You are on page 1of 9

4004 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 55, NO.

11, NOVEMBER 2008

Sliding-Mode Tracking Control of Surface Vessels


Hashem Ashrafiuon, Member, IEEE, Kenneth R. Muske, Member, IEEE,
Lucas C. McNinch, Member, IEEE, and Reza A. Soltan, Member, IEEE

Abstract—A sliding-mode control law is presented and exper-


imentally implemented for trajectory tracking of underactuated
autonomous surface vessels. The control law is developed by in-
troducing a first-order sliding surface in terms of surge tracking
errors and a second-order surface in terms of lateral motion
tracking errors. The resulting sliding-mode control law guarantees
position tracking while the rotational motion remains bounded.
The experimental vessel is a small boat with two propellers in an
indoor pool. The position and orientation of the boat are measured
using a camera that detects two infrared diodes attached near the
front and back ends of the boat. A computer with a capture card
processes the camera image to determine the position, calculates
the control forces and their corresponding input voltages, and
sends the control signals to wireless receivers on the vessel using
Fig. 1. Planar model of a surface vessel with two propellers.
a wireless transmitter. Several experiments are performed where
the vessel accurately follows straight-line and circular trajectories.
pellers are presented using a nonlinear hydrodynamic damping
Index Terms—Camera feedback, sliding-mode control, under- model. Based on this model, an asymptotically stable trajec-
actuated surface vessels.
tory tracking sliding-mode control law is developed using two
sliding surfaces for calculation of the two propeller forces. The
I. I NTRODUCTION first sliding surface is a first-order surface in terms of the surge-
velocity tracking errors. The second surface is second order
A UTONOMOUS surface vessels with two actuator in-
puts are considered as underactuated mechanical systems
since they posses three degrees of freedom (DOF) when mod-
defined in terms of the vessel’s sway velocity and acceleration
tracking errors. It is assumed that only the absolute position
eled as a single planar rigid body. Position control of underactu- and orientation of the vessel are measured and available for
ated systems has received increased attention in the last decade feedback as is the case when a global positioning system is
with most of the research focusing on feedback linearization, used. Hence, the absolute velocities are numerically estimated,
backstepping, controlled Lagrangian, and sliding-mode control and surge and sway velocities are calculated through kinematic
methods. The application of underactuated control to ocean ve- relations between inertial and body-fixed reference frames.
hicles includes hovercraft [1], autonomous underwater vehicles The proposed sliding-mode control approach is implemented
[2]–[4], and surface vessels [5]–[22]. This paper may also be using a small-scale experimental system. The experimental
divided into set point [5]–[10] and trajectory tracking [11]–[22] setup includes a small boat with two light-emitting diodes
position control problems since the respective controllers are (LEDs) and two propellers in an indoor pool, camera, capture
very different. The only experimental work in the underactu- card, computer, data-acquisition card, wireless transmitters and
ated surface vessel control area, to our knowledge, has been receivers, and two dc motors with custom-designed planetary
presented for set point stabilization [10] and ship tracking [20]. gear trains. The motions of the two LEDs are captured and
In this paper, we address the trajectory tracking control filtered to determine the location and orientation of the vessel
problem of autonomous surface vessels. Equations of motion [22]. The motor voltages are estimated through interpolation of
representing the planar model of a 3-DOF vessel with two pro- the calculated control forces and converted to analog signals by
real-time workshop and the dSpace (www.dspaceinc.com) card
which is used for digital-to-analog signal conversion. The sig-
nals are sent to a wireless transmitter which, in turn, transmits
them to dc motor controller receivers. Several experiments are
Manuscript received January 15, 2008; revised August 20, 2008. Current performed where the vessel successfully follows straight-line
version published October 31, 2008. This work was supported in part by the
Office of Naval Research under ONR Grant N00014-04-1-0642 and in part and circular trajectories.
by the Center for Nonlinear Dynamics and Control (CENDAC) at Villanova
University.
The authors are with the Center for Nonlinear Dynamics and Control, II. M ODEL
Villanova University, Villanova, PA 19085 USA (e-mail: hashem.ashrafiuon@
villanova.edu). Fig. 1 shows the model of a planar surface vessel represent-
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. ing surge, sway, and yaw motions with two propeller force
Digital Object Identifier 10.1109/TIE.2008.2005933 inputs f1 and f2 . The geometrical relationship between the

0278-0046/$25.00 © 2008 IEEE


ASHRAFIUON et al.: SLIDING-MODE TRACKING CONTROL OF SURFACE VESSELS 4005

inertial reference frame and the vessel-based body-fixed frame errors


is defined in terms of velocities as
t
ẋ = vx cos θ − vy sin θ S1 = ṽx + λ1 ṽx (τ ) dτ, λ1 > 0 (5)
ẏ = vx sin θ + vy cos θ 0
θ̇ = ω (1)
where “∼” is used to denote the difference between the actual
where (x, y) denote the position of the center of mass, θ is the and desired values, i.e.,
orientation angle of the vessel in the inertial reference frame,
ṽx = vx − vxd . (6)
(vx , vy ) are the surge and sway velocities, respectively, and ω
denotes the angular velocities of the vessel in the body-fixed Note that the integral of vx is used since position variables can-
frame. not be defined in the body-fixed frame. Hence, the desired mo-
In the body-fixed frame, the nonlinear equations of motion tion is specified in the inertial reference frame (xd (t), yd (t)),
for a simplified model of the dynamics of a surface vessel, and its time derivatives (ẋd , ẏd , ẍd , ÿd ) are related to the de-
where the motions in heave, roll, and pitch are neglected, are sired surge and sway velocities (vxd , vyd ) and accelerations
given by (v̇xd , v̇yd ) as
m11 v̇x − m22 vy ω + d1 vxα1 = f
vxd = cos θẋd + sin θẏd
m22 v̇y + m11 vx ω + d2 sgn(vy )|vy |α2 = 0
m33 ω̇ + md vx vy + d3 sgn(ω)|ω|α3 = T. (2) vyd = −sin θẋd + cos θẏd (7)
Note that we only consider forward vessel motion in this v̇xd = cos θẍd + sin θÿd + vyd ω
paper (vx > 0) since the reverse motion dynamics can be quite
different. The general spatial equations for surface vessels, as v̇yd = −sin θẍd + cos θÿd − vxd ω. (8)
well as the planar equations, can be found in [23]. In (2), the
mass elements m22 = m11 due to the added mass effect and In (7) and (8), θ is the actual measured yaw angle provided
md = m22 − m11 > 0 as presented in [23]. The mass moment through feedback in real time, and ω is estimated in real time
of inertia element m33 also includes added mass effects. The based on the θ values.
hydrodynamic damping is modeled as power law with expo- We calculate a nominal surge control law for zero dynamics
nents α1 and α2 and coefficients d1 and d2 for surge and by taking the time derivative of the surface and using the first
sway motions, respectively, and exponent α3 and coefficient equation of motion in (2)
d3 for rotational motion. Assuming that the propellers are
symmetrically located apart at a lateral distance B, the surge Ṡ1 = v̇x + Ṡr1 = 0, Ṡr1 = −v̇xd + λ1 ṽx (9)
control force f and the yaw control moment T are given in
fˆ = − m
 22 vy ω + dˆ1 vxα1 − m
 11 Ṡr1 (10)
terms of propeller forces as

f = f2 + f1 T = (f2 − f1 )B/2. (3) where “∧ ” is used to indicate the estimated model parameters.
The sliding-mode control law is normally derived by subtract-
ing a sign function from the nominal control. However, there are
III. S LIDING M ODE C ONTROL L AW several disadvantages in using a discontinuous control law, such
In the sliding-mode control approach [24], [25], we define as chattering and the inability of the propellers to produce a
asymptotically stable surfaces (S) such that all system trajec- discontinuous control action. Hence, we define an approximate
tories converge to these surfaces in finite time referred to as the sliding-mode control law using a high-slope saturation function
reach time tr and slide along them until they reach the desired as in [26]
destination at their intersection. The reaching conditions are
normally established by defining (1/2)S T S as the Lyapunov f = fˆ − k1 sat(S1 /φ1 ) (11)
function and ensuring that its time derivative is negative. In the 
S1 /φ1 , if |S1 | ≤ φ1
case of underactuated surface vessels, we define two surfaces to sat(S1 /φ1 ) = (12)
sgn(S1 ), if |S1 | > φ1
determine the two control inputs. Hence, the reaching condition
for each surface i may be defined as where φ1 is a positive constant which defines a small boundary
layer around the surface.
Si Ṡi ≤ −ηi |Si |, ηi > 0; i = 1, 2 (4)
In order to determine k1 , let us define the following bounds
where the value of constant ηi determines how fast the trajec- for the model parameters in (2):
tory will reach surface i, i.e., tr ≤ |si |/ηi .
|mii − m
 ii | ≤ Mii |di − dˆi | ≤ Di , i = 1, 2, 3. (13)
A. Surge Control Law
We assume no uncertainty in our estimates of exponents αi
The first sliding surface is a first-order exponentially stable for simplicity. Note that our experiments have shown less than
surface defined in terms of the vessel’s surge motion tracking 1.5% uncertainty for these exponents as presented in Section V.
4006 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 55, NO. 11, NOVEMBER 2008

Define a Lyapunov candidate function similar to (4) that where “∧ ” indicates the estimated model and
guarantees reaching the set {|S1 | ≤ φ1 } in finite time and re-
main inside it thereafter b = m22 vxd − m11 vx (24)
α3
1 h = b (md vx vy + d3 sgn(ω)|ω| )
V1 = m11 S12 . (14)
2 − (m11 vx ω + d2 sgn(vy )|vy |α2 ) α2 d2 sgn(vy )|vy |α2 −1
The time derivative of (14) can be derived using (2), (9), + m33 ω (f − d1 vxα1 + 2λ2 m11 vx + m22 vy )
and (11)
  + 2λ2 m33 d2 sgn(vy )|vy |α2
f + m22 vy ω − d1 vxα1

V̇ = m11 S1 Ṡ1 = m11 S1 + Ṡr1 + m22 m33 vr + 2λ2 v̇yd − λ22 ṽy . (25)
m11
 We define the yaw moment sliding-mode control law as
= S1 (m22 − m  22 )vy ω + (dˆ1 − d1 )vxα1
 
 T = ĥ − k2 sat(S2 /φ2 ) /b̂. (26)
+ (m11 − m
 11 )Ṡr1 − k1 sat(S1 /φ1 ) . (15)
It is interesting to note that the lateral motion control depends
Note that the saturation function is equal to the sign function not only on the yaw moment but also on the surge force f
for the set {|S1 | > φ1 } according to (12). Hence, the following as one would expect. It must also be noted that the control
reaching condition is achieved: law becomes singular if b̂ = 0. Since the control law is valid
only when surge motion is required (nonzero surge velocities),
V̇1 = m11 S1 Ṡ1 ≤ −m
 11 η1 |S1 | (16) singularity could only occur when the tracking surge-velocity
error is more than 20%. This is due to the fact that m22 is
if k1 is selected as
normally more than 20% larger than m11 [23] (see, e.g., data
k1 = M22 |vy ω| + D1 vxα1 + M11 |Ṡr1 | + m
 11 η1 . (17) in Section V). Hence, we assume that b > 0.
In order to determine k2 , let us define a bound for the un-
certainty in h which, in turn, is defined based on the parameter
B. Lateral Motion Control Law uncertainties defined in (13)
We define the second sliding surface as a second-order expo- |h − ĥ| ≤ H. (27)
nentially stable surface in terms of the vessel’s lateral motion
tracking errors We also define the following bounds based on the geometric
mean of b [27]:
t
S2 = ṽ˙ y + 2λ2 ṽy + λ22 ṽy (τ ) dτ, λ2 > 0 (18)
1/β ≤ bb̂−1 ≤ β, b̂ = bmin bmax , β = bmax /bmin (28)
0
 22 + M22 )vxd − (m
bmax = (m  11 − M11 )vx > 0
where ṽy = vy − vyd and ṽ˙ = v̇y − v̇yd . The estimates of these  22 − M22 )vxd − (m
bmin = (m  11 + M11 )vx > 0. (29)
error states are available through feedback and (1), (2), (7),
and (8). We define another Lyapunov candidate function that guaran-
We calculate a nominal yaw control law for zero dynamics tees reaching the set {|S2 | ≤ φ2 } in finite time
by taking the time derivative of the surface and setting it equal
to zero 1
V2 = m22 m33 S22 . (30)
2
Ṡ2 = v̈y + v̈yd + 2λ2 (v̇y − v̇yd ) + λ22 ṽy = 0 (19)
Noting that the saturation function is equal to the sign function
where the time derivatives of the second equations in (2) and for the set {|S2 | > φ2 }, the time derivative of (30) can be
(8) yield derived using (19)–(26) as
α2 d2 sgn(vy )|vy |α2 −1 v̇y + m11 (v̇x ω + vx ω̇)  
v̈y = − (20) bT − h
m22 V̇2 = m22 m33 S2 Ṡ2 = m22 m33 S2
m22 m33
 
v̈yd = vr − vxd ω̇ (21) = S2 bb̂−1 ĥ − k2 sat(S2 /φ2 ) − ĥ
...
vr = (vyd ω − 2v̇xd )ω − sin θ x d + cos θ
...
yd . (22)
= S2 (bb̂−1 − 1)ĥ + ĥ − h − bb̂−1 k2 sat(S2 /φ2 ) . (31)
The nominal control is derived by substituting (2), (20)–(22)
into (19) Hence, the following reaching condition is achieved:

T = ĥ/b̂ (23) V̇2 = m22 m33 S2 Ṡ2 ≤ −m


 22 m
 33 η2 |S2 | (32)
ASHRAFIUON et al.: SLIDING-MODE TRACKING CONTROL OF SURFACE VESSELS 4007

if k2 is selected as

 22 m
k2 = β(H + m  33 η2 ) + (β − 1)|ĥ|. (33)

C. Stability Analysis
The surge force and yaw moment control laws in (11) and
(26) are derived based on the reaching conditions in (16)
and (32), respectively. Integration of these reaching conditions
verifies that the trajectory reaches the two corresponding sur-
faces in a finite time of less than (m11 /m  11 )(S1 (0)/η1 ) and
(m22 m33 /m  22 m
 33 )(S2 (0)/η2 ), respectively. Furthermore, the
two surfaces in (5) and (18) are asymptotically stable. Thus, the
trajectory exponentially slides to the origin at the intersection
of the two surfaces. Thus,
t t
ṽx → 0, ṽx (τ ) dτ → 0, ṽy → 0, ṽy (τ ) dτ → 0 Fig. 2. Vessel with dc motors, receivers, gearboxes, and couplers.

0 0
(34)
and the kinematic relations in (1) guarantee trajectory tracking
in the inertial reference frame based on the real-time feedback
of θ and ω.
We also propose that ω is bounded-input–bounded-output
stable. Let us define the Lyapunov candidate function
1
V3 = m33 ω 2 . (35)
2
By using (2), the time derivative of V3 may be written as

V̇3 = ω [T − md vx vy − d3 sgn(ω)|ω|α3 ]
= ω(T − md vx vy ) − d3 |ω|1+α3 (36)
V̇3 < 0 if d3 |ω|α3 > |T − md vx vy |. (37)

Clearly, if V3 is a decreasing function, |ω| is also a decreasing


function based on (35). Hence, |ω| is a decreasing function in
the set {|ω| > (|T − md vx vy |1/α3 /d3 )}, and since T , vx , and
vy are bounded, ω remains bounded.
It can also be shown that ω reaches equilibrium in the closed-
loop response for simple straight-line and circular trajectories
using the yaw equation of motion in (2)

m33 ω̇ + d3 sgn(ω)|ω|α3 = T − md vx vy

0, linear motion
T − m d vx vy = Fig. 3. Diagram of the experimental setup.
constant, circular motion.

Hence, after the initial transient period, the steady-state solution are only controllable in a high-speed range, we have reduced
will be zero and constant for straight-line and circular trajec- their speed at 1 : 16 ratio with custom-designed planetary gear
tories, respectively. trains. The motors were originally controlled using a joystick
and wireless receivers. However, we have modified the joystick
such that computed control signals can be transmitted through
IV. E XPERIMENTAL S ETUP
a data-acquisition card.
The experiments are performed with a 0.45-m-long 1.614-kg We used a digital black and white 640 × 480-resolution
boat in a 1.9-m × 2.6-m indoor pool (Fig. 2). A diagram of camera for our feedback measurements. The camera rate is
the experimental setup is shown in Fig. 3. Two dc motors are 30 frames/s. The camera is installed 1.83 m above the center
mounted to propeller shafts using couplings built in-house. The of the pool and captures the image of the whole pool area.
propellers are 0.07 m apart (B = 0.07 m). Since the dc motors Since the camera image is distorted and in the form of an image
4008 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 55, NO. 11, NOVEMBER 2008

Fig. 6. Forward and reverse propeller force applied voltage relations.


Fig. 4. Camera calibration points and their image matrix.

Fig. 7. Block diagram of implementation of the feedback control law.

lc = 0.178 m distance from the rear LED. The infrared images


are filtered out, and camera calibration interpolation functions
Fig. 5. Camera calibration measured versus estimated points. are used to find the position of the two LEDs, (x1 , y1 ) and
(x2 , y2 ). These data are then used to determine the vessel
matrix of index numbers, it must be calibrated. We placed a yaw angle, θ = tan−1 ((y2 − y1 )/(x2 − x1 )), and its absolute
board on the top of the pool and installed infrared LEDs at position, x = x1 + lc cos θ and y = y1 + lc sin θ.
known locations over the surface which provided a grid of 35 × Since an accurate model of the motor and propeller dynamics
52 LEDs. We then captured an image and filtered out the LED is not available, we determined a relationship between motor
locations on the image matrix. input voltage and propeller force. We applied several constant
The camera image matrix locations of all the LEDs and voltages to each motor and measured the equivalent propeller
their corresponding actual positions are shown in Fig. 4. Two- forces using a force gauge. As a result of these experiments, we
dimensional cubic interpolation functions were used to fit the found a simple parabolic relationship at voltages of larger than
image matrix. The result of position estimation for the image 0.2 V in positive direction and 0.3 V in the negative/reverse
matrix using the interpolation functions is shown in Fig. 5 rotation. In other words, the dead band of the motors was
where only some selected LEDs are shown for clarity. The between −0.3 and 0.2 V. Fig. 6 shows the relationship between
maximum calibration error was less than 0.004 m (4 mm) in x- motor input voltage and propeller resultant force in positive and
direction and 0.01 m (1 cm) in the y-direction. It is clear that the negative directions and their parabolic curve fits. Although the
calibration is very accurate for the available camera resolution. motors are not perfectly balanced, we assumed that the two
We installed the two LEDs on the vessel at the top of motor force–voltage relations are identical.
plastic columns such that they are at the same height of the The block diagram in Fig. 7 illustrates the implementation of
grid used for camera calibration. The two LEDs are installed the control law through camera data filtering and calibration,
l = 0.343 m apart on the front and rear of the centerline of velocity estimation using the camera position data feedback,
the vessel. The boat’s center of gravity is on the centerline at and propeller force–motor input voltage interpolation. Note that
ASHRAFIUON et al.: SLIDING-MODE TRACKING CONTROL OF SURFACE VESSELS 4009

Fig. 9. Trajectory tracking error for a typical straight-line experiment.


Fig. 8. Desired versus actual vessel path during three experiments.

the surge and sway velocities are estimated using the estimated
velocities in the inertial reference frame and the inverse of the
transformation equations presented in (1).

V. R ESULTS
We performed a series of vessel surge, sway, and yaw motion
experiments by applying various known constant forces and
moments through a pulley system to estimate the parameters
of the model presented in (2). These experiments are not the
subject of this paper and are presented in [28]. The model data
and their uncertainty in SI units for the experimental vessel are

m11 = 1.956 ± 0.019 d1 = 2.436 ± 0.023


m22 = 2.405 ± 0.117 d2 = 12.992 ± 0.297
m33 = 0.403 ± 0.0068 d3 = 0.0564 ± 0.00085 Fig. 10. Vessel’s heading angle during a typical straight-line experiment.

α1 = 1.510 ± 0.0075
Fig. 8 shows that the actual paths followed by the vessel are
α2 = 1.747 ± 0.013 very close to the desired straight-line paths in three different
experiments. The controller is very robust despite the error
α3 = 1.592 ± 0.0285.
in motor and camera calibrations. Although the path error
was very small, the vessel’s trajectory error in the x- and
y-directions was larger, as shown in Fig. 9. It seems that,
A. Straight-Line Trajectory
during the acceleration and cruise period, the boat falls behind
We conducted a series of straight-line trajectory control the trajectory. However, the vessel compensates during the
experiments. In each case, the vessel was commanded to follow deceleration phase, and the errors become smaller again. This
a straight line for 8 s starting from rest at one corner of the pool suggests that the interpolation process underestimates the input
and stopping near the opposite corner for approximately a 2-m voltage for the positive propeller motion, and hence, smaller
travel. We used a fifth-order polynomial to plan the trajectory in voltages result in slower vessel motion. Fig. 10 shows that the
the x-direction. The six boundary conditions used to determine heading angle during a typical experiment is not constant but
the polynomial coefficients were zero initial and final (after 8 s) remains bounded.
velocities and accelerations, and the initial and final position Fig. 11 shows the input voltages for the two motors during
based on the length of travel and initial heading angle. The a typical straight-line experiment and simulation. It can be
motion in the y-direction was defined as a linear function of the observed that the applied experimental voltages during the
x-direction based on the initial heading angle. The control law positive propeller motion are smaller. Another interesting phe-
parameters were selected as λ1 = λ2 = 0.5, η1 = η2 = 0.01, nomenon is that the vessel must maintain a constant heading
and φ1 = φ2 = 0.1. angle to go straight. However, after nearly 7 s, its heading
4010 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 55, NO. 11, NOVEMBER 2008

Fig. 13. Desired versus actual vessel trajectory during a circular experiment.
Fig. 11. Comparison of experimental and simulation control signals.

Fig. 14. Trajectory tracking error for the circular experiment.


Fig. 12. Desired versus actual vessel path during a circular experiment.
the left end of the circle (0.665 m, 1.27 m). Fig. 13 shows
the vessel’s horizontal and vertical time history closely follow-
has deviated about 15◦ . This explains why, after 7 s, one ing the desired trajectory. Fig. 14 shows the tracking errors
motor requires maximum reverse input voltage while the other converging and remaining bounded. Fig. 15 shows the vessel
requires no input and the heading angle deviation is somewhat orientation smoothly changing to allow circular tracking. Note
compensated for during the last second of the experiment. that the discontinuity is due to the inverse tangent calculation at
every 360◦ .
B. Circular Trajectory
We conducted a series of circular trajectory experiments as
C. Discussion of Experimental Error
well. In each case, the vessel was commanded to start from
its initial position at rest and reach and follow a target on a There are several sources of error in this experiment, such
circular trajectory with a constant speed. The selected control as minor camera calibration errors, lack of speed control at low
law parameters were λ1 = λ2 = 1, η1 = 0.1, η2 = 0.002, and voltages due to the dead band of the motors, imbalance between
φ1 = φ2 = 0.01. the performance of the two motors, and motor calibration
Fig. 12 shows the circular path followed successfully by the errors. The two propellers must rotate in opposite direction to
vessel during one of the experiments. In this case, the vessel provide good performance. However, the motor speed control
starts from rest at the initial position (0.407 m, 0.478 m). The hardware does not perform the same in each direction. There-
target (desired trajectory) is rotating clockwise at a constant fore, equal voltages do not produce the same motor speeds in
speed of π/6 rad/s around a circle with radius of 0.3 m and many of the experiments. This problem should be resolved in
center at (0.965 m, 1.27 m). The target’s initial position is at the future through the use of an encoder and speed control.
ASHRAFIUON et al.: SLIDING-MODE TRACKING CONTROL OF SURFACE VESSELS 4011

[6] T. Kim, T. Basar, and I.-J. Ha, “Asymptotic stabilization of an underactu-


ated surface vessel via logic-based control,” in Proc. Amer. Control Conf.,
2002, pp. 4678–4683.
[7] F. Mazenc, K. Pettersen, and H. Nijmeijer, “Global uniform asymptotic
stabilization of an underactuated surface vessel,” IEEE Trans. Autom.
Control, vol. 47, no. 10, pp. 1759–1762, Oct. 2002.
[8] D. Soro and R. Lozano, “Semi-global practical stabilization of an under-
actuated surface vessel via nested saturation controller,” in Proc. Amer.
Control Conf., Denver, CO, 2003, vol. 3, pp. 2006–2011.
[9] W. Dong and Y. Guo, “Global time-varying stabilization of underactuated
surface vessel,” IEEE Trans. Autom. Control, vol. 50, no. 6, pp. 859–864,
Jun. 2005.
[10] K. Y. Pettersen, F. Mazenc, and H. Nijmeijer, “Global uniform asymp-
totic stabilization of an underactuated surface vessel: Experimental re-
sults,” IEEE Trans. Control Syst. Technol., vol. 12, no. 6, pp. 891–903,
Nov. 2004.
[11] G. Indiveri, M. Aicardi, and G. Casalino, “Nonlinear time-invariant feed-
back control of an underactuated marine vehicle along a straight course,”
in Proc. 5th Conf. MCMC, Aalborg, Denmark, 2000, pp. 21–26.
[12] J. Godhavn, “Nonlinear tracking of underactuated surface vessels,” in
Proc. IEEE Conf. Decision Control, Kobe, Japan, 1996, pp. 975–980.
[13] K. Y. Pettersen and H. Nijmeijer, “Tracking control of an under-
Fig. 15. Vessel’s heading angle during the circular experiment. actuated surface vessel,” in Proc. IEEE Conf. Decision Control, 1998,
pp. 4561–4566.
Despite these sources of model mismatch, however, relatively [14] P. Encarnacao and A. Pascoal, “Combined trajectory tracking and path fol-
lowing: An application to the coordinated control of autonomous marine
good tracking control is still obtained due to the robustness of craft,” in Proc. 40th IEEE CDC, Orlando, FL, 2001, vol. 1, pp. 964–969.
the sliding-mode control approach presented in this paper. [15] R. Olfati-Saber, “Exponential -tracking and -stabilization of second-
order nonholonomic SE(2) vehicles using dynamic state feedback,” in
Proc. Amer. Control Conf., 2002, vol. 5, pp. 3961–3967.
VI. C ONCLUSION [16] A. P. Aguiar and J. P. Hespanha, “Position tracking of underactuated
vehicles,” in Proc. Amer. Control Conf., 2003, vol. 3, pp. 1988–1993.
A new sliding-mode control law for trajectory tracking [17] A. Behal, D. M. Dawson, W. E. Dixon, and Y. Fang, “Tracking and
control of autonomous surface vessels was presented, im- regulation control of an underactuated surface vessel with nonintegrable
plemented, and experimentally tested. The control law was dynamics,” IEEE Trans. Autom. Control, vol. 47, no. 3, pp. 495–500,
Mar. 2002.
developed using a first-order surface in terms of the surge [18] Z. Jiang, “Global tracking controller design for underactuated
tracking errors and a second-order surface in terms of the sway ships,” in Proc. IEEE Conf. Control Appl., Mexico City, Mexico,
tracking errors. The vessel’s absolute position and orientation 2001, pp. 978–983.
[19] G. J. Toussaint, T. Basar, and F. Bullo, “Tracking for nonlinear underac-
were measured using a camera. The motor input voltages were tuated surface vessels with generalized forces,” in Proc. IEEE Int. Conf.
estimated from the controller propeller forces and transmit- Control Appl., Anchorage, AK, 2000, vol. 1, pp. 355–360.
ted to the motors using wireless transmitters and receivers. [20] E. Lefeber, K. Y. Pettersen, and H. Nijmeijer, “Tracking control of an
underactuated ship,” IEEE Trans. Control Syst. Technol., vol. 11, no. 1,
Several straight-line and circular experiments were successfully pp. 52–61, Jan. 2003.
performed. Future work will include experiments with envi- [21] H. Ashrafiuon and P. Ren, “Sliding-mode tracking control of under-
ronmental disturbances, obstacle avoidance, and coordinated actuated surface vessels,” in Proc. ASME IMECE, 2005, pp. 11–16.
[22] H. Ashrafiuon and K. Muske, “Sliding-mode tracking control of surface
control of multiple vessels. vessels,” in Proc. Amer. Control Conf., 2008, pp. 556–561.
[23] T. I. Fossen, Guidance and Control of Ocean Vehicles. New York: Wiley,
1994, pp. 168–171.
ACKNOWLEDGMENT [24] V. I. Utkin, “Variable structure systems with sliding-modes,” IEEE Trans.
Autom. Control, vol. AC-22, no. 2, pp. 212–222, Apr. 1977.
The authors would like to thank the undergraduate students, [25] J. Y. Hung, W. Gao, and J. C. Hung, “Variable structure control: A survey,”
namely, T. Flynn, G. Haas, and R. McCloseky, of the College IEEE Trans. Ind. Electron., vol. 40, no. 1, pp. 2–22, Feb. 1993.
of Engineering at Villanova University for helping them to set [26] H. K. Khalil, Nonlinear Systems. Upper Saddle River, NJ: Prentice-Hall,
1996, pp. 552–579.
up the experiments and in troubleshooting. [27] J. E. Slotine and W. Li, Applied Nonlinear Control. Englewood Cliffs,
NJ: Prentice-Hall, 1991, pp. 276–307.
[28] K. Muske, H. Ashrafiuon, G. Haas, R. McCloseky, and T. Flynn, “Iden-
R EFERENCES tification of a control oriented nonlinear dynamic USV model,” in Proc.
[1] I. Fantoni, R. Lozano, F. Mazenc, and K. Y. Pettersen, “Stabilization Amer. Control Conf., 2008, pp. 562–567.
of a nonlinear underactuated hovercraft,” in Proc. IEEE Conf. Decision
Control, 1999, vol. 3, pp. 2533–2538.
[2] N. E. Leonard, “Control synthesis and adaptation for an underactuated
autonomous underwater vehicle,” IEEE J. Ocean. Eng., vol. 20, no. 3, Hashem Ashrafiuon (M’03) received the B.S.,
pp. 211–220, Jul. 1995. M.S., and Ph.D. degrees in mechanical engineering
[3] K. Y. Pettersen and O. Egeland, “Time-varying exponential stabilization from the State University of New York, Buffalo, in
of the position and attitude of an underactuated autonomous underwa- 1982, 1984, and 1988, respectively.
ter vehicle,” IEEE Trans. Autom. Control, vol. 44, no. 1, pp. 112–115, Since 1988, he has been a member of the faculty of
Jan. 1999. the College of Engineering at Villanova University,
[4] A. P. Aguiar and A. M. Pascoal, “Dynamic positioning and way-point Villanova, PA, where he currently holds the position
tracking of underactuated AUVs in the presence of ocean currents,” in of Professor of mechanical engineering. He is also
Proc. 41st IEEE Conf. Decision Control, 2002, vol. 2, pp. 2105–2110. with the Center for Nonlinear Dynamics and Control,
[5] M. Reyhanoglu, “Exponential stabilization of an underactuated au- Villanova University. His research interests include
tonomous surface vessel,” Automatica, vol. 33, no. 12, pp. 2249–2254, sliding-mode control, underactuated systems, and
Dec. 1997. robotic systems.
4012 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 55, NO. 11, NOVEMBER 2008

Kenneth R. Muske (M’05) received the B.S. Reza A. Soltan (M’08) received the B.S. degree in
and M.S. degrees from Northwestern University, mechanical engineering from K. N. T. University of
Evanston, IL, in 1981, and the Ph.D. degree from the Technology, Tehran, Iran, in 2006. He is currently
University of Texas, Austin, in 1994. working toward the M.S. degree in mechanical engi-
He was a Process Control System Consultant neering at Villanova University, Villanova, PA.
prior to his Ph.D. studies and was a Technical Staff He is also with the Center for Nonlinear Dynamics
Member with Los Alamos National Laboratory af- and Control, Villanova University. His current re-
ter receiving the Ph.D. degree. Since 1997, he has search interests are in the areas of dynamic systems,
been with the College of Engineering, Villanova nonlinear control, limit cycles, and stability theory.
University, Villanova, PA, where he is currently the
Mr. and Mrs. Robert F. Moritz Sr. Chair of Systems
Engineering and a Professor of chemical engineering. He is also with the Center
for Nonlinear Dynamics and Control, Villanova University. His current research
interests are in the areas of system modeling, control, and optimization with
application to automotive, chemical, and autonomous systems.

Lucas C. McNinch (M’08) received the B.S. degree


in mechanical engineering from York College of
Pennsylvania, York, in 2007. He is currently working
toward the M.S. degree in mechanical engineering at
Villanova University, Villanova, PA.
He is also with the Center for Nonlinear Dynamics
and Control, Villanova University. His current re-
search interests are in the areas of optimal and model
predictive control and automated systems.

You might also like