Adaptive Identification of Dynamically Positioned Underwater Robotic Vehicles

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO.

4, JULY 2003 505

Adaptive Identification of Dynamically Positioned


Underwater Robotic Vehicles
David A. Smallwood, Student Member, IEEE, and Louis L. Whitcomb, Senior Member, IEEE

Abstract—This paper reports a stable online adaptive identifica- evaluation of this approach are new and have not been previ-
tion technique for the identification of finite-dimensional dynam- ously reported.
ical models of dynamically positioned underwater robotic vehi- The paper is organized as follows. Section I-A reviews the
cles. Proofs for the identifier’s global stability, and for the input-to-
state stability of this class of plants are reported. A direct compar- finite-dimensional modeling of actuated marine vehicles. Sec-
ison of the adaptive identification method to a conventional, off- tion I-B briefly outlines standard techniques for model param-
line, least-squares method is reported. Using experimental data ob- eter identification. Section II reviews the application of conven-
tained with the Johns Hopkins University Remotely Operated un- tional least-squares parameter identification to this plant. Sec-
derwater robotic Vehicle (JHUROV), both methods are employed tion III reports a stable adaptive identifier for this plant, and ex-
to identify decoupled, single-degree-of-freedom dynamical plant
models. Performance of the resulting identified dynamical plant amines its performance in numerical simulations. Section IV re-
models is quantitatively compared to the experimentally observed ports our experimental setup. Section V quantitatively compares
motion of the actual vehicle. the performance of the two approaches with actual experimental
Index Terms—Adaptive estimation, dynamics, least-squares data, and investigates several critical implementation issues.
methods, robot dynamics, underwater vehicles.
A. Finite-Dimensional Dynamical Modeling of Marine
Vehicles
I. INTRODUCTION
Exact analysis of a rigid-body underwater vehicle’s dy-

T HIS paper addresses the problem of experimental identifi-


cation of finite-dimensional nonlinear dynamical models
for open-frame remotely operated underwater vehicles. The de-
namics includes both the finite-dimensional dynamics of the
vehicle body itself and the infinite-dimensional dynamics of
the fluid surrounding the vehicle. The former is a finite-dimen-
velopment of model-based control techniques for the dynamic sional dynamical system represented by an ordinary differential
positioning of underwater robotic vehicles has been limited by equation (ODE), but the latter is a continuous (infinite-dimen-
the lack of experimentally validated plant models. sional) dynamical system represented by the incompressible
Very few previously reported vehicle plant identification Navier–Stokes equation, a partial differential equation (PDE).
studies, [4], [18], both report the experimentally determined The overwhelming computational complexity of PDE
model parameters and quantitatively compare experimentally dynamical models has motivated the widespread use (by
observed vehicle dynamical response with that predicted by the naval architects and others) of finite-dimensional approximate
proposed analytical models. models for marine vehicles, sometimes termed “lumped pa-
This paper first reviews an approximate finite-dimensional rameter models.” The expedient of experimentally determining
dynamical model for open-frame underwater vehicles. This approximate finite-dimensional models for complex fluid
class of plants is then shown to possess input-to-state stability phenomenon is widely employed in naval architecture and the
(ISS). Using this property, we report a stable on-line adaptive field of underwater robotic vehicles [6]–[9], [11]–[13].
identification method for this class of plants. The performance Most reported finite-dimensional plant models for holonomic
of the proposed adaptive identification technique is evaluated (fully actuated) underwater vehicles take the general form
in comparison to conventional least-squares parameter identi-
fication with experimental data obtained by the authors. The (1)
quality of the resulting identified plant models is evaluated
by quantitatively comparing the model’s predicted dynamical where is a vector of the vehicle position and
behavior to actual experimentally observed vehicle behavior. orientation in world inertial coordinates; and
To the best of our knowledge, the application of this general are, respectively, vectors of the vehicle velocity
identification approach to this specific class of plants, the ISS and acceleration in vehicle body coordinates; is a
stability property of this class of plants, and the experimental vector of control forces and moments generated by thrusters and
control surfaces in body coordinates; is a mass
matrix representing both rigid-body mass and velocity-depen-
Manuscript received September 17, 2001; revised August 8, 2002. Manu-
script received in final form February 13, 2003. Recommended by Associate dent “added mass;” is a vector of vehicle drag
Editor D. S. Bayard. This work was supported by the Office of Naval Research and coriolis forces; and is a vector of vehicle buoy-
and the National Science Foundation under Grant CMS-0100783. ancy forces. Note that the velocity in inertial (world) coordinates
The authors are with The Johns Hopkins University, Baltimore, MD
21218-2686 USA. , is related to the velocity in body coordinates by a linear
Digital Object Identifier 10.1109/TCST.2003.813377 transformation of the form [9].
1063-6536/03$17.00 © 2003 IEEE
506 IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO. 4, JULY 2003

At present, there is no uniform consensus within the research TABLE I


community on the exact analytical form of the terms comprising NOMENCLATURE
, , and , nor for the force
vector .
Although not theoretically justified, in this report we adopt
the common practice of further approximating the 6-degree of
freedom (DOF) equations by neglecting off-diagonal entries and
coupling terms, tether dynamics, as well as assuming a constant
added mass [4], [18]. These simplifications are empirically jus-
TABLE II
tified by the slow velocities and modest attitude changes typical DEFINITION OF LUMPED PARAMETERS
of this class of underwater vehicles. The resulting decoupled
1-DOF dynamical equations take the form

(2)

where, for each DOF , is the net control force, is the


effective mass, and represent the hy- Defining and
drodynamic drag, and is the buoyancy. Using lumped param- , where represent
eters, (2) can be written as the sample times, then from (4) we have

(3) (5)

Thus if is full rank, then the least-squares solution for


or written in vector form as
the unknown parameter vector is given by the standard
Moore–Penrose pseudoinverse
(4)
(6)
where is the vector of lumped plant
parameters and is a nonlinear This method is computationally simple but requires that the
vector of state and the control input . A nomenclature is velocity and acceleration of the vehicle are instrumented in ad-
given in Table I. The lumped parameters are defined in Table II. dition to the applied force. Instrumentation of acceleration is
not standard on most underwater vehicles. Further, calculation
B. Experimental Identification of Plant Parameters of acceleration via direct numerical differentiation, an acausal
operation, is easily corrupted by sensor noise.
Several methods of determining the parameters of dy- For completeness, the authors would like to point out that
namical equations of motion for underwater vehicles have there are valid alternative approaches to the least-squares
been presented in the literature. These include the method of method in the literature [2], [15], [22], that eliminate the need
least-squares [4], [5], [10] and the use of extended Kalman for acceleration measurement. Essentially these approaches
filters (EKFs) [1], [5]. A method based on numerical mini- propose low-pass filtering (5) to eliminate the need for acceler-
mization of the error between the trajectories of the vehicle ation measurement.
and several models during free decay, in 1-DOF, was presented
in [18]. In special idealized cases—e.g., symmetric vehicles III. ONLINE ADAPTIVE PARAMETER IDENTIFICATION
in inviscid flow [17]—plant parameters can be determined
analytically, but this approach does not extend to general This section presents a stable adaptive technique for param-
vehicle shapes in viscous fluid. eter identification as an alternative to the offline least-squares
This paper will present two identification methods and technique. This approach can be implemented online, and
compare the results obtained with data from free running exper- does not require instrumentation of acceleration. The general
imental trials. The first is a conventional offline least-squares approach to plant parameter identification employed herein is
scalar identification method. The second is a novel application well known [19], [20], the application to this class of plants and
to this class of plants of a well known online identification experimental evaluation reported herein is new. This adaptive
method utilizing a stable adaptive identifier [19], [20]. approach to plant parameter identification utilizes a) the actual
plant with full state access to state vector, , (4), and b)
an “identification” plant, with state vector , (7), and c) a
II. OFFLINE LEAST-SQUARES PARAMETER IDENTIFICATION
parameter update law, (11). The goal of this approach is to
Parameters for plants of the form (4) can be identified ex- ensure that all signals remain bounded and that, under sufficient
perimentally using the standard least-squares technique. This conditions of persistent excitation, the plant parameter esti-
method requires values for , , and for each DOF mates converge to the true plant parameter values. Note that the
. purpose of this identifier is to identify dynamical plant model
SMALLWOOD AND WHITCOMB: ADAPTIVE IDENTIFICATION OF DYNAMICALLY POSITIONED UNDERWATER ROBOTIC VEHICLES 507

parameter values, not the state of the plant. The application 4) In consequence, , , and are bounded.
of this general identification approach to this specific class of 5) From (12) and that and are bounded, we can
plants, the ISS stability property of this class of plants, and the then conclude that . Then using a corollary to
experimental evaluation of this approach are new and have not Barbalat’s lemma from [19], since is bounded, we
been previously reported. can conclude .
Given the scalar plant (4), the chosen adaptive estimator is 6) Given that , and is bounded, then
from (11) we conclude .
(7) In summary, all signals remain bounded, the identifier
state error, , converges asymptotically to zero, and the
where and
time-derivative of the parameter error, , converges
. was used in the adaptive iden-
asymptotically to zero. Although the parameter estimation
tifier instead of because of the typical noise associated with
error remains bounded, absent additional arguments we cannot
the measurement of actual plant velocity, . is an esti-
conclude that the parameter error converges asymptotically
mate of the unknown parameter vector , and is a scalar
to zero. It is well known, however, that the parameter error
gain. The identifier error coordinates are defined as
for this general type of system will converge to zero given a
sufficiently rich input signal [19], [20], [21]. Numerical
simulation of the full system, reported in the next section,
indicates that asymptotic convergence of the plant parameters
occurs for most time varying inputs.
(8)
A. Performance of Adaptive Identifier on a Simulated
Consider the Lyapunov function candidate of the form Analytical Plant
How does this scalar adaptive identifier perform on an “ideal”
(9) simulated analytical plant? Consider the scalar single DOF an-
alytical plant model of the form
where , for i , with , whose time
derivative is (13)

where , , subject
to a propulsion thrust profile of
[N].
(10)
This model represents a vehicle of 250 kg total mass (140 kg
If we choose the parameter update law to be dry mass, 110 kg added mass) that is experiencing both inertial
forces and quadratic drag forces, and is undergoing a 1/4 Hz sin
(11) wave propulsion thrust profile of peak magnitude 100 N. The
velocity error feedback gain was set as , and the
then (10) becomes adaptation gains were set to and .
A fourth order Runge–Kutta ODE integrator was used to con-
duct this simulation and identification. The performance of the
adaptive identifier on the analytical plant is shown in Fig. 1. As
predicted by theory, , , and are bounded and
goes to zero. In addition we see that goes to zero,
(12)
and the estimated parameters converge to the actual simulated
Examining the term we plant values, .
see that when we have which The adaptive identifier offers two advantages in comparison
results in . Likewise to conventional off-line least-squares techniques. First, it does
if then we have , resulting in not require acceleration instrumentation. Second, the adaptive
. Since , , and identifier can be implemented on-line and suffers from no rank
are negative scalar constants, it is clear that (12) is negative deficiency singularity problems.
semidefinite.
We conclude the following points. IV. EXPERIMENTAL SETUP
1) From (9) and (12), we conclude that and The Johns Hopkins University Remotely Operated under-
are bounded. water robotic Vehicle (JHUROV), Fig. 2, is a tethered, remotely
2) Given that is bounded and is a constant, then operated underwater robot. The JHUROV is powered by an
is bounded as well. isolated 10 kW dc power supply. The mass of the vehicle is
3) Given that (4) is bounded-input bounded-state (BIBS) 140 kg and it’s dimensions are 1.5 m long 1 m wide 0.6 m
stable (see Appendix A), and that the input is high. The XYZ position and heading of the vehicle are actively
bounded, then and are bounded. controlled and the vehicle is passively stable in roll and pitch.
508 IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO. 4, JULY 2003

(a)

(b)
Fig. 1. Numerical simulation showing the performance of the adaptive identifier, (7) and (11), on an “ideal” simulated plant, (13). (a) The simulated plant velocity
1 () 1 () 1 ()
(top) and the adaptive identifier estimated plant velocity (bottom). (b) Identifier velocity error v t versus time (top); Parameter error t and t versus
()
time (middle); and Lyapunov function W t versus time (bottom).

Actuation is provided by five dc brushless electric thrusters. A analog pressure transducer, sampled at 20 Hz. A KVH Azimuth
complete description of the JHUROV is reported in [23]. Digital Gyro Compass (ADGC) measured the vehicle’s roll
The ROV is equipped for full 6-DOF position measurement. and pitch. Heading was measured using a prototype Litton
For this set of experiments, XYZ position was measured LN200 IMU. Thruster specifications are given in Table III. The
using a 300 kHz Sonic High Accuracy Ranging and Posi- vehicle’s entire sensor suite is listed in Table 4.
tioning System (SHARPS) time-of-flight acoustic hardwired While not directly instrumented, velocity was obtained by
transponder system. Depth was instrumented via a Foxboro/ICT directly numerically differentiating the position and running it
SMALLWOOD AND WHITCOMB: ADAPTIVE IDENTIFICATION OF DYNAMICALLY POSITIONED UNDERWATER ROBOTIC VEHICLES 509

aware of more advanced thruster models [3], [14] that are more
precise, however they require a dynamical thruster characteri-
zation not available at the time of the experiments.
The experiments reported in this paper were conducted using
the JHUROV at the United States Naval Academy’s Hydrome-
chanics Laboratory, Annapolis, MD. The work was done in col-
laboration with Dr. D. Stilwell of the Virginia Polytechnic In-
stitute and State University, who deployed his inertial measure-
ment unit (IMU) on the vehicle during these experiments to ex-
amine the issue of IMU calibration.

V. EXPERIMENTAL RESULTS
This section reports a comparative experimental evaluation of
the following two techniques for determining plant-model pa-
Fig. 2. Physical layout of the JHUROV. rameters from free-motion experimental data: 1) conventional
least-squares parameter identification and 2) the proposed adap-
TABLE III tive parameter identification technique. For each 1) and 2), we
THRUSTER SPECIFICATIONS report the following steps: First, we state the procedure followed
to execute the proposed identification technique. Second, we re-
port the resulting identified plant parameters. Third, we evaluate
the “quality” of the identified plant parameter values by quan-
titatively comparing experimentally observed plant dynamical
performance to that predicted by the identified plant model.
In addition, this section also reports the effect of different
TABLE IV thrust profiles on the dynamic performance of the identified dy-
JHUROV EXPERIMENTAL INSTRUMENTATION namical plant models.

A. Least-Squares Identified Parameters


We employed the conventional least-squares parameter iden-
tification outlined in Section II, to experimentally determine
plant parameter values for the JHUROV. First, dynamic vehicle
through a low-pass filter. Acceleration was obtained by directly trials were conducted with motion in the , , and DOF.
numerically differentiating the velocity. At the time we per- For each experiment, estimates for the plant parameters were de-
formed the experiments reported herein, our vehicle did not have termined using the least-squares method. Next numerical sim-
on-board accelerometers. However, in subsequent experiments, ulations were run on the identified parameters. The simulation
we have found that the vehicle acceleration measurements ob- was run using the actual logged force profile from the exper-
tained as reported in this paper (by differentiating the highly ac- iment. The output of the simulation was a velocity, ,
curate SHARPS LBL positions) compare favorably with accel- which was compared to the actual logged experimental plant
eration measurements obtained directly with the accelerometers velocity, . The error, for each DOF , was calculated as
in an IXSEA OCTANS north-seeking fiber-optic gyrocompass mean . Plots of , , and are
unit mounted on the vehicle. shown in Fig. 3. The parameters that performed best (i.e., had
For these experiments, the vehicle was actuated by two the lowest “error,” ) of all the experiments in numerical simu-
longitudinal thrusters and one vertical thruster. There were lations are listed in Table V. Fig. 3 shows that the simulated plant
no thrusters arranged to provide pure lateral movement. Free model performance agrees closely with actual experimental data
running experiments were conducted in the , , and when the model employs the least-squares identified parameters
DOF. Multiple trials in each DOF, one DOF at a time, were listed in Table V.
conducted with sinusoidal thrust profiles of varying magnitude
and frequency. The thrust profiles employed in the single DOF B. Adaptively Identified Parameters
experiments were chosen to minimize the effects of cross We employed the stable adaptive scalar identification method
coupling in the dynamics. General maneuvers involving cross outlined in Section III to experimentally determine plant pa-
coupled dynamics were not performed. rameters for the JHUROV. Using the same experimental data
A static thrust control law was used for these experiments to as used in the least-squares method in Section V-A, the scalar
model the thrust produced by each thruster; i.e., the thrust pro- adaptive identifier was run on each experimental data set using
duced is assumed to be proportional to the current commanded, a fourth-order Runge–Kutta ODE integrator. The output of this
where is the motor torque constant with units of process is an estimated plant velocity, , as well as estimates
[N-m/A] and is the current commanded in amperes. Current for the unknown plant parameters, . A plot of the output of
was controlled using current mode amplifiers. The authors are the adaptive identifier for one experimental trial in the DOF
510 IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO. 4, JULY 2003

(a)

(b)
Fig. 3. x DOF: Comparison of the simulated performance of (a) the least-squares and (b) adaptively identified plant models, (4), and actual experimental data.
The plots show (top) model-simulated and actual velocity versus time and (bottom) the velocity error between simulated and experimentally observed velocities.
Parameter values used in the simulation are listed in Tables V and VI.

can be seen in Fig. 4. Next a simulation was run using the adap- plant parameters and the “true” plant parameters. These plots
tively identified parameters. The output was a velocity, . correspond to the same adaptive identification seen in Fig. 4. In
The logged plant velocity, , was then compared to . As plotting the plant parameter error, the parameters identified by
done in Section V-A, the error for each DOF , is calculated as the least-squares method were considered to be the “true” plant
mean . The best identified parameters were parameters. The data indicates that the simulated performance
those with the lowest “error,” , and are listed in Table VI. Fig. 3 of the plant model employing the adaptively identified parame-
contains plots of and for the DOF. Fig. 5 shows ters, listed in Table VI, equals or exceeds that of the plant model
the velocity error ( ), and the error between the estimated employing the least-squares identified parameters.
SMALLWOOD AND WHITCOMB: ADAPTIVE IDENTIFICATION OF DYNAMICALLY POSITIONED UNDERWATER ROBOTIC VEHICLES 511

TABLE V
LEAST-SQUARES LUMPED PARAMETERS

Fig. 4. Performance of adaptive identifier on experimental data for a trial in the x DOF. Plot shows identifier velocity v^(t) and the plant parameter estimates
^ (t), ^ (t), ^ (t), and ^ (t). Adaptive identifier is attempting to identify a model of the form reported for the least-squares technique in Table V. 8 are the
least-squares identified parameters.

Note that while care was taken to be methodical and precise set of experimental data accurately predicts the experimental
in examining the data and using the adaptive identifier, choosing dynamic performance observed under different experimental
of the velocity error feedback gain, , and the adaptation gains conditions. For example, in the case of linear plant models for
remains a trial and error process. At present, there is no re- nonlinear systems it is well known that identified linear plant
ported analytical technique for “optimally” tuning these adap- models are typically only valid in a neighborhood around the
tation gains. operating point (an equilibrium point) at which the model was
In this section, the adaptive identifier was run on the data identified. In this paper we have adopted a nonlinear plant
using a fourth-order Runge–Kutta ODE integrator and logged model, (3), which should be valid for a wide range of operating
experimental plant state values. For online implementation, conditions.
however, fourth-order Runge–Kutta ODE is not feasible. The This section examines how an identified plant model’s per-
online implementation of the adaptive identifier was explored formance compares to experimentally observed plant behavior
using Euler integration and was found to perform comparably under thrust profiles differing from those employed to obtain the
to the Runge–Kutta ODE integrator implementation. model parameters. We performed this comparison as follows:
First, we identified a set of plant parameters , from experi-
C. Effect of Different Thrust Profiles mental data, for example “data set A,” corresponding to a partic-
When model parameters are obtained from a particular set of ular thrust profile . The response of the identified model was
experimental data, it is reasonable to expect that the dynamic then simulated both using original thrust profile , and also
performance of the resulting identified model will agree, to a substantially different experimental thrust profile, , from
some degree, to that observed in the original experimental data. “data set B.” In each case, the simulated model response was
An important question arising in experimental model identifi- compared to the actual experimentally observed response. This
cation is whether or not the a plant model identified from one was done for each thrust profile.
512 IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO. 4, JULY 2003

1 ()
Fig. 5. x DOF: Adaptive identifier performance on actual experimental data using Runge–Kutta ODE integration. The plot shows identifier velocity error v t
1 () 1 () 1 () 1 ()
versus time (top graph); Parameter error t , t ,  t , and  t versus time (four middle graphs); and Lyapunov function W t versus time ()
(bottom graph). The parameter errors and Lyapunov function value are computed with respect to the least-squares identified plant parameters in Table V.

TABLE VI
ADAPTIVE ID LUMPED PARAMETERS

We observed that a model employing parameters identified 1) Physical Interpretation of Parameters: Do the identified
using high-magnitude thrust profiles performed uniformly plant parameters seem reasonable? An examination of the
well—the single model accurately predicts experimentally parameters obtained by the two (very different) identification
observed response at all thrust levels. In contrast, a model techniques reveals that the two methods result in nearly iden-
employing parameters identified using low-magnitude thrust tical plant parameter values. Moreover, the identified parameter
profiles performed poorly—the single model predicts experi- values are consistent with our physical understanding of the
mentally observed response only for thrust profiles similar to vehicle’s design.
its model-identification data set. The JHUROV, Fig. 2, has a total dry mass of 140 kg. In the
In summary, both identification methods (least-squares and (surge) direction, in which the vehicle presents a minimal cross-
adaptive) were observed to perform best when employing data section, it exhibits an added mass of about 150 kg. As expected,
from trials containing thrust profiles of higher magnitude and in the (heave) direction, in which the vehicle presents a large
lower frequency. Under these conditions, the vehicle experi- flat 1.7 m cross section, the vehicle exhibits an added mass
enced the largest variation in position, velocity, and accelera- about ten to 20 times larger than in the (surge) direction.
tion. We surmise that this improves the signal-to-noise ratio of The identified parameters indicate that drag is highest in the
the experimental data and, in consequence, provides for more DOF. This is physically consistent with the ROV’s layout, as
accurate model parameter identification. seen in Fig. 2. Once again this is due to the fact that the vehicle
is more or less a flat plate moving in the DOF, while the
D. Discussion of Results vehicle’s layout has a smaller cross-sectional area in the and
DOF.
This section first discusses the physical interpretation of the The buoyancy term represents a combination of the actual
results and, second, articulates a comparison of the least-squares buoyant force and any possible thruster zero-calibration offset
and adaptive identification techniques. error. Only in the DOF did we identify a significant amount
SMALLWOOD AND WHITCOMB: ADAPTIVE IDENTIFICATION OF DYNAMICALLY POSITIONED UNDERWATER ROBOTIC VEHICLES 513

TABLE VII
DIRECT COMPARISON OF ADAPTIVE AND LEAST-SQUARES IDENTIFIED LUMPED PARAMETERS

of buoyant force. This is consistent with our ballasting of the TABLE VIII
vehicle, the ROV was deliberately trimmed to be slightly nega- MAP OF NOMENCLATURE BETWEEN THIS PAPER AND [28] AND [29]
tively buoyant. The values listed for the buoyancy term in the
and DOF are much smaller in magnitude than the value for
the DOF. We surmise that these small fictitious “buoyancy”
forces are the result of a small zero-offset in the thruster current
amplifier calibration.
2) Comparison of Least-Squares and Adaptive Identification
Techniques: How does the performance of the two identifica-
tion (adaptive versus least-squares) methods compare? Com- acceleration instrumentation. One disadvantage of the adaptive
paring the “error,” as defined in Sections V-A and V-B, of method is that it possesses several gains that must be empirically
each method in predicting the dynamical response of the ROV tuned.
listed in Tables V and VI, it is clear that the plant models iden-
tified by the adaptive method do a better job predicting the VI. CONCLUSION
ROV dynamical motion. Note that the models identified by each
method do not necessarily contain the same lumped parameters. This paper has presented to the best of our knowledge, the first
experimental evaluation of a provably stable, on-line adaptive
How does the performance of the two identification methods
identification technique for an underwater robotic vehicle and
compare when they are both used to identify plant models
a direct comparison to a conventional, off-line, least-squares
consisting of the same lumped parameters? Table VII directly
identification technique. Decoupled, single DOF dynamical
compares the dynamic response of plant models consisting of
plant models identified by both techniques were presented and
the same lumped parameters (with parameters identified by
evaluated. The adaptive method was shown to produce superior
the adaptive and least-squares approaches, respectively) to the
dynamical plant models to the least-squares technique. The effect
actual observed experimental dynamic response. The “error,”
, is the mean absolute difference between the identified of different thrust profiles on model parameter identification
model simulated response and the actual logged experimental was also investigated, confirming that both the least-squares and
response. The data shows the adaptively identified parameters the adaptive identification techniques performed better when
to yield a plant model whose behavior corresponds more used on thrust profiles of higher magnitude. When nonlinear
exactly to the experimental data than do the least-squares iden- plant models of the form (3) are properly identified, they exhibit
tified parameters. Although there is no theoretical algorithmic both steady-state and dynamic response that closely agrees
justification for the adaptive method’s apparently superior with the experimentally observed response over a wide range of
performance, we surmise that the difference is principally operating conditions. Such performance would not be expected
due to their respective instrumentation requirements. In our of a linearized plant model.
experiments, vehicle position was measured directly with a As mentioned in Section I, the development of model-based
300 kHz long-baseline tracking system at 10 Hz to an accuracy control techniques for the dynamic positioning of underwater
of several millimeters. The vehicle velocity and acceleration robotic vehicles has been limited by the lack of experimentally
signals were obtained by direct differentiation of the position validated plant models. This paper has reported experimentally
signals. This is theoretically objectionable, but in practice it validated plant models for the JHUROV. These reported models
is a necessity. Given that the least-squares method requires served as the foundation for the design of a family of model-based
both velocity and acceleration signals, while the adaptive controllers for the JHUROV. Extensive experimental and nu-
technique only requires velocity signals, we surmise that the merical simulation of the performance of these model-based
apparently poorer performance of the least-squares approach is controllers has been reported by the authors in [24]–[26]. The
a consequence of numerical differentiation “noise.” accurate trajectory tracking performance of the model-based
In short, both methods perform well. It should be noted here controllers reported in these papers serves as an additional tes-
that both methods identified simple, nonlinear dynamical plant tament to the validity of their ability to represent the dynamical
models that were validated over a range of operating conditions, motion of this class of underwater robotic vehicles.
something that would not be expected with the use of a lin-
earized plant model. The adaptive technique performs slightly APPENDIX
better and has the advantage of not requiring acceleration in- The decoupled dynamical system (3) has the property that
strumentation, whereas the least-square method does require bounded control inputs gives rise to bounded state. Note
514 IEEE TRANSACTIONS ON CONSTROL SYSTEMS TECHNOLOGY, VOL. 11, NO. 4, JULY 2003

that the state of the system in this formulation is the plant ve- then
locity . This property can be established directly, by a Lya-
punov-like analysis, or indirectly, by employing recent general
results on ISS. The latter approach is reported here. First, in Ap-
pendix A, the vehicle plant with zero-buoyancy is shown to be and
ISS. Second, in Appendix B, the general vehicle plant with ar-
bitrary buoyancy is shown to be BIBS but not ISS.
which results in
A. Input to State Boundedness: Case With Zero Buoyancy
Recently reported advances on ISS provide a well-defined (19)
general definition of input-output stability for general nonlinear
systems, [27] as well as a family of general Lyapunov-like nec-
Applying (19)
essary and sufficient conditions for ISS systems, e.g., [29]. A
map of the nomenclature used here and that used by Sontag [28],
[29], is provided in Table VIII for clarification.
First consider the case where in which the plant buoyancy
term is zero, i.e., . The decoupled dynamical system (3)
then takes on the form (20)

(14) Thus we conclude the system (14) is ISS.

where , and as defined in (2) and Table II, B. Input to State Boundedness: Case With Nonzero Buoyancy
. In this case, the function The nonzero buoyancy system (3), of the form (16) is not ISS
stable because, among other reasons, it no longer satisfies the
(15) equilibrium condition . The system (3) can, how-
ever, be shown to possess the weaker property of BIBS stability
is an ISS Lyapunov function in the sense reported in [28] and in the usual sense, as follows: Consider the system (3) rewritten
[29]. This is established by showing that the system (14) satis- with a new control input, , defined as
fies the following conditions.
1) The system (14) has the form (21)

(16) The resulting system takes the form

where is continuously differentiable and (22)


.
2) The function is a smooth map from to . identical to the original no-buoyancy case (14) which, from Ap-
3) The functions and pendix-A, is ISS and, in consequence, BIBS. Given that (22) is
are, by inspection, radially unbounded compar- BIBS, and (21) ensures that, since is finite by assumption,
ison functions, , [16], and trivially satisfy then if is finite, is also finite. In summary, finite
the relation implies finite , and finite implies finite , thus (3)
is BIBS.
(17)
ACKNOWLEDGMENT
4) The functions and The authors gratefully acknowledge our ongoing collabora-
are both radially unbounded com- tion with Prof. D. Stilwell of the Virginia Polytechnic Institute
parison functions, , satisfying the relation and State University, who tested his inertial navigation system
that aboard the JHUROV during these experiments. The authors are
indebted to the USNA Hydromechanics Laboratory for their
(18) world-class testing facilities, and to Dr. D. Yoerger of the Woods
Hole Oceanographic Institution, who graciously provided use of
for any and any such that his 300 kHZ SHARPS acoustic transponder system.
. That is, if
REFERENCES
[1] A. Alessandri, M. Caccia, G. Indiveri, and G. Veruggio, “Application
of LS and EKF techniques to the identification of underwater vehicles,”
in Proc. IEEE Int. Conf. Control Applications, Trieste, Italy, 1998, pp.
1084–1088.
SMALLWOOD AND WHITCOMB: ADAPTIVE IDENTIFICATION OF DYNAMICALLY POSITIONED UNDERWATER ROBOTIC VEHICLES 515

[2] M. Amestegui, R. Ortega, and J. Ibarra, “Adaptive computed torque con- [21] J. Slotine and W. Li, Applied Nonlinear Control. Englewood Cliffs,
trol for rigid link manipulators,” in Proc. 25th Conf. Decision Control, NJ: Prentice-Hall, 1991.
Athens, Greece, 1986, pp. 68–73. [22] J.-J. E. Slotine and W. Li, “Composite adaptive control of robot manip-
[3] R. Bachmayer, L. L. Whitcomb, and M. Grosenbaugh, “An accurate ulators,” Automatica, vol. 25, no. 4, pp. 509–519, July 1989.
four-quadrant nonlinear dynamical model for marine thrusters: Theory [23] D. Smallwood, R. Bachmayer, and L. Whitcomb, “A new remotely op-
and experimental validation,” IEEE J. Oceanic Eng., vol. 25, pp. erated underwater vehicle for dynamics and control research,” in Proc.
146–159, Jan. 2000. 11th Int. Symp. Unmanned Untethered Submersible Technol., Durham,
[4] M. Caccia, G. Indiveri, and G. Veruggio, “Modeling and identification NH, 1999, pp. 370–377.
of open-frame variable configuration underwater vehicles,” IEEE J. [24] D. A. Smallwood and L. L. Whitcomb, “Preliminary experiments in the
Oceanic Eng., vol. 25, pp. 227–240, Apr. 2000. adaptive identification of dynamically positioned underwater robotic ve-
[5] S. Eiani-Cherif, G. Lebret, and M. Perrier, “Identification and control hicles,” in Proc. 2001 IEEE/RSJ Int. Conf. Intelligent Robots Systems,
of a submarine vehicle,” in Proc. 5th IFAC Symp. Robot Contr., Nantes, Maui, HI, 2001, pp. 1803–1810.
France, 1997, pp. 307–312. [25] , “Toward model based dynamic positioning of underwater robotic
[6] J. P. Feldman, “State-of-the-art for predicting the hydrodynamic char- vehicles,” in Proc. IEEE OCEANS’ 2001, Honolulu, HI, 2001, pp.
acteristics of submarines,” in Proc. Symp. Control Theory Naval Appli- 1106–1114.
cations, Monterey, CA, June 1975, Defense Tech. Inform. Center Doc. [26] , “The effect of model accuracy and thruster saturation on tracking
A071 804, pp. 87–127. performance of model based controllers for underwater robotic vehi-
[7] , “DTNSRDC revised standard submarine equations of motion,” cles: Experimental results,” in Proc. IEEE Int. Conf. Robot. Automation,
David Taylor Naval Ship Research and Development Center, Tech. Rep., Washington, DC, 2002, pp. 1081–1086.
Defense Tech. Inform. Center Doc. A071 804, June 1979. [27] E. D. Sontag, “Smooth stabilization implies coprime factorization,”
[8] T. I. Fossen, Guidance and Control of Ocean Vehicles. New York: IEEE Trans. Automat. Contr., vol. 34, pp. 435–443, Apr. 1989.
Wiley, 1994. [28] E. D. Sontag and Y. Wang, “On characterizations of the input-to-state
[9] M. Gertler and G. Hagen, “Standard equations of motion for subma- stability property,” IEEE Syst. Contr. Lett., vol. 24, no. 5, pp. 351–359,
rine simulation,” David Taylor Naval Ship Research and Development Apr. 1995.
Center, Tech. Rep., Defense Tech. Inform. Center Doc. 653 861, June [29] , “New characterizations of input to state stability,” IEEE Trans.
1967. Automat. Contr., vol. 41, pp. 1283–1294, Sept. 1996.
[10] K. Goheen and E. Jefferys, “The application of alternative modeling
techniques to ROV dynamics,” in Proc. IEEE Int. Conf. Robot. Automat.,
May 1990, pp. 1302–1309.
[11] A. Goodman, “Experimental techniques and methods of analysis used in
submerged body research,” in Proc. 3rd Symp. Naval Hydrodynamics.
Washington, DC, 1960, pp. 379–449. David A. Smallwood received the B.S. degree from the University of Miami,
[12] A. J. Healey, “Model-based maneuvering controls for autonomous un- Coral Gables, FL, in 1997 and the M.S. degree from The Johns Hopkins Univer-
derwater vehicles,” ASME J. Dyn. Syst., Meas. Contr., vol. 114, no. 4, sity, Baltimore, MD, in 1999, where he is currently pursuing the Ph.D. degree
pp. 614–622, Dec. 1992. in the Dynamical Systems and Controls Laboratory, Department of Mechanical
[13] A. J. Healey and D. Lienard, “Multivariable sliding mode control for au- Engineering.
tonomous diving and steering of unmanned underwater vehicles,” IEEE His Ph.D. research involves the dynamics and control of underwater robotic
J. Oceanic Eng., vol. 18, pp. 327–339, July 1993. vehicles. As part of this work, he designed and built the JHUROV, a new ROV
[14] A. J. Healey, S. M. Rock, S. Cody, D. Miles, and J. P. Brown, “Toward an used for research in underwater vehicle navigation, dynamics, and control. His
improved understanding of thruster dyamics for underwater vehicles,” research interests are in the dynamics and control of electromechanical systems.
IEEE J. Oceanic Eng., vol. 20, pp. 354–361, Oct. 1995.
[15] P. Hsu, M. Bodson, S. Sastry, and B. Paden, “Adaptive indentification
and control for manipulators without using joint accelerations,” in
Proc. IEEE Int. Conf. Robotics Automation, Raleigh, NC, 1987, pp.
1210–1215.
[16] A. Isidori, Nonlinear Control Systems II. London, U.K.: Louis L. Whitcomb (S’87–M’87–SM’02) received the B.S. and Ph.D. degrees
Springer-Verlag, 1999. from Yale University, New Haven, CT, in 1984 and 1992, respectively.
[17] N. Leonard, “Stability of a bottom-heavy underwater vehicle,” Auto- From 1984 to 1986, he was an R&D Engineer with the GMFanuc Robotics
matica, vol. 33, no. 3, pp. 331–346, March 1997. Corporation. From 1992 to 1994, he was a Postdoctoral Fellow with the Woods
[18] A. T. Morrison, III and D. R. Yoerger, “Determination of the hydrody- Hole Oceanographic Institution, Woods Hole, MA, and the University of Tokyo,
namic parameters of an underwater vehicle during small scale, nonuni- Tokyo, Japan. He is an Associate Professor with the Department of Mechanical
form, 1-dimensional translation,” in Proc. IEEE/MTS OCEANS’93, Oct. Engineering, with joint appointment in the Department of Computer Science,
1993, pp. 277–282. The Johns Hopkins University, Baltimore, MD. He is codeveloper of naviga-
[19] K. Narendra and A. Annaswamy, Stable Adaptive Systems. New York: tion and control systems for underwater vehicles including the robotic vehicles
Prentice-Hall, 1988. JHUROV, Jason II, DSL120A, Isis, Hercules, and the D.S.V. Alvin inhabited
[20] S. Sastry and M. Bodson, Adaptive Control: Stability, Convergence, and submersible. His research focuses on the navigation, dynamics, and control of
Robustness. Englewood Cliffs, NJ: Prentice-Hall, 1989. robotic vehicles and systems.

You might also like