Vous êtes sur la page 1sur 19

Transactions of the Institute of

Measurement and Control


http://tim.sagepub.com/

Lateral track control of UAVs using the sliding mode approach: from design to flight testing
M Zamurad Shah, Raza Samar and Aamer I Bhatti
Transactions of the Institute of Measurement and Control published online 24 July 2014
DOI: 10.1177/0142331214543093
The online version of this article can be found at:
http://tim.sagepub.com/content/early/2014/07/24/0142331214543093.1

Published by:
http://www.sagepublications.com

On behalf of:

The Institute of Measurement and Control

Additional services and information for Transactions of the Institute of Measurement and Control can be found at:
Email Alerts: http://tim.sagepub.com/cgi/alerts
Subscriptions: http://tim.sagepub.com/subscriptions
Reprints: http://www.sagepub.com/journalsReprints.nav
Permissions: http://www.sagepub.com/journalsPermissions.nav
Citations: http://tim.sagepub.com/content/early/2014/07/24/0142331214543093.1.refs.html

>> OnlineFirst Version of Record - Jul 31, 2014


OnlineFirst Version of Record - Jul 24, 2014
What is This?

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Article

Lateral track control of UAVs using the


sliding mode approach: from design to
flight testing

Transactions of the Institute of


Measurement and Control
118
The Author(s) 2014
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0142331214543093
tim.sagepub.com

M Zamurad Shah, Raza Samar and Aamer I Bhatti

Abstract
This paper develops sliding-mode-based nonlinear logic for guidance of unmanned aerial vehicles (UAVs) for curved and straight path following. UAV
trajectories generally consist of straight path segments, curved arcs, circular loiters and other manoeuvres; tight ground track control is desired
throughout the trajectory. This is achieved by controlling the lateral (cross-track) deviation of the vehicle in flight. The main objective of the guidance
algorithm is to keep the lateral track error of the vehicle as small as possible while performing graceful and stable manoeuvres despite the presence of
uncertainties and disturbing winds. Lateral track control is usually achieved by banking the vehicle, that is, by executing roll manoeuvres. The scheme
must perform well without saturating the roll angle of the vehicle, which serves as the control input for the guidance algorithm. The algorithm proposed here is shown to perform well for both straight and circular path tracking while ensuring control boundedness, and hence no saturation.
Crosswinds are a major source of disturbance for the guidance problem. This is incorporated into the design formulation and guidance gains are
selected to provide the desired performance despite the presence of disturbing winds. A sliding-mode-based scheme is developed which includes a
feedforward component related to the rate of change of the desired path heading. Stability of the algorithm is proved using an appropriate Lyapunov
function. The algorithm is implemented in the flight control computer of a scaled YAK-54 research aircraft; flight test results are presented and compared with those from other guidance algorithms. Flight results demonstrate the effectiveness and performance of the proposed guidance scheme. The
algorithm considers guidance in the 2D lateral plane only and minimizes deviations from the desired ground track of the vehicle.

Keywords
UAV guidance, sliding mode control, unmanned aerial vehicles, lateral guidance, track control, control boundedness, guidance of UAVs in the presence
of winds.

Introduction
In recent years, the use of unmanned aerial vehicles (UAVs)
has increased significantly in military, commercial and civilian applications due to their low cost, low human risk and
operational effectiveness. UAVs offer a unique range of features, such as ultra-long endurance, demanding trajectoryfollowing and high-risk mission acceptance, which cannot be
easily performed by manned aircraft. This has been made
possible due to the technological advancements of the last
two decades; UAVs have now obtained a permanent and critical role in high-tech military arsenals. Beside the military, the
use of UAVs has risen sharply in civilian and commercial
applications as well. Such applications include surveillance
and law enforcement, search and rescue, environmental studies, mapping and surveying, media and traffic reporting, forest fire monitoring and control, gas/oil/water pipeline and
power line monitoring, agriculture growth patterns monitoring and field spraying/crop dusting, radio/communications
relays and disaster surveying and management operations. In
all these applications, the requirement of closely following the
given ground track, despite complex mission trajectories,
remains a fundamental requirement. For example, in the
applications of agricultural field spraying, forest fire control,

rescue operations, border monitoring, etc., the need for tight


lateral track control is obvious.
The guidance and control design problems for
unmanned aircraft are usually treated separately, and implemented in an inner-loop, outer-loop configuration. The guidance law resides in the outer loop, uses ground track
measurements and generates appropriate commands for the
inner loop to follow. These commands can be in the form of
reference roll angle or reference lateral acceleration commands. The inner loop consists of a tracking controller that
accepts the reference commands generated by the guidance
loop, and performs control and stability augmentation tasks.
This approach to design, being simple and intuitive, can be
found in the literature in various papers; see for example
Dadkhah and Mettler (2012), Niculescu (2001), Rysdyk
(2006), Samar et al. (2008) and Shah et al. (2011a). Design of
the outer loop guidance logic has been inspired by techniques
Mohammad Ali Jinnah University, Islamabad, Pakistan
Corresponding author:
M Zamurad Shah, Department of Electronic Engineering, Mohammad Ali
Jinnah University, Near Kakpul, Sihala, Islamabad, Pakistan.
Email: m.zamurad.2013@ieee.org

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Transactions of the Institute of Measurement and Control

as diverse as proportional navigation, cognitive and intelligent methods, vision-based techniques, and others (Cesetti et
al., 2009; Ren and Beard, 2004; Siouris, 2004; Yamasaki and
Balakrishnan, 2010; Zhang et al., 2013). The objective
remains to drive the lateral ground track deviation of the
vehicle to a minimum. The inner loop control design problem
is beyond the scope of this paper; numerous well-established
techniques for flight control design exist such as linear robust
design and nonlinear and intelligent control design techniques. The reader may refer to any text on advanced control
for further reading.
Some work has also been done on the so-called integrated
approach in which the guidance and control problems are
dealt with in an integrated and unified framework (Dadkhah
and Mettler, 2012; Park et al., 2004; Yamasaki and
Balakrishnan, 2010; Yu et al., 2011). This approach is more
complicated due to coupling of the different guidance and
control variables and needs more research (Kaminer et al.,
1998; Yamasaki et al., 2012a,b). Most of the practising aerospace community favours the two-step approach, and that is
what is considered in this paper.
A tracking guidance law for a pursuer UAV is considered
in Regina and Zanzi (2011). The UAV tracks a moving target
and the guidance law generates lateral acceleration commands
which the control system follows. A back-stepping-based guidance law is proposed in Ahmed and Subbarao (2010); firstorder dynamics are assumed for elevation and heading angles.
Design of a path-following controller for a small UAV is presented in Sasongko et al. (2011); the cross-track and course
angle errors are used to generate heading angle corrections.
Heading change is accomplished using bank-to-turn manoeuvres; a linear quadratic regulator (LQR)-based controller
provides inner-loop roll control. Simulation results are presented to demonstrate the working of the system. Guidance
based on the concept of vector fields for curved path following has also been studied (Griffiths, 2006; Pisano et al., 2007).
Vehicle position relative to the desired path is computed, and
a vector field of heading commands is generated to drive the
error to zero. This may, in some cases, generate large heading
commands thereby saturating the control system of the
vehicle.
Proportional-derivative (PD) logic for outer loop guidance
has been employed in some UAV applications (Pappoullis,
1994; Siouris, 2004). Modifications to the PD scheme for disturbance rejection and saturation avoidance are proposed in
Samar et al. (2007, 2008); however, no formal stability proofs
are provided. A nonlinear guidance algorithm is proposed in
Park et al. (2004), and its stability discussed in Deyst et al.
(2005); the performance is better than the PD scheme for both
straight and curved path following. When the track errors
become large, however, the control input saturates and stability is not guaranteed. Also, the algorithm needs a reference
point for computation, which is a point from the mission
plan ahead of the current position of the vehicle; if the mission gets changed online, it can give rise to a discontinuity.
For a more complete survey of different approaches to UAV
guidance, the reader is referred to references Dadkhah and
Mettler (2012) and Goerzen et al. (2009).
Crosswinds are a major source of disturbance for accurate
lateral track control. Wind effect is ignored by most authors

in the design of the guidance logic. UAV dynamics in the presence of winds is discussed in references Ceccarelli et al. (2007),
McGee and Hedrick (2006) and Osborne and Rysdyk (2005);
however, the emphasis here is not on guidance but on how to
modify the planned mission to best cater for wind conditions.
Beard and McLain (2012) discuss the effect of wind on guidance performance of UAVs and aircraft. In Brezoescu et al.
(2013) an adaptive backstepping-based guidance law (considering the effect of winds) for UAVs is presented using a skidto-turn concept.
Direct application of sliding mode control in the outer guidance loop is normally not feasible; interested readers are
referred to Shtessel et al. (2003, 2007) for a detailed discussion. In the authors previous work (Shah et al., 2011b, 2014),
a novel nonlinear sliding surface is proposed for lateral guidance of UAVs. The main theme of that work was a highperformance sliding surface and algorithm for simple straight
paths; the effect of wind was not considered. Derivation of the
algorithm was based on the assumption that the reference
course angle will not change rapidly (i.e. no sharp circular or
curved paths). Flight tests for circular and curved paths in the
presence of wind indicated a need for performance improvement, and hence an extension of that work is required from
an applications point of view. Subsequently assumption of no
rapid change in reference course angle was relaxed in Shah et
al. (2012) and theoretical aspects along with simulation results
discussed there. In this work, a generalized nonlinear guidance
law is presented for both straight and curved path following
in the presence of winds. Performance is shown to be quite
good both for small and large track errors. Flight test results
with the proposed algorithm are also presented; these demonstrate excellent performance for both straight and rapidly
turning paths even in windy conditions. The work presented
considers guidance in the 2D horizontal plane only (lateral
guidance): this can be extended to the more general 3D guidance problem which is the subject of ongoing research.
This paper is organized as follows. Problem formulation
describes the problem under consideration in relative detail.
We define the reference coordinate system and different parameters relevant for circular and straight path segments, state
assumptions and develop state equations for guidance design.
In Sliding surface for cross-track control, a brief review of
previous work regarding nonlinear sliding mode guidance is
given; this forms the basis for further development of the guidance algorithm. Guidance law design describes design of a
nonlinear guidance law for straight and circular path following. Stability and control boundedness proofs for the proposed
law are also discussed. The guidance law is implemented on a
test aircraft and various flight tests carried out. Flight results
show efficacy of the proposed law; these are presented in
Flight test results. Comparison with previous algorithms is
also made. The final section concludes the paper.

Problem formulation
Path planning for UAVs is an active area of research. The
output of a path planning algorithm is in general a sequence
of waypoints joined by a curved path that the UAV should
follow. The curved path can be a combination of straight lines

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Shah et al.

WP-2

WP-1

Figure 2. The guidance problem definition for a straight path.

Centre of turn (O)

Figure 1. Definition of various angles.

rn
f tu
so
diu

Ra

and circular arcs of different radii. The guidance algorithm


generates suitable commands for the vehicle control system so
that the desired path is followed in the presence of disturbances. In this paper, we consider the lateral track control (or
guidance) problem of UAVs during straight as well as circular
paths in the presence of winds.
Before proceeding to the guidance problem, we first define
some important parameters as shown in Figure 1. The body
heading angle c is the angle which the nose of the aircraft
(usually referred to as the body x -axis) makes with respect to
~a denotes the velocity of the airnorth. The airspeed vector V
craft relative to air, and is not necessary aligned with the body
axis. The difference between the body heading and the relative
air velocity vector angle is called the sideslip angle b. In steady
~a is aligned
level flight without crosswind, the airspeed vector V
with the nose of the aircraft, implying zero sideslip angle.
Another important parameter is the ground velocity vector
~g , which gives the direction of travel of the vehicle relative to
V
~g is equal to V
~a .
ground. In the absence of wind, the vector V
~
~
In the presence of wind, Vg is the vector sum of Va and the
~w , as shown in Figure 1. The course
wind velocity vector V
~g with respect to
angle cG is the angle of the ground velocity V
north. Generally in an innerouter loop design strategy, kinematics is considered in the outer guidance loop and dynamics
is considered in the design of the inner control loop. The
dynamical variables respond much more quickly in time scale,
and hence the inner control loop is usually kept much faster
than the outer guidance loop. The sideslip angle is related to
the dynamics of the aircraft and must be considered in the
design of the inner control loop. The course angle cG is a
kinematic variable, and considered in guidance loop design.
Different guidance variables and their sign conventions are
shown in Figures 2 and 3 for straight and circular paths,
respectively. Both figures depict the positive sense of all indicated variables. Consider a straight path segment (Figure 2),
let WP1 and WP2 be two successive waypoints, and let cR be
the angle of the line WP1 WP2 with respect to north (called
the reference or desired course angle). The lateral (or crosstrack) deviation of the vehicle from the desired path is
denoted by y. In the case of a curved path (Figure 3), let WP1
and WP2 be two successive waypoints on a circular arc of

(R

WP-1

WP-1
P
y

Figure 3. The guidance problem definition for a circular path.

radius R centred at point O. Point P is the nearest point to


the vehicle on the arc. The reference or desired course angle
cR is defined as the angle of the tangent line at P with respect
to north. The cross-track y is defined as in the case of the
straight path. Another parameter of interest is cE =cG 2 cR,
usually referred to as the intercept course. The main task of
the guidance algorithm is to keep the cross-track error y as
small as possible, and also to keep cE 0 when y 0. In
the case of a non-zero y, the guidance algorithm should
manipulate cE by banking the vehicle and bring y to zero.
Generally for very large track errors, a constant cE (  p/2)
is desired, and when the track error reduces, cE is adjusted
accordingly (a criterion known as the good helmsman in literature; Pettersen and Lefeber, 2001; Rysdyk, 2006).
The block diagram of the overall guidance and control
system is shown in Figure 4. The outer guidance block gets
mission information from the mission plan in terms of waypoints, and feedback from sensors measuring the instantaneous position and course (cG) of the vehicle. The guidance
algorithm generates a roll angle command (fr) for the inner
control loop to track; the inner loop generates commands for
the control surfaces, that is, the ailerons and the rudder. Here
we shall focus on the design of the outer guidance logic that
ensures minimum cross-track error y for both straight and
curved paths by generating appropriate roll angle commands
in the presence of disturbing winds. This will be done using
the sliding mode approach.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Transactions of the Institute of Measurement and Control


35
Current position and heading angle (G)

34

North wind (5 m/s)


East wind (5 m/s)

33
Aileron

Control

+-

Rudder

Waypoints,
R and R

32

UAV lateral
dynamics

Vg [m/s]

Ref. roll angle (r)


Guidance

Yaw rate

Washout filter

Mission
Roll angle ()

31
30
29
28
27

Figure 4. The guidance and control scheme.

26
25

Assumptions

50

100

150

200
G []

250

300

350

400

We shall make the following simplifying assumptions:


Figure 5. Variation of Vg for constant Va in the presence of wind.

1. It is assumed that a control law for the inner loop is


available. The inner loop controller may be designed
using well-established techniques, such as robust linear
control design methods (Skogestad and Postlethwaite,
2005).
2. We also assume that the inner loop dynamics from fr
to f are significantly faster (e.g. 5 ; 10 times faster)
than the outer (guidance) loop dynamics.
3. The states y and cE (equation (18)) are assumed to be
perfectly measurable.

System dynamics
~g is the vector sum
As discussed earlier, the ground velocity V
~a and the wind velocity vector V
~w , in
of the airspeed vector V
~g = V
~a + V
~w . Considering a fixed and level
other words, V
northeastdown inertial coordinate system, the wind triangle
can be expressed as (Beard and McLain, 2012)
0

where Vg and Va are the magnitudes of the ground and airspeed vectors, respectively, and g and g a denote the inertial
referenced and air-mass-referenced flight-path angles respectively, in the longitudinal plane. For level flight (g = 0) the
projection of these vectors on the northeast (2D ground)
plane is written as


cos cG
sin cG


=

Vwn
Vwe

cos c
sin c

+ Va


2

or alternatively

Vg

cos cG
sin cG




Vwn
Vwe


= Va

Vwn sin cG  Vwe cos cG = Va  sin cG cos c + cos cG sin c

cos cG cos g
cos c cos g a
Vwn
Vg @ sin cG cos g A = @ Vwe A + Va @ sin c cos g a A 1
 sin g
Vwd
 sin ga

Vg

This equation gives the relation between Vg and Va that


depends on the magnitude and direction of wind and the
course angle cG. For UAVs usually a constant airspeed Va is
maintained at a given altitude either through a closed-loop
speed controller, or using a preset open-loop throttle-setting
table. In the presence of winds, a constant Va can still be
approximately maintained, but Vg will vary depending on the
course angle cG and the direction and magnitude of the wind.
As an example consider a constant airspeed of 30 m/s: the
variation of Vg with cG is shown in Figure 5 for a north wind
~w = (5, 0)T ), and also for an east wind of same
of 5 m/s (V
~w = (0, 5)T ). The relation between c and cG can
magnitude (V
also be derived using (3); multiplying both sides of the equation on the left-hand side by the row vector ( 2sin cG,cos cG)
yields

cos c
sin c


3

Taking the square of equation (3), we have


Vg2  2Vg (Vwn cos cG + Vwe sin cG ) + Vw2  Va2 = 0

5
which can be written in simplified form as
c  cG = sin1


1
Vwn sin cG  Vwe cos cG
Va

The concept of coordinated turns in the presence of winds


is discussed in detail in Beard and McLain (2012). The term
coordinated turn is used in the sense that there is no side
force in the body frame of the vehicle, thus implying zero
sideslip angle. Aerial vehicles use a component of the aerodynamic lift to generate lateral accelerations to correct the lateral (cross-track) errors during flight; these are balanced by
centrifugal accelerations in a coordinated turn. Lateral acceleration is produced by banking the vehicle so that a component of the lift vector is tilted in the required direction (the
bank-to-turn manoeuvre). The guidance loop generates
appropriate bank (or roll) commands fr for the inner control
loop to follow: fr therefore serves as the control input in our
case. Consider a coordinated turn (b = 0) in the presence of
wind: the aerodynamic lift vector L (Figure 6) can be resolved
into two components as shown in Figure 6. The component

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Shah et al.

5
Referring to Assumption 2 of Assumptions, where we
assume a significantly faster inner loop, we have fr f, and
so (9) can be approximated as

L cos

Vg c_ G
g

10

Vg (c_ E + c_ R )
g

11

tan fr cos (cG  c) =


Now since cE =cG 2 cR we have

tan fr cos (cG  c) =


Rearranging (11), we get

g tan fr
c_ E =
cos (cG  c)  c_ R
Vg

(a) Back view during coordinated turn

12

Using (6), the above state equation becomes

g tan fr
c_ E =
Vg



1
cos sin1
Vwn sin cG + Vwe cos cG
 c_ R
Va

13

where cE is a state variable and c_ R is the rate of change of


the desired heading: this information is available from mission data. During straight flight, c_ R is equal to zero, but it
can become appreciable during sharp turns.
The inertial position (Pn, Pe, Pd)T of the UAV in the fixed
northeastdown frame can be derived using

cos

1
0
1 0
1
Vwn
cos c cos ga
P_ n
@ P_ e A = Va @ sin c cos g a A + @ Vwe A
V ed
sin g a
P_ d

Centre of turn O
(b) Top view during coordinated turn

14

provided the airspeed and wind information is available.


Alternatively, we can derive this from the ground velocity
vector directly:

Figure 6. Components of the lift vector L during a steady turn of


radius R.

0
1
1
cos cG cos g
P_ n
@ P_ e A = Vg @ sin cG cos g A
sin g
P_ d
0

(L cos f) balances the weight of the vehicle, and a part of the


other component balances the centrifugal force during a turn
(Beard and McLain, 2012). Summing the forces in the vertical
and radial directions:
L cos f = mg,

L sin f cos (cG  c) =

mVg2
R

Here m denotes the mass, g is the gravitational acceleration


and R is the radius of turn of the vehicle. From (7) above, we
get
tan f cos (cG  c) =

Vg2
Rg

Vg c_ G
g

For level flight (g = 0), the northeast position of the


UAV is


P_ n
P_ e


= Vg

cos cG
sin cG


16

To compute the cross-track deviation (y) at any point (see


Figures 2 and 3) we can use the relation
17

y_ = Vg sin cE

Equations (17) and (13) represent the overall dynamics for the
outer loop guidance design problem. For simplicity, we may
define u = tan fr as the control input. In state-space form we
can then write

During a steady turn Vg = Rc_ G , so (8) becomes


tan f cos (cG  c) =

15

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Transactions of the Institute of Measurement and Control

y_ = Vg sin cE



_cE = gu cos sin1 1 Vwn sin cG + Vwe cos cG
 c_ R
Vg
Va
18
Here y and cE are the state variables, and u = tan fr is the
control signal that the guidance loop has to generate for
cross-track control.

The guidance problem


The objective of the guidance logic is to provide smooth and
stable control of the lateral deviation of the vehicle and to
keep it as small as possible in the presence of disturbances. If
a track error develops (for example in the initial phase after
take-off when the guidance loop is first activated, or when the
navigation information gets corrected after a long GPS outage), it should be brought to zero in a graceful manner without saturating the control input to the vehicle. Also, the error
should be minimized quickly with little or no overshoot for
both small and large track errors.

Sliding surface for cross-track control


In our previous work (Shah et al., 2011a,b, 2014), sliding surface design for the guidance loop was discussed in detail.
Limitations of linear surfaces were discussed, and thereafter a
piecewise linear sliding surface was proposed in Shah et al.
(2011b) and a nonlinear sliding surface was proposed in Shah
et al. (2011a, 2014). It was shown that linear surfaces do not
perform well for both small and large track errors. If a linear
surface is designed to give good performance for small errors,
it causes control saturation for large errors, and if on the

other hand the surface is designed to avoid saturation, poor


performance results for small errors. A novel nonlinear sliding surface (Figure 7) was proposed in Shah et al. (2011a,
2014) that overcomes these limitations and performs well for
both small and large error scenarios. The following equation
represents the nonlinear sliding surface:
s = cE + a arctan by = 0

Here the two adjustable parameters a, b 2 < (the set of real


numbers), and for stability of the sliding surface we require
ab . 0 (Shah et al., 2014). It is clear from (19) that we need
jaj  1 in order to ensure jcEj  p/2, which is required for
correct direction of approach while guiding the vehicle
towards the desired path. We now derive a guidance law
based on the sliding surface (19) that works for both straight
and circular paths in the presence of winds.

Guidance law design


Sliding-mode-control design is essentially a two-step process:
sliding-surface design, and derivation of a control law to
ensure that the phase trajectory is attracted towards the sliding surface. A suitable, stable sliding surface is presented in
Sliding surface for cross-track control above. We discuss a
sliding-mode-based guidance law in this section; conditions to
ensure sliding and control boundedness are also derived.

Equivalent lateral control


Equivalent control is the control input which, when applied
to the system, enables the system to continue sliding once it is
on the sliding surface (Bandyopadhyay and Janardhanan,

0.5
0.4
0.3
0.2

E [rad]

0.1
0
0.1
0.2
0.3
0.4
0.5
1000

500

19

0
Cross track error (y) [m]

500

Figure 7. The nonlinear sliding surface (19) for a particular a and b.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

1000

Shah et al.

2006). The equivalent control that maintains the sliding mode


is therefore the input ueq satisfying s_ = 0. In our case
s_ = c_ E +

ab
y_
1 + b2 y2

20

is for compensation of wind; in the absence of wind it is equal


to 1. Generally wind measurement is not available on board
UAVs, and hence this term has to be estimated. The above
guidance law (23) can be written in the form
~2
~g
V
ab V
g
u=G
c_ R 
sin cE
2
2
g~
1 + b y g~

Using state equations (18), s_ = 0 can be written as






g~ueq
1  ~
~we cos cG
0=
cos sin1
Vwn sin cG + V
~a
~g
V
V
ab
~g sin cE
21
V
c_ R +
1 + b2 y2
~ are the measured values of gravity and velowhere g~ and V
city, respectively. After simplification, the expression for
equivalent control ueq becomes
ueq =

1

 

1 1
~wn sin cG + V
~we cos cG
V
cos sin
V~a
!
~
~g2
ab V
Vg _
cR 
sin
c
E
g~
1 + b2 y2 g~

22

This control has two parts: a feedback part and a feedforward


~g =~
part ((V
g)c_ R ) that is directly related to the rate of change of
the reference course angle. It is seen that the lateral acceleration generated by the feedforward component is equal to the
centripetal acceleration during a turn, provided the velocity
Vg is known exactly. Equivalent control can maintain sliding
motion only if the state trajectory is on the sliding manifold,
and the dynamics are perfectly known (no uncertainty). A formal control law, possibly variable structure, has to be formulated to take care of the uncertainty, and to bring the system
states onto the sliding manifold in the first place.

1

fr = tan

1

cos sin

V~g2
g~
1

 1 +abb2 y2
fr

= tan

1
~a
V

24

(u)

G=

1

 

1 1
~
~
cos sin
~ Vwn sin cG + Vwe cos cG
V
a

Here we assume that both state variables y and cE are accurately measured and available for feedback control.
Conditions on k are derived in the following sections to cater
for uncertainties in different parameters, including G.
From an applications perspective, two approaches are possible to deal with the wind disturbance. In the first approach
G is taken as 1, but its maximum variation due to extreme
winds is computed beforehand during the design phase; this
variation is treated as an uncertainty and is considered in
selection of the gain k. In the second approach, an estimator
is used to estimate the wind speed and direction, and hence G
is estimated in real time and used in the guidance law (25);
any estimation error is considered while selecting k. Here in
this paper we adopt the first approach and take G equal to 1;
the wind disturbance is considered in design of the gain k.
The simplified guidance law therefore becomes
~2
~g
V
ab V
g
c_ R 
sin cE  k sgn(s)
2
2
g~
1 + b y g~

25

fr = tan1 (u)

The guidance algorithm must ensure sliding in the presence of


parametric uncertainties and input disturbances and from an
arbitrary initial condition, in finite time. For this a discontinuous control term of the form 2 k sgn (s) is added to the
equivalent control term to ensure sliding in the presence of
uncertainties. The total guidance law is then the sum of the
equivalent and discontinuous control terms. The gain k is
selected based on Lyapunov theory to ensure stable sliding
motion. The lateral guidance law therefore becomes
u

 k sgn(s)

where

u=

The complete guidance law

V~wn sin cG + V~we


cos cG

~

Vg
g~

c_ R


sin cE  k sgn(s)

(u)

where ab . 0 and jaj  1. The term


1

 

~wn sin cG + V
~we cos cG
cos sin1 V~1 V

23

Reachability condition
Reachability implies that the state trajectories are attracted
towards the sliding surface, and, once achieved, it maintains
that sliding motion for subsequent time periods
(Bandyopadhyay et al., 2009; Perruquetti and Barbot, 2002).
To check for reachability, let us take V = s2/2 as the
Lyapunov candidate function. The derivative of V is
V_ = s_s




gu
1 1
=s
cos sin
Vwn sin cG + Vwe cos cG
Vg
Va

ab
26
V
sin
c
c_ R +
g
E
1 + b 2 y2
Substituting for the control input u from (25), we get

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Transactions of the Institute of Measurement and Control




g
1
Vwn sin cG + Vwe cos cG
V_ = s cos sin1
Vg
Va
#
~
~2
Vg _
ab V
g
c 
sin cE  k sgn(s)
g~ R 1 + b2 y2 g~


ab
27
V
sin
c
+ s c_ R +
g
E
1 + b2 y2
or
"
#
~g2
~
V
_V = s g t Vg c_ R  ab
sin cE  k sgn(s)
Vg
g~
1 + b2 y2 g~


ab
+ s c_ R +
V
sin
c
g
E
1 + b 2 y2

With k as above, the phase trajectory will be attracted


towards the sliding surface across the entire flight envelope of
the vehicle. The parameters affecting the value of k are the
sliding surface variables, velocity and the radius of turn.
Keeping the product ab large will give better performance in
terms of cross-track error regulation, but will result in a bigger value of k, and hence a larger control input u. Similarly,
keeping the radius of turn R small would demand a larger u.
In the above discussion, a maximum error of 10% is assumed
in the measurement of velocity; if a different error magnitude
is expected for a particular application, expression (33) may
be modified accordingly.

28

Control boundedness

where



1
t = cos sin1
Vwn sin cG + Vwe cos cG
Va
~g and g~ (with some measurement
The uncertain variables V
error) appear from substitution of the control input u above.
Neglecting the error in the gravity term (g g~), we have



s
ab
~g2  Vg2  c_ R (tV~g  Vg ):
V_ = 
sin
c
t
V
E
V g 1 + b 2 y2

+ gkt sgn(s)
29

As discussed earlier, the aerodynamic lift required to balance


the weight of the vehicle decreases by a factor of cos f as the
UAV banks to follow the roll command generated by the guidance law. The output of the guidance block fr = arctan u
must therefore be bounded in magnitude by an upper bound,
say fmax, during both the reaching and sliding phases. This
will ensure the availability of the aerodynamic lift required to
sustain flight. Here we will derive conditions so that jfrj 
fmax throughout the reaching and sliding phases. In order to
avoid control saturation, we must have


2

V
~g
V
ab

~g
_
sin
c
+

k
sgn(s)
c
 tan fmax

E

g~ 1 + b2 y2
g~ R
or in the worst case

V_ will be negative definite if





ab

~ 2  V 2 + c_ R (t V
~g  Vg ) 30
sin
c
t
V
jgkt j .
E
g
g

1 + b2 y2
or




2
V
V
ab
~g _
~g
+ jk sgn(s)j  tan fmax 34
+
sin
c
c


E
g~ R
g~ 1 + b2 y2
After simplification, we have

k.




1
ab
1 _ ~
~2
2
cR t Vg  Vg
jsin cE j t V
g  Vg +
2
2
gt 1 + b y
gt

31

To find the value of k which satisfies the above for the entire
flight envelope, that is, to ensure reachability from any initial
condition, we find the maximum of the right-hand side in the
above inequality. The maximum value of ab/(1 + b2y2) is
ab at y = 0, and the maximum value of jsin cEj is 1 at
cE = 6p/2. Hence, the maximum value in the entire phase
portrait occurs at the point: (cE, y) = (6p/2, 0). Let us also
assume a worst-case error of 10% in the measurement of
~g 1:1 Vg ); now V_ will be negative defiground velocity (i.e. V
nite, if




Vg2 ab
Vg _
1
1
k.
+
1:21 
cR 1:1 
t
t
g
g

32

Using the minimum value of 1/t = 1 and the relation


Vg = c_ R R, we have
Vg2 ab
0:1 Vg2
+
k . 0:21
g R
g

33




V~ 2 ab
~g
V


g
sin cE
k  tan fmax  c_ R 
2
2

g~ 1 + b y
g~

35

The maximum value of the term




2

V
ab

~g
sin
c


E
2

g~ 1 + b y2
~g2 =~
cannot exceed (V
g )ab in the entire phase plane (for all values of y and cE). So in order to bound the control input for
all conditions, we must have
k  tan fmax 

~2
~g
V
V
g
c_ R 
ab
g~
g~

36

Constraints on control gain k


For both reachability and saturation avoidance, we combine
conditions (33) and (36) to get

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Shah et al.

s
|s | +

0.5

= 0.7
= 0.3
= 0.1

0.5

1
10

10

Figure 8. Approximation of sgn (s) function.

0:21

~g
Vg2 ab
V~g2
V
0:1 Vg2
+
\k  tan fmax  c_ R 
ab
g R
g
g~
g~

37

or
0:21

Vg2 ab
g

1:21Vg2
1:1Vg _
0:1 Vg2
\k  tan fmax 
cR 
ab
g R
g
g
38

The gain k and the sliding surface parameters a and b should


be selected considering the UAV flight conditions (Vg, g), the
maximum allowable roll angle (fmax) and the mission requirement (R). First the sliding parameters a and b may be chosen,
followed by the selection of the gain k that satisfies (38). If no
feasible value of k is found, the above procedure may be reiterated with comparatively smaller values of the sliding parameters a and b.

Implementation issues

Figure 9. Gain k required for reachability over the phase plane.

From an implementation point of view, several solutions have


been developed to avoid chattering in the control signal due
to the signum function. One method is to replace the discontinuous switching action sgn (s) with a continuous sigmoid
approximation s/(jsj + e) ; this is known as the boundary
layer approach (Boiko, 2013; Burton and Zinober, 1986). As
a result of this approximation, system trajectories are confined to a small vicinity of the sliding surface (the boundary
layer) and not exactly to s(t) = 0, which is the ideal sliding
mode.
Figure 8 shows the sigmoid approximation of sgn (s) for
different values of e. It is seen that the effective gain becomes
smaller as s approaches zero, and hence chattering is reduced
in the control signal. Since the gain is reduced, control
boundedness (condition (36)) is not affected in this case. The
reachability condition also needs to be analysed for this
approximation. The reachability condition for the worst-case
scenario (33) gives the maximum gain required so that the
states are attracted towards the sliding surface from any point
in the phase plane. The maximum gain corresponds to the

point (cE, y) = (6p/2, 0). In the generalized reachability


condition (31), the dominant term is
1 ab
j sin cE j
g 1 + b 2 y2
Figure 9 shows a coloured mesh plot of this term for the
entire phase plane. Thus we can easily see the gain required
to ensure reachability from any point of the phase portrait.
The sliding surface is also plotted in the figure in black line
for a = 0.7 and b = 0.002. It is evident that the maximum
gain required corresponds to (6p/2, 0). It is also seen that
the gain requirement in the close vicinity of the sliding surface
is not too high: in fact, close to the origin the required gain
reduces to one-fifteenth of its maximum value. The effect of
gain reduction (due to the approximation sgn (s) s/(jsj + e))
on the reachability condition needs to be analysed for each
specific application; it may be acceptable to bound the state
trajectory within a defined vicinity of the sliding surface. For

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

10

Transactions of the Institute of Measurement and Control

Figure 10. A photograph of the test vehicle during take-off.

Figure 11. Main interfaces of the flight control computer.


AHRS: Attitude and heading reference system.

example let us assume that we want to slide with an accuracy


of 0.1, that is, we want to keep jsj \ 0.1 (the boundary layer
width will be 0.2 in this case). Choosing e = 0.3 yields s/(jsj
+ e) = 60.25 for s = 60.1, which implies that the actual
gain delivered to the system at the edges of the boundary
layer will be one-fourth of the chosen value. Figure 9 shows
the required gain to be approximately one-fourth of the maximum in the neighbourhood of the sliding surface: hence for
this case we can have jsj \ 0.1 for e = 0.3.

Flight test results


Experimental setup
The proposed guidance law is programmed in the flight control computer of a scaled YAK-54 UAV (Figure 10) to
demonstrate its effectiveness; comparisons are also made with

other guidance laws. The test vehicle is designed and produced


by EG Aircraft and is powered by a DLE-55 engine. The
structure is suitably modified to accommodate the flight computer and related sensors. The layout of different avionics
modules is shown in Figure 11. An MPC-565 microcontrollerbased generic board forms the heart of the control computer
and generates Pulse Width Modulation (PWM) commands
for the actuators; it communicates with different on-board
sensors and the ground terminal through serial ports. With
the available fuel capacity, the UAV can fly autonomously for
about one hour. Basic data of said vehicle is listed in Table 1.

Parameter selection
For implementation of the guidance law (25), the sliding surface is s =cE + a arctan by = 0, where a and b are selected

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Shah et al.

11

Table 1: Geometrical and mass properties of the scaled YAK-54 UAV.


#

Parameter

Value

1
2
3
4
5
6
7
8

Wing span
Length
Wing area
Wing root/tip airfoil
Horizontal tail root/tip airfoil
Vertical tail root/tip airfoil
Take-off mass
Moments of inertia: Ixx, Iyy, Izz

2235 mm
2080 mm
943,869 mm2
NACA 0016 / 0017
NACA 0015 / 0012
NACA 0009 / 0010
11 kg
1.36,2.848,4.07 kg-m2

as 0.7 and 0.002 respectively. The minimum turn radius is


taken as 400 m, the speed during flight is approximately
35 m/s and the cruise altitude is 800 m above sea level. With
these parameter values, conditions for reachability and control boundedness (38) become 0.06 \ k \ 0.212. A boundary layer approximation sgn (s) s/(jsj + e) is used (see
Implementation issues) to avoid chattering in the computation of fr.
For flight we want to achieve a sliding accuracy of 0.1 (i.e.
jsj \ 0.1) and this can be achieved with e = 0.3 for k = 0.06
as discussed in Implementation issues. But here we choose a
larger gain of 0.21 (;3.5 times more than that required) for
implementation. With this larger gain we will have greater disturbance rejection outside the boundary layer, and can
achieve jsj \ 0.1 even with e = 0.7, which gives a much
smoother control signal. Hence for this application we choose
k = 0.21 and e = 0.7.

Flight results
Effectiveness of the proposed algorithm for straight path following is equivalent to that of the algorithm presented in
Shah et al. (2011a, 2014). Here we shall focus on test results
for different cases of curved path following; comparisons will
also be drawn with our previously proposed algorithm (Shah
et al., 2014), and with Parks algorithm (Park, 2004; Park et
al., 2004). Parks algorithm generally performs well for
straight flight and mild turns. However, as discussed in
Introduction, if track errors become large, the control input
(roll angle) can saturate. The lateral acceleration command
generated by Parks algorithm for a circular loiter requires
perfect knowledge of the vehicle velocity for path following
with zero steady-state error. In the case of measurement
errors in velocity, a steady-state error will develop. Also, the
performance of this algorithm is sensitive to uncertainty in
the input channel (the roll channel), for example, a bias or
shift in roll angle following.
Flight results for two different circular loiters are discussed
first. The take-off point is taken as the origin, the northward
distance travelled is denoted by x (or Posx), while the eastward distance is denoted by y (or Posy). The UAV has an
open loop speed control which is adjustable from the ground.
Initially a circular loiter is performed with a fixed throttle setting; ground speed Vg versus course angle cG is plotted in
Figure 12. A nearly constant airspeed is maintained with the

fixed throttle; variation in ground speed is seen due to wind.


From Figure 12 it is estimated that there was a wind of ;5 m/
s during the flight; all subsequent flight results show the performance of the guidance algorithm in the presence of wind of
approximately 5 m/s.
Flight results for a circular loiter of radius 800 m are
shown in Figures 13 to 15, with three different guidance algorithms. In Figure 13, the reference trajectory (the direction of
flight is clockwise) is shown in dashed line and the actual trajectory flown by the vehicle is shown in solid line. The initial
part of the circular loiter is executed with our previous algorithm (presented in Shah et al., 2014): the trajectory-following
is not accurate. After some time, the newly proposed algorithm (25) is activated, the point of transition is marked in
the figure, and thereafter the tracking accuracy improves substantially. Parks algorithm (Park et al., 2004) is brought
online in a later part of the flight. Figure 14 shows the crosstrack error y and the heading error angle cE versus time for
the same part of the flight. Initially (13401412.5 s), the guidance law of Shah et al. (2014) is online and yields a crosstrack error of ~ 55 m. This improves to around 2 m when the
proposed law (25) is activated (from 1412.5 s to 1494 s). In
the last section (14941527 s), Parks algorithm is activated; it
keeps the error within a 3 m band. Figure 15 shows the reference roll angle generated by the three guidance algorithms,
and also the actual roll angle of the vehicle along with aileron
deflections. There is no control saturation, the aileron deflections being nominal.
Flight results of the three guidance algorithms for a circular
loiter of radius 400 m (close to the maximum capability of the
UAV) are shown in Figures 16 to 18. For the first 337 s to
413 s of flight our previous algorithm (Shah et al., 2014) is activated, followed by Parks algorithm from 413 s to 461.5 s, and
then the algorithm proposed here is brought online in the last
part of the flight (461.5495 s). It is seen from Figure 17 that
trajectory-following with our previous algorithm is not good,
and a cross-track error of ;80 m develops in the initial part of
the flight. This reduces to ;20 m with Parks algorithm. With
the newly proposed algorithm (25), trajectory-following
improves and the error reduces to around 5 m. The commanded and actual roll angles along with aileron deflections
are shown in Figure 18.
Flight results of the proposed algorithm for a straight path
followed by a sharp heading change of approximately 135+
are shown in Figures 19 to 21. Figure 19 shows the desired
mission plan along with the actual trajectory flown by the
UAV. The mission consists of three parts: a straight path
WP1 WP2, a sharp turn (heading change of ;135), followed by another straight segment WP2 WP3. To see
robustness of the proposed algorithm, the speed of the UAV
is varied during the mission and ascend/descend commands
are also given, as shown in Figure 20. Figure 21 shows the
reference roll angle, the actual roll angle and the aileron
deflections. With the proposed guidance algorithm, the aircraft follows the entire mission successfully. In the last segment of the mission (straight path WP2 WP3), the following
remains very good despite ascend/descend manoeuvres and
variations in speed.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

12

Transactions of the Institute of Measurement and Control


40

Speed (Vg ) [m/s]

38

36

34

32

30
200

100

150

50

G []

100

50

150

200

Pos y (west) [m]

400
200
0
200
400
500

0
Pos x (north) [m]

500

Figure 12. Estimated wind on flight day.

1000
800
600
Switched to new algorithm

200
0

Pos (west) [m]

400

200
400
Switched to Parks algorithm
600
800

Start here with


previous algorithm

1000
1000 800

600

400

200
0
200
Pos (north) [m]

400

600

800

Figure 13. Trajectory-following for a circular loiter of radius 800 m.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

1000

Shah et al.

13

Cross-range (y) [m]

20
0

with previous algorithm

20

with Parks
algorithm

40
with new algorithm
60
80

1340

1360

1380

1400

1420 1440
Time [s]

1460

1480

1500

1520

1340

1360

1380

1400

1420 1440
Time [s]

1460

1480

1500

1520

15
10

E []

5
0
5
10

Figure 14. Cross-track error y and heading error cE vs time for the circular loiter of radius 800 m.

40

30

roll ref.
roll

with previous algorithm

[]

20

10
with Parks
algorithm

with new algorithm

10

20

1340

1360

1380

1400

1420

1440
Time [s]

1460

1480

1500

1520

1340

1360

1380

1400

1420

1440
Time [s]

1460

1480

1500

1520

Aileron (a) []

4
2
0
2
4
6

Figure 15. Roll angle and aileron deflection (da) vs time for the circular loiter of radius 800 m.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

14

Transactions of the Institute of Measurement and Control


600

400

Pos y (west) [m]

200
Switched to new algorithm
0
Switched to Parks algorithm
Start here
with previous algorithm

200

400

600
600

600

400

200
0
Pos (north) [m]

200

400

Figure 16. Trajectory-following for a circular loiter of radius 400 m.

Cross-range (y) [m]

20
0
with previous algorithm
20
with new
algorithm

40
with Parks
algorithm

60
80

340

360

380

400

420
Time [s]

440

460

480

500

340

360

380

400

420
Time [s]

440

460

480

500

15
10

E []

5
0
5
10
15

Figure 17. Cross-track error y and heading error cE vs time for the circular loiter of radius 400 m.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Shah et al.

15
40
35
roll ref.
roll

30
25
20

[]

15
10

with new
algorithm

with previous algorithm

with Parks
algorithm

0
5
10
15

340

360

380

400

420
Time [s]

440

460

480

500

340

360

380

400

420
Time [s]

440

460

480

500

Aileron (a) []

4
2
0
2
4
6

Figure 18. Roll angle and aileron deflection (da) vs time for the circular loiter of radius 400 m.

33.525
Start here

Latitude []

33.52

WP1

WP3

33.515

33.51

33.505

Reference path
UAV trajectory

WP2

33.5
73.162 73.164 73.166 73.168 73.17 73.172 73.174 73.176 73.178 73.18
Longitude []
Figure 19. Trajectory-following for a straight path followed by a sharp turn.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

16

Transactions of the Institute of Measurement and Control


150

[m]

100

Reference alt
Actual altitude

50

0
1180

1200

1220

1240

1260

1280

1300

1200

1220

1240
Time [s]

1260

1280

1300

Speed [m/s]

50
40
30
20
10
1180

Figure 20. Altitude and speed for straight path and sharp turn.

40
roll ref.
roll

30

[]

20
10
0
10
20
1180

1200

1220

1240

1260

1280

1300

[ ]

2
0
2
4
1180

1200

1220

1240
Time [s]

1260

1280

Figure 21. Roll angle and aileron deflection for straight path and sharp turn.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

1300

Shah et al.

17

Conclusion
In this paper a new lateral guidance law for cross-track control of UAVs is developed. The guidance law is based on the
sliding mode technique. For good performance and accurate
tracking while following curved paths, a feedforward term
which depends on the rate of change of the reference path
heading is included in the guidance law. The algorithm uses
the current mission information (reference path) and sensor
measurements to generate roll commands. Conditions for stability and control boundedness are derived in terms of upper
and lower bounds on the control gain k. The guidance algorithm is implemented in the flight computer of a scaled YAK54 research aircraft, and various flight tests conducted.
Comparison with other algorithms is also made. It is seen
that the performance of the proposed algorithm is better than
the previously suggested algorithm (Shah et al., 2014), especially during circular loiters. The sliding-mode-based guidance law performs well in the presence of wind, parametric
and input channel uncertainties; it does not saturate the control input, and meets the criterion of a good helmsman.
Generalization of this 2D lateral guidance problem to the full
3D trajectory tracking problem is the subject of ongoing
work.

Funding
This research received no specific grant from any funding
agency in the public, commercial, or not-for-profit sectors.

References
Ahmed M and Subbarao K (2010) Nonlinear 3-D trajectory guidance
for unmanned aerial vehicles. In: Proceedings of the 11th international conference on control, automation, robotics and vision.
Bandyopadhyay B and Janardhanan S (2006) Discrete-Time Sliding
Mode Control: A Multirate Output Feedback Approach. Berlin/
Heidelberg: Springer-Verlag.
Bandyopadhyay B, Deepak F and Kim KS (2009) Sliding Mode Control Using Novel Sliding Surfaces. Berlin/Heidelberg: SpringerVerlag.
Beard RW and McLain TW (2012) Small Unmanned Aircraft: Theory
and Practice. Princeton, NJ: Princeton University Press.
Boiko I (2013) Chattering in sliding mode control systems with
boundary layer approximation of discontinuous control. International Journal of Systems Science 44(6): 11261133.
Brezoescu A, Espinoza T, Castillo P, et al. (2013) Adaptive trajectory
following for a fixed-wing UAV in presence of crosswind. Journal
of Intelligent and Robotic Systems 69(14): 257271.
Burton JA and Zinober ASI (1986) Continuous approximation of
variable structure control. International Journal of Systems Science 17(6): 875885.
Ceccarelli N, Enright J, Frazzoli E, et al. (2007) Micro UAV path
planning for reconnaissance in wind. In: American control conference, pp. 53105315.
Cesetti A, Frontoni E, Mancini A, et al. (2009) A vision-based guidance system for UAV navigation and safe landing using natural
landmarks. Journal of Intelligent and Robotic Systems 57(14):
233257.
Dadkhah N and Mettler B (2012) Survey of motion planning literature in the presence of uncertainty: Considerations for UAV guidance. Journal of Intelligent and Robotic Systems 65(1): 233246.

Deyst J, How J and Park S (2005) Lyapunov stability of a nonlinear


guidance law for UAVs. In: AIAA atmospheric flight mechanics
conference and exhibit, San Francisco, CA.
Goerzen C, Kong Z and Mettler B (2009) A survey of motion planning algorithms from the perspective of autonomous UAV guidance. Journal of Intelligent and Robotic Systems 57(14): 65100.
Griffiths S (2006) Vector-field approach for curved path following for
miniature aerial vehicles. In: AIAA guidance, navigation, and control conference and exhibit, Keystone, CO.
Kaminer I, Pascoal A, Hallberg E, et al. (1998) Trajectory tracking
for autonomous vehicles: An integrated approach to guidance and
control. AIAA Journal of Guidance, Control, and Dynamics 21(1):
2938.
McGee T and Hedrick J (2006) Path planning and control for multiple point surveillance by an unmanned aircraft in wind. In: American control conference, p. 6.
Niculescu M (2001) Lateral track control law for Aerosonde UAV.
In: 39th AIAA aerospace sciences meeting and exhibit, Reno, NV.
Osborne J and Rysdyk R (2005) Waypoint guidance for small UAVs
in wind. AIAA Infotech@ Aerospace 193: 112.
Pappoullis FA (1994) Cross track errors and proportional tuning
rate guidance of marine vehicle. Journal of Ship Research 38:
123132.
Park S (2004) Avionics and control system development for mid-air rendezvous of two unmanned aerial vehicles. PhD Thesis, Department
of Aeronautics and Astronautics, Massachusetts Institute of Technology, MA.
Park S, Deyst J and How J (2004) A new nonlinear guidance logic for
trajectory tracking. In: AIAA guidance, navigation, and control
conference and exhibit, Providence, RI.
Perruquetti W and Barbot JP (2002) Sliding Mode Control in Engineering. New York, NY: Marcel Dekker Inc.
Pettersen K and Lefeber E (2001) Way-point tracking control of
ships. In: Proceedings of the 40th IEEE conference on decision and
control, Orlando, FL.
Pisano W, Lawrence D and Gray P (2007) Autonomous UAV control
using a 3-sensor autopilot. In: AIAA infotech aerospace 2007 conference and exhibit, Rohnert Park, CA.
Regina N and Zanzi M (2011) UAV guidance law for ground-based
target trajectory tracking and loitering. In: Proceedings of the
IEEE aerospace conference.
Ren W and Beard RW (2004) Trajectory tracking for unmanned air
vehicles with velocity and heading rate constraints. IEEE Transactions on Control Systems Technology 12(5): 706716.
Rysdyk R (2006) Unmanned aerial vehicle path following for target
observation in wind. Journal of Guidance, Navigation, and Control
29(5): 10921100.
Samar R, Ahmed S and Aftab MF (2007) Lateral control with
improved performance for UAVs. In: The 17th IFAC symposium
on automatic control in aerospace, Toulouse, France.
Samar R, Ahmed S and Nzar M (2008) Lateral guidance & control
design for an unmanned aerial vehicle. In: 17th IFAC world congress, Seoul, South Korea.
Sasongko R, Sembiring J, Muhammad H, et al. (2011) Path following
system of small unmanned autonomous vehicle for surveillance
application. In: Proceedings of the 8th Asian control conference.
Shah MZ, Samar R and Bhatti AI (2011a) Lateral control for UAVs
using sliding mode technique. In: 18th IFAC world congress,Milano, Italy.
Shah MZ, Samar R and Bhatti AI (2011b) Sliding mode based lateral
control for UAVs using piecewise linear sliding surface. In: International conference on communications, computing and control
applications (CCCA11), Hammamet, Tunisia.
Shah MZ, Samar R and Bhatti AI (2012) Cross-track control of
UAVs during circular and straight path following using sliding

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

18

Transactions of the Institute of Measurement and Control

mode approach. In: 12th international conference on control, automation and systems (ICCAS), pp. 185190.
Shah MZ, Samar R and Bhatti AI (2014) Guidance of air vehicles: A
sliding mode approach. IEEE Transactions on Control Systems
Technology. Epub ahead of print 3 June 2014. DOI: 10.1109/
TCST.2014.2322773.
Shtessel Y, Shkolnikov I and Brown M (2003) An asymptotic secondorder smooth sliding mode control. Asian Journal of Control 4(5):
959967.
Shtessel YB, Shkolnikov IA and Levant A (2007) Smooth secondorder sliding modes: Missile guidance application. Automatica
43(8): 14701476.
Siouris GM (2004) Missile Guidance and Control Systems. Berlin:
Springer-Verlag.
Skogestad S and Postlethwaite I (2005) Multivariable Feedback Control Analysis and Design. 2nd edn. New York, NY: John Wiley &
Sons.

Yamasaki T and Balakrishnan SN (2010) Preliminary assessment of


flying and handling qualities of mini-UAVs. Journal of Aerospace
Engineering 224(10): 10571067.
Yamasaki T, Balakrishnan S and Takano H (2012a) Integrated guidance and autopilot for a path-following UAV via high-order sliding modes. In: American control conference (ACC), pp. 143148.
Yamasaki T, Balakrishnanb S, Hallberg E, et al. (2012b) Integrated
guidance and autopilot design for a chasing UAV via high-order
sliding modes. Journal of the Franklin Institute 349(2): 531558.
Yu J, Xu Q and Zhi Y (2011) A TSM control scheme of integrated
guidance/autopilot design for UAV. In: 3rd international conference on computer research and development (ICCRD), Shanghai,
China.
Zhang Z, Li S and Luo S (2013) Composite guidance laws based on
sliding mode control with impact angle constraint and autopilot
lag. Transactions of the Institute of Measurement and Control 35:
764776.

Downloaded from tim.sagepub.com at NATL AEROSPACE LABORATORIES on November 13, 2014

Vous aimerez peut-être aussi