You are on page 1of 12

IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO.

3, JULY 1990 167

The Influence of Thruster Dynamics on Underwater


Vehicle Behavior and Their Incorporation Into
Control System Design

Abstmct-The system dynamics of underwater vehicles can he greatly Estimated and Desired Position
influenced by the dynamics of the vehicle thrusters. In this paper a non- O2 7
1
linear parametric model of a torque-controlled thruster is developed and
experimentally confirmed. The model shows that the thruster behaves
o,,5
-I Desired position

like a sluggish nonlinear filter, where the speed of response depends on


the commanded thrust level. A quasi-linear analysis utilizing describ
ing functions shows that the dynamics of the thruster produce a strong
bandwidth constraint and a limit cycle, both of which are commonly
seen in practice. Three forms of compensation are tested, utilizing a
hybrid simulation which combined an instrumented thruster with a real-
time mathematical vehicle model. The first compensator, a linear lead
network, is easy to implement and greatly improves performance over
the uncompensated system, but does not perform uniformly over the en-
tire operating range. The second compensator, which attempts to cancel
the nonlinear filtering effect of the thruster, is effective over the entire
-0.15 1
operating range but depends on a n accurate thruster model. The final
-0.21
-0.2
'
-0.15
'
-0.1
'
-0.05
'
0 0.05
' '
0.1
'
0.15
I
0.2
compensator, an adaptive sliding controller, is effective over the entire X (meters)
operating range and can compensate for uncertainties or the degradation Fig. I The ROV Jason shows a limit cycle when in the closed-loop position
of the thruster. control.

Good control of a vehicle at low speed is also an im-


I. INTRODUCTION
portant design problem which must be solved to permit
important operations like automatic docking and combined
T HE AUTOMATIC control of underwater vehicles repre-
sents a difficult design problem due to the nature of the
dynamics of the system to be controlled. Controllers based
vehicle-manipulator control. In this regime, nonlinearities re-
lated to thruster dynamic behavior can be very important and
on simple models of vehicle mass and drag usually yield dis- influence overall system behavior in a manner fundamentally
appointing performance. In this paper, it is shown that for a different from most hydrodynamic and inertial nonlinearities.
wide class of vehicles, the dynamics of the thrusters dominate This paper focuses specifically on the dynamic behavior
the control problem and must be properly considered to obtain of thrusters and its implication for control. The dynamics of
good results. an electrically powered thruster are modeled, and the model
The general underwater vehicle control system design prob- verified in a tank test. It is shown that the dynamics of the
lem includes a variety of nonlinearities and modeling un- thruster dominate behavior of the vehicle by restricting the
certainties. These include hydrodynamic nonlinearities, iner- maximum closed-loop bandwidth and creating a limit cycle.
tial nonlinearities, and problems related to coupling between Such limit cycles are commonly seen in most closed-loop
the degrees of freedom [l], [2]. Additionally, the ability of marine vehicles, ranging from dynamically positioned ships
thrusters to produce force is greatly influenced by axial- and to small servo-controlled remotely operated vehicles (ROV's).
cross-flow effects. These types of nonlinearities are especially Fig. 1 shows the closed-loop positioning performance of Ja-
prominent at high speed. son, a 1200-kg ROV, during a dockside test in calm wa-
ter. With no compensation for thruster dynamics, the system
Manuscript received January 6, 1990; revised March 1990. This work shows a stable oscillation about the desired position with a
was supported by the ONR under Contract Nos. N00014-86-C-0038 and fixed frequency and a magnitude of about 10 cm. Fig. 2 shows
N00014-88-K-2022, and by ONR Grant N00014-87-J-1111. This paper
represents Woods Hole Oceanographic Institution Contribution Number 7307.
similar behavior for the RV Knorr, and 1800-ton oceano-
D. R. Yoerger and J. G. Cooke are with the Deep Submergence Labora- graphic research ship, while under control of its dynamic po-
tory, Department of Applied Physics and Ocean Engineering, Woods Hole sitioning system in calm water. As in the case of the ROV,
Oceanographic Institution, Woods Hole, MA 02543. the oscillation is of fixed magnitude and frequency and is not
J.-J. E. Slotine is with the Department of Mechanical Engineering, Mas-
sachusetts Institute of Technology, Cambridge, MA 02139. caused by a forced input, such as waves. For a simple con-
IEEE Log Number 9036202. troller without an integral term or outside disturbances, such

0364-9059/90/0700-0167$01.OO 0 1990 IEEE


168 IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO. 3, JULY 1990

:“i
‘ID
Position of RV KNORR under

4638

I t:t
1
4632

4630

4628

4626
L 6080
6065 6070 6075

X (meters)
Fig. 2. RV Knorr, an 1800 ton research vessel, shows a prominent limit Fig. 3. Jason’s thrusters consist of a pressure-compensated brushless dc
cycle when under closed-loop position control. motor and controller driving a propeller mounted in a shroud.

limit cycles cannot be produced by simple nonlinearities like effects can be reasonably approximated by the modeling of
stiction, deadband, or pure delays. For the Jason example, the thruster unit alone, the velocity of the fluid entering the
it is shown that such behavior is produced by the nonlinear thruster shroud effectively changes the angle of attack of the
response of a torque-controlled thruster. propeller, thus altering the force produced. Cross-flow effects
A series of control system design procedures are outlined are much more difficult to model and have been shown experi-
that reduce the effects of the thruster dynamics. First, a linear mentally to be highly dependent on the position of the thruster
compensation scheme is shown to improve performance com- on the vehicle [3]. In both cases, the reduced amount of force
pared to the uncompensated system, but it results in nonuni- produced by the thruster will reduce the overall gain of a con-
form behavior over the entire operating regime. Nonlinear trol system unless these effects are specifically included in the
compensation techniques are shown to improve the overall controller design.
system performance over the entire vehicle operating range. Even when a thruster is not moving through the water, its
Finally, the use of a nonlinear adaptive sliding controller is force-producing behavior is not simple. In the rest of this pa-
shown to preserve system performance even as the thruster per it will be shown that these types of thrusters act like non-
performance degrades or in the presence of significant uncer- linear filters that significantly influence closed-loop behavior
tainties. at low speeds and hover. This effect is largely responsible for
the notoriously poor performance of the automatic heading
II. A DYNAMIC
THRUSTER
MODEL
and depth servoes for most small underwater vehicles.
A . General Description of a Thruster Torque-controlled thrusters predominate in electrically
Most small-to-medium-sized underwater vehicles are pow- powered vehicles and will be considered exclusively in this
ered by electric motors driving propellers mounted in ducts. paper. Thrusters with speed-controlled motors have signifi-
Some of these vehicles utilize reduction gears between the cantly different dynamics, but are less common for several
motor and propeller, but direct drive is the most common. reasons. The steady-state relationship between torque and
In virtually all cases, the propeller is mounted in a duct or thrust is nearly linear, while the steady-state relationship be-
shroud which increases the static and dynamic efficiency of tween propeller speed and thrust is nonlinear. Additionally,
the thruster. the speed-controlled thruster requires a high-quality angular
The thruster unit from the Jason ROV (Fig. 3) will be con- speed sensor (typically much better than the sensors used to
sidered here. These thrusters utilize a 1/3 hp brushless dc commutate a brushless motor), and the thruster performance
motor contained in a pressure compensated oil-filled housing, is more sensitive to axial flow effects.
along with a brushless motor controller. Motor torque is com-
B . Lumped Parameter Model Development
manded by applying an analog voltage ( f 10 V) to the motor
controller. A typical thruster, as shown in Fig. 4, consists of a static
These motors have been utilized down to the full operating shroud and a propeller driven by a torque source ( 7 ) at angular
depth of Jason, 6000 m. They were also used on the Jason Jr. velocity (0).The thruster shroud has a cross-sectional area
vehicle which explored the Titanic and on the Robin ROV, ( A )and encloses a volume ( V). The ambient fluid has a density
which also operates to 6000 m. Similar motors have been ( p ) and a volumetric flowrate within the thruster ( Q ) .
used on the UNH Eave series of autonomous vehicles. Con- The model development is simplified by the following as-
ceptually, these thrusters resemble the electric units on most sumptions:
low-cost ROV’S. The energy stored is due solely to the kinetic energy of
These types of thrusters are not ideal force-producing ac- the fluid in the duct.
tuators. It is well known that thrusters are subject to serious The kinetic energy of the external ambient fluid is negli-
degradation due to axial- and cross-flow effects. Axial-flow gible.
YOERGER et al.: INFLUENCE OF THRUSTER DYNAMICS ON UNDERWATER VEHICLE BEHAVIOR 169

SHROUD

I : pV/A2
Fig. 5 . Bond graph representing thruster dynamics.

Fig. 4. Schematic of the thruster. The enclosed volume is V , the crosssec- oped:
tional area A , the angular velocity of the propeller is 0, and the applied
torque is r .
Thrust = yQ. (3)
Friction losses are negligible.
Assuming that the propeller does not cavitate, the volumet-
Ambient fluid is incompressible.
ric flowrate can be related to the thruster/propeller charac-
Fluid flow at the thruster intake and exhaust is parallel,
teristics and angular velocity R. The difference between the
one dimensional, and at ambient pressure. Rotational flow ef-
theoretical and actual advance per revolution of a propeller is
fects are ignored.
referred to as the slip and is typically expressed as a ratio U
Gravity effects are negligible.
The thruster is completely symmetric with respect to the
as follows:
flow direction. U =
PA - Q
The kinetic coenergy T * of the fluid in the thruster can be QPA
expressed as a state function of the volumetric flowrate Q: where p represents the axial distance traveled by the propeller
blades with each unit of rotation (1 rad) and is referred to as
T * ( Q )= i p V [:I2.
L A
the pitch,
It follows from the equation above that Q(R) can be ex-
pressed as
A generalized momentum can be defined as
Q =VPAR (4)
where v = 1 - U and is referred to as the propeller efficiency.
The above relation is that of an inertia (momentum related From (1)-(4), the following thruster dynamic state and out-
by a static constitutive law to the flow [4]) in bond-graph put equations are formed:
nomenclature with the effort variable r and flow variable Q.
r has units of momentum/area and is referred to as the pres-
sure momentum. Since the energy relations are linear, the
coenergy and energy have equal magnitudes [ 5 ] , and the ki- Thrust = yQ.
netic energy T can be expressed as a state function of the
pressure momentum I?: The thruster dynamic state and output equations can be rep-
resented topographically in the bond graph of Fig. 5.
A2 If we assume that the propeller efficiency ( v ) , pitch (p),
T ( r ) = -r2. and duct area ( A ) are constant, the thruster dynamic state and
2PV
output equations may be expressed with the propeller angular
A power balance yields the following pressure momentum re- velocity R as the thruster dynamic state variable:
lation;
dT A2 .
- = -rr = or - K Q
dt PV
Thrust = Apy2p2R101.
where Or represents the power input from the thruster pro-
peller, K represents the exiting kinetic energy per volume, Note that the steady-state thrust force is proportional to the
and represents the time rate of change of the pressure input torque, as mentioned earlier.
momentum. The normalized step response of the thruster model to
The exiting kinetic energy per volume K can be expressed torques of 2, 1/3, and 1/4 Nm is presented in Fig. 6,
as graphically emphasizing the nonlinear dynamics. The thruster
presents a more complex control problem than the linear first-
K = -A 2 r 2 - -
y2
order lag associated with most actuators. The thruster time
2pv2 - 2 p
response performance actually degrades as the magnitude of
where y = A r / V is the fluid momentum per volume within the input decreases. This dynamic behavior, when coupled
the thruster. with other dynamic nonidealities such as a pure delay, could
The thruster and ambient fluid are coupled through the con- result in the limit-cycle behavior associated with underwater
vected linear momentum, which is equal to the thrust devel- vehicle hover control.
170

8 80 -
! M)-

" * ' " "


1 1.5 2 25 3 3.5 4 45 5
FmpellerAngular Velonty Squared
Time (seconds)
Fig. 8. Experimental results show excellent correlation between the steady-
Fig. 6. The nonlinear nature of the thruster response is illustrated in this
set of normalized step responses. The larger the input, the more rapid the state thrust force and the absolute squared propeller speed.
dynamics of the thruster.

I I

vc (volrs)
Fig. 9. The steady-state absolute squared propeller speed is shown expen-
Fig. 7. The test setup allowed the thruster control voltage to be set from a mentally to be a linear function of the applied motor control voltage.
computer, while the thrust force, flow rate, and propeller speed were logged.
firm the model presented earlier and to determine the specific
C. Model Summary parameter values a,0, and Ct . Given the available measure-
Using an energy-based physical system approach, a lumped ments, it was possible to confirm that these corresponded to
parameter dynamic model of the thruster was developed. With reasonable values for the physical parameters such as the pro-
the propeller angular velocity Cl as the thruster dynamic state, peller efficiency and involved volume.
the model and the output equation take the form: Static tests provided values for Ct and the ratio @/CY.
These
tests were conducted for thrust in the forward direction only,
= 07 - (YfllRI ignoring the asymmetry of the thruster. Fig. 8 shows a plot
of output force versus the absolute squared propeller speed,
Thrust = CfCllCl;21
yielding a value for C t . Fig. 9 shows absolute squared pro-
where 7 is the input torque, (Y and 0 are constant model peller velocity as a function of applied motor control voltage,
parameters, and C f is a proportionality constant. which is proportional to torque. This curve yields the ratio
This model will be used subsequently as a dynamic struc- p / a . Both plots show excellent agreement to the model. The
ture of determining the constant model parameters, and ulti- identified values were:
mately in the design of controller systems to compensate for Ct = 0.022 Ns*
the thruster dynamics.
fila= 1.0. io3 v-l s-~.
III. MODELVERIFICATION
A dynamic test was conducted to determine the individual
The model was verified using a thruster mounted in a tank values of the parameters (Y and @. The thruster was com-
which was instrumented for output force, propeller speed, and manded with a random input signal passed through a filter
output flow velocity. As shown in Fig. 7, the thrust was mea- with a 1-hz cutoff. By measuring propeller speed as a func-
sured by a load cell, the propeller speed was measured using tion of the applied motor control voltage, least-squares values
an electro-optic device similar to a simple optical shaft en- of CY and 6 were determined:
coder but suitable for immersion, and the output flow velocity
CY = 0.037
was measured by an analog electromagnetic current meter.
A series of static and dynamic tests were conducted to con- fl = 42 V-' sV2.
YOERGER et al.: INFLUENCE OF THRUSTER DYNAMICS ON UNDERWATER VEHICLE BEHAVIOR 171

Nonlinear System

QuasilinearSystem

30

0 2 4 6 8
Time (sec)
10 12 14
-I
16 7 p I . : :

S+JlQl

Fig. 10. The measured (dashed line) and simulated thrust force (solid line) Fig. 11. Replacing the absolute squared nonlinearity with a describing func-
responses to a random input (dotted line) are shown. The plot shows good tion permits the sinusoidal steady-state response of the system to be analyzed.
agreement and emphasizes the phase lag exhibited by both the real thruster
and model.
a function of the input magnitude. Describing functions will be
used to approximate the behavior of the system by representing
Note that the ratio of (1.1 . lo3) for /3/afor the dynamic test
it as a first-order filter which is a function of the magnitude
matches the results of the static tests, even though the dynamic
and frequency of a steady input sinusoid. Describing functions
tests included flow in both directions.
are ideal for analyzing a limit cycle, as they apply directly to
Fig. 10 shows the recorded propeller speed data and the
steady-state sinusoidal conditions.
simulated response using the identified parameters.
The block diagram in Fig. 11 shows a thruster model which
The ratio of these two numbers (1.1 . lo3 V-' sc2) com-
substitutes a describing function [6] for the absolute-square
pared favorably to the statically determined value of 1.0 . lo3
nonlinearity. For this nonlinearity, the describing function has
V-' sc2). Likewise, a least-squares value for C , determined
a gain which is a function of only the magnitude of the input,
from the recorded dynamic output force and propeller speed
in this case the propeller speed R. However, the steady-state
yielded a value of 0.019 Ns2, compared to the value 0.022
sinusoidal relationship between the commanded thrust ( F e )
Ns2 determined from the static tests.
and output thrust force ( F , ) is that of a first-order filter for
As all the parameters in the model are based on the physical
which the break frequency is a function of both the input
parameters of the motor, propeller, and duct, it is possible to
magnitude and frequency.
check the experimentally determined values of CY, /3, and C ,
Given that the system is first order, the magnitude of the
to see if they make physical sense. The only two physical
input and output sinusoids are related by
parameters of the model that are not directly determinable are
the involved fluid volume V and the propeller efficiency r ] . IFt(j4l w
--
Given the experimentally determined parameters a,/3 , and C t
as well as the known physical parameters for propeller pitch
IFc(jw)I - (U2 x;h)'12 +
( p = 0.101 m), duct area ( A = 0.0527 m2), and fluid density where the break frequency x t h is a function of the propeller
( p = loo0 kg/m3), the values of Vand r] can be implied. The speed R,
computed value of is 0.22, which is reasonable for propellers
of this type, and the computed involved fluid volume is 0.016
m3, which is twice the actual volume enclosed by the duct.
In summary, the identification procedure provided good val- Combining these two equations, recalling that the output
ues for the parameters and also supported the basis for the force F t = CtRIRI, and solving a quadratic yields the break
model. Static and dynamic tests were consistent and the iden- frequency Xth as a function of the input magnitude IFc(jw)I
tified parameters made physical sense. and frequency w :

IV. EFFECT
OF THRUSTER
DYNAMICS
ON CLOSED-LOOP
BEHAVIOR
In this section it will be shown that the thruster dynamic
model can be used to explain several difficulties encountered in where
controlling many types of underwater vehicles. First, a strong
8a
constraint on the closed-loop bandwidth is encountered which y=-.
is much lower than the limit implied by sampling or time 3lr
delays. Also, the vehicle will exhibit a limit cycle about the Fig. 12 shows a contour plot of the break frequency x t h for
desired setpoint. a Jason thruster, using the identified values of a and /3 as a
A quasi-linear approximation of the dynamic model will be function of the input sinusoid magnitude and frequency w . The
used for this analysis. As seen in Fig. 6, the model can be break frequency is strongly dependent on the magnitude of the
approximated as a first-order system where the time constant is input sinusoid, and weakly dependent on the frequency of the
172 IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO. 3, JULY 1990

20

15

10

z 5
9P 0

1 -5

-10

-I5

-20
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 5 10 15 20 25 30 35 40 45 50

Input frequency (l/s) tune (S)

Fig. 12. The quasi-linear break frequency, as determined from the &crib- Fig. 14. This plot of simulated thrust is consistent with the movement of
ing function analysis, is a strong function of the magnitude of the input the closed-loop poles shown in Fig. 13. Additionally, the magnitude and
and a weaker function of the input frequency. frequency of the limit cycle are very close to the values predicted by the
describing function analysis.

Desired
Clmed-loop
describing function model can be used to compute approxi-
mate values for the magnitude and frequency.
The desired closed-loop system transfer function relates the
Open-Imp desired position Xd to the actual position X :
2
X(s> - - *n
(a) (b) (C) +
X d ( s ) s2 2 r ~ n +U,’
s
Fig. 13. (a) If no thruster dynamics are present, the open-loop poles are
easily placed to produce the desired closed-loop response. (b) The unac- where on is the desired natural frequency, and is the desired r
knowledged presence of the slow thruster pole results in the actual closed- damping ratio.
loop poles moving into the right half plane. (c) The resulting instability Unfortunately, if the thruster dynamics are ignored when
causes the thrust level to increase, which speeds up the thruster pole. The
thruster pole frequency increases until the pole pair reaches the j w axis to
the gains are chosen, the following third-order closed-loop
form a pure oscillator. transfer function is obtained for the system, including the
thruster pole:
input sinusoid. For the Jason thruster, the value of the thruster
break frequency is less than 2 for all magnitudes, which is
X-
- (s) - k-4
certain to wreak havoc with any effort to place the closed-loop X ~ ( S ) s3 +X ~ +S2r~thwns
~ +
poles at a reasonable bandwidth, unless specifically factored The value of Ath which puts the two complex poles on the
into the design. j w axis can be computed by substituting s = j w l c into the
Now the describing function model can be used to ana- characteristic equation and finding the roots, where utc is the
lyze the effect of the thruster dynamics on closed-loop per- frequency of the limit cycle. This calculation yields:
formance. In the simplest case, a single degree of freedom
w c =*n
of a vehicle can be described as a pure inertia (Fig. 13(a)).
Assume that the goal of the control system design is to place A* =wn/2{.
the closed-loop poles at a location specified by the natural
frequency Act with a damping ratio { = 0.707. Both position This means that the magnitude of the thruster input will in-
and velocity gains are required to achieve a damped response, crease until the break frequency of the thruster dynamics
and these can be computed algebraically. reaches a value of w n / 2 r . Recalling (3,
the magnitude of
However, unless the break frequency of the thruster dy- the commanded thruster force can be related to the break fre-
namics is much higher than the desired value of &I, the poles quency of the thruster and the thruster command frequency:
will not move as expected. With the system initially at rest,
the thruster action will be low, and the filter break frequency,
very low. The presence of the low-frequency pole associated
with the thruster dynamics will drive the two other poles into Fig. 14 shows that the limit cycle produced a numerical
the right half plane. The resulting instability will then increase simulation of a single axis of a vehicle the size of Jason Jr.,
the magnitude of the input to the thrusters until the filter break including the full nonlinear thruster dynamics. The magnitude
frequency increases to the point where the poles lie directly of the commanded thruster signal during the limit cycle is 15.6
on the j w axis. This process is illustrated in Fig. 13. N, which corresponds closely to the value of 15.7 N predicted
This situation describes a classic limit cycle. The resulting by (6). Likewise, the frequency of the limit cycle corresponds
oscillation will be of a specific and constant frequency and almost exactly tol the desired closed-loop frequency of 11s
magnitude that are independent of the initial conditions. The (period = 2n s).
YOERGER et al.: INFLUENCE OF THRUSTER DYNAMICS ON UNDERWATER VEHICLE BEHAVIOR 173

V. COMPENSATION
FOR THRUSTER
DYNAMICS 0.08 ,-.’
,on‘.*, \ compensaaon
, I

Several controller structures were investigated to overcome 0.07 ’ :


;
,,
,, ,lead compensator
the effects of neglecting the thruster dynamics. Recall that
there were two bad effects that resulted from neglecting the
thruster dynamics: First, overall system bandwidth was lim-
ited, and secondly, a limit cycle was produced.
A . Hybrid Simulation
A hybrid simulation was used to verify the various controller
designs. The hybrid simulation was an extension of the test
setup used to verify the thruster model (see Fig. 7). Rather a 01
than just logging the output of the thruster, the measured value a 02
2 4 6 8 10 12
was used to drive a real-time numerical simulation of a single
Tune (seconds)
degree of freedom vehicle. This permitted the behavior of the
Fig. 15. For a sinusoidal tracking task, the lead compensator improved
actual thruster and its effect on the control loop to be observed performance considerably over the uncompensated case, as shown in these
while acting on a “pure” vehicle with well-known character- hybrid simulation results.
istics. For the controllers that were implemented digitally, the
states of the vehicle simulation (xand i )were made available
to the controllers. Additionally, the adaptive sliding controller
required the propeller speed measurement 0. A more detailed
description of the hybrid simulation is given in [7].
An important limitation of the hybrid simulation is that the
thruster does not actually move when the simulated vehicle
velocity is nonzero. This means that axial-flow effects are ig-
nored. For this reason, the experiments concentrated on hov-
ering and small movements.
The vehicle simulation had the form:

MX = thrust - CoilXI
2 4 6 8 10 12
where M is the effective mass (actual mass plus “added
Time (seconds)
mass”), and CO is the drag parameter. The numerical val- Fig. 16. The effectiveness of the lead compensator diminished as the op-
ues used were chosen to represent a single DOF of Jason Jr., eration conditions change from those for which the lead compensator was
M = 340 kg, and CO = 67 Ns2/m2. designed. Here, for a low magnitude input, performance has degraded
significantly.
A vehicle position controller was implemented which com-
pensates for the vehicle drag with a nonlinear feed-forward
selected as representative of the dynamics to be compensated
term which would produce first-order tracking behavior:
is that observed at low control signals. From Fig. 6, we ob-
U = c~xlxl-2 G - X2P serve that a 0.75s time constant matches the observed step
response at a 1-V control signal. The lead compensator in-
where P = x - x d is the tracking error, U is the computed creases the system order to two, and we select our desired
control force, and X is the desired closed-loop bandwidth. dominant poles with an attenuation twice that of the first-order
This control law was used for all the control schemes that pole and a damping ratio of 0.707. Following the root locus
follow. However, in each case, a different type of compensa- method of [8], the lead compensation pole and zero are placed
tion was employed. These compensators operate on the control at - 6.5 and - 2.43, respectively. The resulting improvement
force U computed by the position controller (which is naive due to the addition of lead compensation on the closed-loop
to the existance of the thruster dynamics) to produce an im- system response is shown in Fig. 15.
proved overall system response. Fig. 16 shows the results for a trajectory which calls for
a much lower level of thrust (resembling hover). The closed-
B . Lead Compensation loop response is oscillatory and shows a phase lag not seen
Given the resemblence of the nonlinear thruster dynamics when the system is operating closer to the design point.
to a first-order system, it is sensible to compensate with a lead In summary, the lead compensator represents a substantial
network [8]. If the thruster actually behaved like a first-order improvement over no compensator, but performance degrades
linear system, a zero could be placed at the frequency of the substantially if the operating conditions vary greatly from the
thruster pole and a complementary pole placed at a higher design point.
frequency where it would not interfere with the controller.
Fig. 15 shows the results using a lead compensator com- C . ‘%le” Cancellation
pared to the results with no compensation. The lead com- The “pole” cancellation controller is so named because it
pensator was implemented in analog form. The time constant compensates for thruster dynamics by canceling the apparent
174 IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO. 3, JULY 1990

lag pole with a corresponding “zero.” The corresponding


control law which is implemented digitally is of the form:

-0.02‘ J
where U * is the compensated thrust command, and hu(u)is 2 4 6 8 10 12

the inverse of the thruster time constant which is chosen as


a function of the uncompensated control input U. Note that
although the term “pole” us used here loosely, the validity
B
01:
of this time-varying “pole-zero cancellation” can be justified 2 0.05
for this first-order system as long as Xu is bounded away from
P o
zero. 3
I
The pole cancellation method requires the evaluation of the 2 4 6 8 10 12
time derivative of the commanded control signal U . The value
of U can be determined explicitly if the commanded control
signal for the thruster is known a priori, as when there is
a predefined force trajectory for the associated vehicle; how-
ever, this is not normally the case. The value of U may also be
determined by differentiating the vehicle control law, resulting
in some terms that are known explicitly and some that may
2 4 6 8 10 12
be determined by numerical differentiation. For example, the
Fig. 17. The “pole” cancellation technique provides good performance over
vehicle control law is a wide range of operating conditions, as shown in these hybrid simulation
results.
U = -KpZ - K d Z
U = -Kpk - Kdi.
The velocity error k is known from the state measurement,
and the acceleration error 2 may be determined by numerical
differentiation with a known error bound taking the form of a
disturbance.
Xu(U) represents the location of the apparent first-order
thruster “pole” and is considered a function of the magnitude
of the commanded input; X,(u) was determined empirically
by matching the step response of a first-order system to the
observed thruster step response; X, (U ) was determined for the
full range of thruster operation and matched to the functional
form: Time (seconds)
Fig. 18. The “pole” cancellation technique suffers when the actual thruster
&(U) = 0.49fi. statics or dynamics differ significantly from the model. Here, a blockage
is placed in the flow for the hybrid simulation and control performance
In [7], it is shown that this relationship provides a good fit degrades significantly.
to the observed thruster behavior and that it is consistent with
the model identified earlier. D . Sliding Controller
Combining this functional form for X,(u) with the “pole”
Unlike the previous two compensators, a model-based com-
cancellation differential equation yields the differential equa-
pensator can directly take advantage of the previously identi-
tion for the “pole” compensation:
fied controller model. Given the perfect measurement of the
U thruster dynamic state and a perfect model, the input torque
u*=u+- can be directly computed and applied; however, this com-
0.49fi‘
puted torque method may quickly fail in the presence of
Fig. 17 shows the hybrid simulation results for the closed- model uncertainty, measurement noise, computational delays,
loop system with “pole” compensation. Unlike the lead com- and disturbances.
pensator, the nonlinear compensator provides uniform perfor- Analysis of the effects of these nonidealities are further
mance over the entire operating range. complicated by nonlinear dynamics. The issue becomes one
Fig. 18 illustrates a fundamental limitation of the “pole” of ensuring that a nonlinear dynamic system remains robust
compensation scheme. In this example, an obstruction was to nonidealities while minimizing tracking error.
placed in the thruster duct, altering the static and dynamic One method able to deal directly with system uncertain-
properties of the thruster. Clearly, the “pole” compensation ties is referred to as “Sliding Mode Control” and has re-
technique relies on good knowledge of the thruster properties. ceived considerable attention in the Soviet literature, where it
YOERGER e/ al.: INFLUENCE OF THRUSTER DYNAMICS ON UNDERWATER VEHICLE BEHAVIOR 175

is applied to the control of linear systems with discontinuous where


control action. Sliding control theory has been extended to 1, s>o
systems with nonlinear systems with continuous control ac- sign(s)
tion [9], and has been further extended to included adaptation -1, s<o
to system parametric uncertainty [101. Application of adap- and K is selected to be large enough to offset the model dy-
tive sliding mode theory permits precise tracking performance namic uncertainty, ensuring that the sliding condition remains
while being robust to modeling errors and disturbances. The satisfied.
controller design presented follows closely the development of Although the discontinuous control law provides for perfect
sliding controllers in [ 101 and [9]. tracking despite nonidealities, it exhibits undesirable chatter-
With the propeller angular velocity R as our thruster state ing around the sliding point. Chattering can be avoided by
measurement, the thruster dynamic model can be written as smoothing the control law within a thin boundary of the slid-
ing point [9]. This smoothing introduces behavior which de-
R = p7 - (YRIR1 parts from the dynamics specified by the sliding surface in
the form of a first-order filter in s when inside a region about
or
the sliding surface called the "boundary layer." This can be
hfi = 7 - c R ~ R ~ accomplished by replacing the discontinuous control law term
by the continuous term K sat (s/d), where
where 7 is the input torque, while a and /3 are the thruster sign(z), for 121 2 1
parameters.
A sliding surface s (in the example in this paper, a point)
is selected based on simple linear dynamics:
sat ( 2 )
{ 2, otherwise
and q5 is the boundary layer width.
Inside the boundary layer, the dynamics of s should take
s = = 0 - f&j
the form:
where fld is the desired angular velocity. Perfect tracking is S+Xs=bu
then achieved by the state trajectory remaining at the sliding
point (s = 0). where U is the input, b is the input gain, and X is the band-
Having selected the sliding point, the condition (referred to width.
as the sliding condition) necessary to constrain trajectories This can be verified by substituting the modified control law
to be directed toward the sliding point must be determined. inside the boundary layer into the dynamic equation for the
The sliding condition is satisfied if a quadratic form of s can sliding point:
be established as an Lyapunov-like function of the closed loop hS = 7 - CRlRl - h&
system [9]:
S
h i = fRlRl +hhd - hK- - CRlRl - h&
1 4
V = -hs2
2 S
hS = -hK- +Ya"
V = shS 5 0 . 4
Imposing the sliding condition on the dynamic model and where a = [c hIT,a" a^ -a is the parametric uncertainty, and
sliding point yields the control law: Y = [RIRI&].
The dynamics of s can be seen to be first order:
s=R
K
hS = h a = h ( d - Qj)
s + -s = h-lya".
4
and substituting the system dynamics: This implies that inside the boundary layer the bandwidth is
established by
hi =T - CRlRl - hfid
K
-
the best approximation of a continuous control law which 4
would achieve h i = 0 is
with s dynamics driven by the dynamic uncertainty, h-'YG.
? = fRlRl +h& The sliding controller with a boundary layer yields the con-
trol law:
where ( * ) indicates an estimate.
To satisfy the sliding condition despite model uncertainty,
a discontinuous control law is required. This can be achieved
by adding a term to the approximately continuous control law
E . Adaptive Sliding Controller
which is discontinuous across the sliding point:
The sliding controller design can be made adaptive to para-
7 = C R ( R I+ A { & -Ksign(s)} metric uncertainty by coupling it to on-line parameter estima-
176 IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO. 3, JULY 1990

tion which is based on the algebraic distance of the current 006

state to the boundary layer [lo]. This structure leads natu-


rally to active adaptation only when the system is outside the
boundary layer, avoiding the long-term drift frequently expe-
:I o,02

rienced in parameter estimation schemes [lo]. The developed


adaptive controller structure is based on the premise that there
p -0.02
2 4 6 10 12
be no adaptation to that which can be modeled, but adaptation
should occur only in response to complex dynamic effects that 0 15
cannot be simply modeled. Examples of appropriate triggers 3 o1
to the adaptation mechanism include thruster fouling, motor .$2 005
or propeller degradation, or complex flow effects.
Inside the boundary layer for steady-state conditions, (8) $ o
yields: 005
2 4 6 8 10 12

s = h. - 14
- Ya".
K
However, K was selected to be greater than the dynamic un-
certainty h-'Ya", which implies that s 5 4.
This indicates that once in the boundary layer, s remains
there. The value of s will be driven out of the boundary layer I
only when the actual dynamic uncertainty exceeds our cur- 2 4 6 8 10 12
Time (seconds)
rent estimate. Therefore we should only adapt when outside
Fig. 19. These hybrid simulation results show that the adaptive sliding
the boundary layer and in proportion to the distance to the controller performs very well over the entire operation range. Here, the
boundary layer. thruster parameters are initialized at their proper values and no adaptation
To ensure the stability of the adaptive closed-loop system, takes place.
the parameter estimation scheme must preserve the attractive-
ness of the boundary layer. With this in mind, a Lyapunov The model-based adaptive sliding controller was shown to
function candidate is selected of the form: avoid the limitations of the other controllers evaluated through
use of the hybrid simulation. Good vehicle-tracking perfor-
mance over the full range of thruster operation is obtained as
shown in Fig. 19. In addition, the adaptive sliding controller
and sliding condition: is robust to significant parametric uncertainty with the ability
to adapt on-line to changes in model parameters.
The ability of the contoller to adapt is best demonstrated by
where assuming that the model parameters are initially known. As
shown in [7], when parameter estimates were initially set to
zero, they converge rapidly and directly to the proper values.
Fig. 20 shows an example of parameter convergence for the

{ 0,
= s - 4 sign (s),
inside the boundary layer
outside the boundary layer
hybrid simulation for a sinusoidal-commanded position input
with a magnitude of 0.2 m and a frequency of 0.3 Hz. The
parameters do not converge exactly to the actual values, but
0, inside the boundary layer are estimated only well enough to achieve a tracking error
SA = within the specified boundary layer. Adaptation ceases within
S, outside the boundary layer
. . the boundary layer, preventing long-term drift.
a" = 6, since a 0 The adaptive sliding controller is also robust to changes
in model parameters as demonstrated by the tracking perfor-
and is a symmetric, positive definite matrix.
mance achieved when an obstruction is placed in the thruster
Imposing the sliding condition outside the boundary layer
duct outlet shown in Fig. 21. This is the same desired tra-
yields the adaptation recursion relation:
jectory and obstruction used in the evaluation of the "pole"
V = [sa { ~ a - " LK sign (s)) +Pro2 cancellation controller shown in Fig. 18. Performance after
imposing the obstruction ( t M 8 s) is not quite as good as
V = [ { S A Y +hTrp- S A ~ Ksign(s)l without the obstruction, but is far superior to the performance
of the "pole" cancellation compensator.
which satisfies the sliding condition:
S A Y + P r= o VI. CONCLUSION
The dynamics of underwater vehicles can be dominated by
the dynamics of the thrusters, particularly at low speed or
hover. Unless compensation for these dynamics is included in
/
YOERGER et al.: INFLUENCE OF THRUSTER DYNAMICS ON UNDERWATER VEHICLE BEHAVIOR 177

001
A quasi-linear analysis utilizing describing functions was
0009 performed to investigate the influence of the thruster dynam-
0008 ics on the closed-loop system behavior. The analysis reveals
WO7 that the thruster dynamics place a strong constraint on the
0006
closed-loop system bandwidth and produce a limit cycle. Fur-
thermore, the frequency and magnitude of the limit cycle can
0005
be precisely predicted.
WO4
Three forms of compensation were tested utilizing a hybrid
0003 simulation which combined an instrumented thruster with a
WO2 real-time mathematical vehicle model. The first compensator,
0001
a linear lead network, was easy to implement and greatly im-
0
proved performance over the uncompensated system, but did
5 10 15 20 25 30 not perform uniformly over the entire operating range. The
Tune (seconds) second compensator, which attempts to cancel the nonlinear
0 035 filtering effect of the thruster, was effective over the entire op-
erating range but depends on an accurate thruster model. The
0 03
final compensator, an adaptive sliding controller, was effec-
0.025
tive over the entire operating range and can compensate for
uncertainties or degradation of the thruster.
0 02

REFERENCES
0015
M. Abkowitz, Stability and Motion Control of Ocean Vehicles.
Cambridge, MA: MIT Press, 1969.
0 01
M. Gertler and G. R. Hagen, “Standard equations of motion for sub-
marine simulations,” Naval Ship R&D Center, Bethesda, MD, NSRDC
0 005 Rep. No. 2510, 1967.
J. Feldman, “Model investigations of stability and control character-
0 istics of a preliminary design of the deep submergence rescue vehicle
5 10 15 20 25 30
(DSRV scheme A),” NSRDC Rep., June 1966.
Tune (seconds) D. C. Karnopp and R. C. Rosenberg, System Dynamics: A Unified
Fig. 20. With the parameters initialized at zero, the adaptive sliding con- Approach. New York: Wiley, 1975.
troller shows rapid parameter convergence, after which adaptation ceases. S. H. Crandall, D. C. Karnopp, E. F. Kurtz, and D. C. Pridmore-
Brown, Dynamics of Mechanical and Electromechanical Systems.
New York: McGraw-Hill, 1968.
D. Graham and D. McRuer, Analysis of Nonlinear Control Systems.
New York: Wiley, 1961.
J. G. Cooke, “Incorporating thruster dynamics in the control of an
underwater vehicle,” Engineers degree thesis, MIT-WHO1 Joint Pro-
gram in Oceanogr. Eng., Woods Hole, MA, Sept. 1989.
K. Ogata, Modern Control Engineering. Englewood Cliffs, NJ:
Prentice-Hall, 1970.
J.-J. E. Slotine, “Sliding controller design for nonlinear systems,” Znt.
J. Contr., vol. 40, no. 2, 1984.
J.-J. E. Slotine and J. A. Coetsee, “Adaptive sliding controller syn-
thesis for nonlinear systems,” Znt. J. Contr., vol. 42, no. 6, 1986.
V. I. Utkin, Sliding Modes and their Application in Variable Struc-
ture Systems. Moscow: MIR, 1974.

I
5 10 15 20 25 30
Time (seconds)
Fig. 2 1. Unlike the “pole” cancellation technique shown in Fig. 18, the use
of an adaptive sliding controller allows the system to maintain performance
even when an obstruction is placed in the duct. Dana R. Yoerger (M’87) obtained the S.B., S.M.,
and Ph.D. degrees in mechanical engineering from
the Massachusetts Institute of Technology, Cam-
the control system, the closed-loop system will be limited in bridge.
bandwidth and prone to limit cycle. This behavior is seen in He is an Associate Scientist at the Woods Hole
Oceanographic Institution, Department of Applied
the heading and depth control of most commercial ROV’s, and Physics and Engineering. His interests are in the
also in the position control of advanced ROV’s and AUV’s. areas of underwater vehicles and manipulators. He
A lumped parameter model of a torque-controlled thruster has published papers on vehicle and tether dynam-
ics, the application of modem nonlinear and adap-
shows the dynamics to be nonlinear. The thruster represents a tive control techniques to underwater vehicles, su-
fairly sluggish filter, although the speed of response increases pervisor control methodologies, and underwater manipulator design and per-
with the magnitude of the thruster input. The lumped param- formance. He is a Principal Investigator in the design and implementation of
Argo/Jason, an advanced telerobotic system for scientific seafloor survey, and
eter model was identified in a test tank using an instrumented ABE, and autonomous vehicle for long-term monitoring of the deep ocean.
thruster from the full ocean depth ROV Jason. He has participated In numerous oceanographic cruises, including the discov-
178 IEEE JOURNAL OF OCEANIC ENGINEERING, VOL. 15, NO. 3, JULY 1990

ery of the Titanic, full-scale dynamic testing of the Argo system, and the Cooke is currently Officer-in-Charge of the U.S. Navy’s nuclear-powered
initial deep-ocean deployments of the Jason vehicle and manipulator. deep-submergence research vessel NR-1.
Dr. Yoerger is a member of the IEEE Oceanic Engineering Society
Administrative Committee and of the Editorial Advisory Board of the journal,
Mechatronics. *
* Jean-Jacques E. Slotine (S’82-M’83) received the
John G . Cooke completed his undergraduate stud- Ph.D. degree from the Massachusetts Institute of
ies at the U.S. Naval Academy, Annapolis, receiv- Technology (MIT), Cambridge, in 1983.
ing the Bachelor of Science degree in mechanical He is Associate Professor of Mechanical Engi-
engineering, and was commissioned in 1978. In neering at MIT, Director of the nonlinear Systems
1989 he received the degree of ocean engineer from Laboratory, Doherty Professor of Ocean Utiliza-
the Massachusetts Institute of TechnologylWoods tion and CO-Director and CO-Founder of the Cen-
Hole Oceanographic Institution Joint Program in ter for Information-Driven Mechanical Systems.
Oceanographic Engineering. His research focuses on applied nonlinear control,
He is a Lieutenant commander in the U.S. Navy. robotics, and learning systems. Professor Slotine is
He completed nuclear propulsion and submarine the co-author of the textbooks, Robust Analysis
training in 1980 and was assigned to the commis- and Control and Applied Nonlinear Control, based on graduate courses he
sioning crew of the USS Ohio ( G B N 726). He served asEngineering Officer developed at MIT. He is an Associate Editor of several professional journals
of the commissioning crew of the USS Georgia (SSBN 729) and a Naviga- and is a frequent Consultant to industry and the Woods Hole Oceanographic
tor and Operations Officer of the USS San Francisco (SSN 711). LCDR Institution.

You might also like