN AUV integratedinertial navigation system. In Proceedings of the OCEANS’03 MTS/IEEE, San Diego, CA.

Kinsey, J., Eustice, R., & Whitcomb, L. (2006). A survey of un-derwater vehicle navigation: Recent advances and newchallenges. In Proceedings of the 7th Conference on Ma-noeuvring and Control of Marine Craft (MCMC2006), Lis-bon, Portugal. IFAC.

Kinsey, J. C., & Whitcomb, L. L. (2004). Preliminary field expe-rience with the DVLNAV integrated navigation system foroceanographic submersibles. Control Engineering Prac-tice, 12(12), 1541–1548. Invited paper.

Knight, D. (1997). Rapid development of tightly-coupledGPS/INS systems. IEEE Aerospace and Electronic Sys-tems Magazine, 12(2), 14–18.

Koifman, M., & Bar-Itzhack, I. (1999). Inertial navigation sys-tem aided by aircraft dynamics. IEEE Transactions onControl Systems Technology, 7(4), 487–493.

Lee, P., Jeon, B., Kim, S., Choi, H., Lee, C., Aoki, T., & Hyaku-dome, T. (2004). An integrated navigation system for au-tonomous underwater vehicles with two range sonars, in-ertial sensors and Doppler velocity log. In Proceedings ofthe OCEANS ’04 MTS/IEEE, vol. 3, pp. 1586–1593, Kobe,Japan.

Lurton, X., & Millard, N. (1994). The feasibility of a very-longbaseline acoustic positioning system for AUVs. In Pro-ceedings of the OCEANS ’04 IEEE Europe, vol. 3, pp. 403–408, Brest, France.

Macmillan, S., McLean, S., & Maus, S. (2004). The US/UKworld magnetic model for 2005-2010. Technical report,NOAA.

Menold, P., Pearson, R., & Allgower, F. (1999). Online outlierdetection and removal. In Proceedings of the 7th Inter-national Conference on Control and Automation MED99,pp. 1110–1134, Haifa, Israel.

Miller, P., Farrell, J., Zhao, Y., & Djapic, V. (2010). Autonomousunderwater vehicle navigation. IEEE Journal of OceanicEngineering, 35(3), 663–678.

Milne, P. (1983). Underwater acoustic positioning systems.Gulf Pub. Co.

Journal of Field Robotics DOI 10.1002/rob

Page 29: Tightly coupled ultrashort baseline and inertial navigation system for underwater ... · 2015. 7. 2. · ern navigation systems with application to underwater ve-hicles, focusing

170 • Journal of Field Robotics—2013

Morgado, M., Oliveira, P., & Silvestre, C. (2010). Design and ex-perimental evaluation of an integrated USBL/INS systemfor AUVs. In Proceedings of the 2010 IEEE InternationalConference on Robotics and Automation (ICRA), pp.4264–4269, Anchorage, AK.

Morgado, M., Oliveira, P., & Silvestre, C. (2011). A closed-loopdesign methodology for underwater transducers pulse-shaping. In Proceedings of the 2011 IEEE InternationalConference on Mechatronics and Automation (ICMA),pp. 2014–2019, Beijing, China.

Morgado, M., Oliveira, P., Silvestre, C., & Vasconcelos, J. (2006).USBL/INS tightly-coupled integration technique for un-derwater vehicles. In Proceedings of the 9th IEEE Interna-tional Conference on Information Fusion, Florence, Italy.

Napolitano, F., Cretollier, F., & Pelletier, H. (2005). GAPS, com-bined USBL + INS + GPS tracking system for fast de-ployable and high accuracy multiple target positioning.In Proceedings of the OCEANS ’05 IEEE Europe, Brest,France.

Papoulis, A. (1984). Probability, Random Variables, andStochastic Processes. McGraw-Hill, second edition.

Pascoal, A., Oliveira, P., & Silvestre, C., et al. (2000). Roboticocean vehicles for marine science applications: The Euro-pean ASIMOV Project. In Proceedings of the OCEANS ’00MTS/IEEE, Rhode Island.

Roscoe, K. (2001). Equivalency between strapdown inertialnavigation coning and sculling integrals/Algorithms.AIAA Journal of Guidance, Control, and Dynamics, 24(2),201–205.

Rugh, W. (1996). Linear system theory. 2nd ed. Prentice-Hall.Sanz, P., Ridao, P., Oliver, G., Melchiorri, C., Casalino, G., Sil-

vestre, C., Petillot, Y., & Turetta, A. (2010). TRIDENT:A framework for autonomous underwater interventionmissions with dexterous manipulation capabilities. InProceedings of the 7th Symposium on Intelligent Au-tonomous Vehicles IAV-2010. IFAC.

Sarwate, D., & Pursley, M. (1980). Cross-correlation propertiesof pseudorandom and related sequences. Proceedings ofthe IEEE, 68(5), 593–619.

Savage, P. (1998a). Strapdown inertial navigation integra-tion algorithm design part 1: Attitude algorithms. AIAAJournal of Guidance, Control, and Dynamics, 21(1),19–28.

Savage, P. (1998b). Strapdown inertial navigation integrationalgorithm design part 2: Velocity and position algorithms.AIAA Journal of Guidance, Control, and Dynamics, 21(2),208–221.

Simandl, M., Kralovec, J., & Tichavsky, P. (2001). Filtering, pre-dictive, and smoothing Cramer-Rao bounds for discrete-time nonlinear dynamic systems. Automatica, 37(11),1703–1716.

Smith, S., & Kronen, D. (1997). Experimental results of aninexpensive short baseline acoustic positioning system

for AUV navigation. In Proceedings of the OCEANS ’97MTS/IEEE, vol. 1, pp. 714–720, Halifax, Nova Scotia,Canada.

Sukkarieh, S., Nebot, E., & Durrant-Whyte, H. (1999). A highintegrity IMU/GPS navigation loop for autonomous landvehicle applications. IEEE Transactions on Robotics andAutomation, 15(3), 572–578.

Tichavsky, P., Muravchik, C. H., & Nehorai, A. (1998). Poste-rior Cramer-Rao bounds for discrete-time nonlinear filter-ing. IEEE Transactions on Signal Processing, 46(5), 1386–1396.

Tolstoy, A. (1993). Matched field processing for underwateracoustics. River Edge, NJ: World Scientific.

Van Trees, H. L. (1966). Bounds on the accuracy attainablein the estimation of continuous random processes. IEEETransactions on Information Theory, 12, 298–305.

Van Trees, H. L. (1968). Detection, estimation, and modulationtheory. Part 1. Detection, estimation, and linear modula-tion theory. New York: Wiley.

Van Trees, H. L., & Bell, L. K. (2007). Bayesian bounds for pa-rameter estimation and nonlinear filtering/tracking. 1sted. John Wiley & Sons–IEEE Press.

Vasconcelos, J. F., Silvestre, C., & Oliveira, P. (2011). INS/GPSaided by frequency contents of vector observations withapplication to autonomous surface crafts. Journal ofOceanic Engineering, 36(2), 347–363.

Vickery, K. (1998). Acoustic positioning systems. Newconcepts—The future. In Proceedings of the 1998 Work-shop on Autonomous Underwater Vehicles, AUV’98,Cambridge, MA.

Whitcomb, L. L. (2000). Underwater robotics: Out of the re-search laboratory and into the field. In Robotics and Au-tomation, 2000. Proceedings, IEEE International Confer-ence on ICRA’00, vol. 1, pp. 709–716.

Yi, Y., & Grejner-Brzezinska, D. (2006). Tightly-coupledGPS/INS integration using unscented Kalman filter andparticle filter. In Proceedings of the 19th InternationalTechnical Meeting of the Satellite Division of The Instituteof Navigation (ION GNSS), pp. 2182–2191, Fort Worth,TX.

Yli-Hietanen, J., Kalliojarvi, K., & Astola, J. (1996). Low-complexity angle of arrival estimation of wideband sig-nals using small arrays. In Proceedings of the 8th IEEESignal Processing Workshop on Statistical Signal and Ar-ray Processing.

Yun, X., Bachmann, E., McGhee, R., Whalen, R., Roberts, R.,Knapp, R., Healey, A., & Zyda, M. (1999). Testing and eval-uation of an integrated gps/ins system for small auv nav-igation. IEEE Journal of Oceanic Engineering, 24(3), 396–404.

Zhong, Z., Meng, H., & Wang, X. (2010). A comparison of pos-terior Cramer-Rao bounds for point and extended targettracking. IEEE Signal Processing Letters, 17(10), 819–822.

Journal of Field Robotics DOI 10.1002/rob