scispace - formally typeset
Search or ask a question
Proceedings ArticleDOI

Geometric tracking control of a quadrotor UAV on SE(3)

TL;DR: New results for the tracking control of a quadrotor unmanned aerial vehicle (UAV) are provided and a nonlinear tracking controller is developed on the special Euclidean group SE(3), shown to have desirable closed loop properties that are almost global.
Abstract: This paper provides new results for the tracking control of a quadrotor unmanned aerial vehicle (UAV). The UAV has four input degrees of freedom, namely the magnitudes of the four rotor thrusts, that are used to control the six translational and rotational degrees of freedom, and to achieve asymptotic tracking of four outputs, namely, three position variables for the vehicle center of mass and the direction of one vehicle body-fixed axis. A globally defined model of the quadrotor UAV rigid body dynamics is introduced as a basis for the analysis. A nonlinear tracking controller is developed on the special Euclidean group SE(3) and it is shown to have desirable closed loop properties that are almost global. Several numerical examples, including an example in which the quadrotor recovers from being initially upside down, illustrate the versatility of the controller.

Summary (2 min read)

Introduction

  • A quadrotor unmanned aerial vehicle (UAV) consists of two pairs of counter-rotating rotors and propellers, located at the vertices of a square frame.
  • Due to its simple mechanical structure, it has been envisaged for various applications such as surveillance or mobile sensor networks as well as for educational purposes.
  • Little attention has been paid to constructing nonlinear control systems for them, particularly to designing nonlinear tracking controllers.
  • By characterizing geometric properties of nonlinear manifolds intrinsically, geometric control techniques provide unique insights to control theory that cannot be obtained from dynamic models represented using local coordinates [15].
  • The authors construct a tracking controller to follow prescribed trajectories for the center of mass and heading direction.

II. QUADROTOR DYNAMICS MODEL

  • Consider a quadrotor vehicle model illustrated in Fig. 1.
  • The first and the second axes of the body fixed frame, ~b1,~b2, lie in the plane defined by the centers of the four rotors, as illustrated in Fig.
  • The authors develop a controller to follow a prescribed trajectory of the location of the center of mass, xd(t), and the desired direction of the first body-fixed axis, ~b1d(t).
  • As a result, the authors are able to achieve almost global exponential attractiveness to the zero equilibrium of tracking errors.

A. Tracking Errors

  • The authors define the tracking errors for x, v,R,Ω as follows.
  • From this, the attitude tracking error eR is chosen to be eR = 1 2 (RTdR−RTRd)∨. (10) The tangent vectors Ṙ ∈ TRSO(3) and Ṙd ∈ TRdSO(3) cannot be directly compared since they lie in different tangent spaces.
  • The authors choose the tracking error for the angular velocity as follows: eΩ = Ω−RTRdΩd. (11).

C. Exponential Asymptotic Stability

  • The authors first show exponential stability of the attitude dynamics in the sublevel set L2 = {Rd, R ∈ SO(3) |Ψ(R,Rd) < 2}, and based on this result, they show exponential stability of the complete dynamics in the smaller sublevel set L1 = {Rd, R ∈ SO(3) |Ψ(R,Rd) < 1} Proposition 1: (Exponential Stability of Attitude Dynamics).
  • (17), (18) represent a region of attraction for the attitude dynamics.
  • More precisely, the position tracking performance is affected by the difference between ~b3 = Re3 and ~b3d = Rde3.
  • In the proceeding stability analysis, it turns out that for the stability of the complete translational and rotational dynamics, the attitude error function Ψ should be less than 1, which states that the initial attitude error should be less than 90◦.
  • Proposition 2: (Exponential Stability of the Complete Dynamics).

D. Almost Global Exponential Attractiveness

  • Proposition 2 requires that the initial attitude error is less than 90◦ to achieve exponential stability of the complete dynamics.
  • Suppose that this is not satisfied, i.e. 1 ≤ Ψ(R(0), Rd(0)) < 2. From Proposition 1, the authors are guaranteed that the attitude error function Ψ exponentially decreases, and therefore, it enters the region of attraction of Proposition 2 in a finite time.
  • Therefore, by combining the results of Proposition 1 and 2, the authors can show almost global exponential attractiveness.
  • This should be distinguished from the stronger notion of exponential stability, in which the constant α(δ) in the above bound is replaced by α(δ) ‖z(0)‖.
  • (Almost Global Exponential Attractiveness of the Complete Dynamics) Consider a control system designed according to Proposition 2, also known as Proposition 3.

E. Properties and Extensions

  • One of the unique properties of the presented controller is that it is directly developed on SE(3) using rotation matrices.
  • It also avoids the ambiguities that arise when using quaternions to represent the attitude dynamics.
  • Without these considerations, a quaternion-based controller can exhibit an unwinding phenomenon, where the controller unnecessarily rotates the attitude through large angles [15].
  • The attitude error function defined in (8) has the following critical points: the identity matrix, and rotation matrices that can be written as exp(πv̂) for any v ∈ S2.
  • The presented controller can be modified accordingly.

IV. NUMERICAL EXAMPLE

  • The parameters of the quadrotor UAV are chosen according to a quadrotor UAV developed in [2].
  • Simulation results are presented in Figures 4 and 5.
  • This corresponds to Proposition 3, which implies almost global exponential attractiveness.
  • After that instant, the position tracking error and the angular velocity error converge to zero as shown in Figures 5(c) and 5(d).
  • The region of attraction of the proposed control system almost covers SO(3), so that the corresponding controlled quadrotor UAV can recover from being initially upside down.

V. CONCLUSION

  • The authors presented a global dynamic model for a quadrotor UAV, and they developed a geometric tracking controller directly on the special Euclidean group that is intrinsic and coordinate-free, thereby avoiding the singularities of Euler angles and the ambiguities of quaternions in representing attitude.
  • This controller can be extended as follows.
  • But, without changing the controller structure, they can be used to follow arbitrary three-dimensional attitude commands.
  • The remaining one input degree of freedom can be used to maintain the altitude as much as possible.

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

Geometric Tracking Control of a Quadrotor UAV on SE(3)
Taeyoung Lee
, Melvin Leok
, and N. Harris McClamroch
Abstract This paper provides new results for the tracking
control of a quadrotor unmanned aerial vehicle (UAV). The
UAV has four input degrees of freedom, namely the magnitudes
of the four rotor thrusts, that are used to control the six
translational and rotational degrees of freedom, and to achieve
asymptotic tracking of four outputs, namely, three position
variables for the vehicle center of mass and the direction of
one vehicle body-fixed axis. A globally defined model of the
quadrotor UAV rigid body dynamics is introduced as a basis
for the analysis. A nonlinear tracking controller is developed
on the special Euclidean group SE(3) and it is shown to
have desirable closed loop properties that are almost global.
Several numerical examples, including an example in which the
quadrotor recovers from being initially upside down, illustrate
the versatility of the controller.
I. INTRODUCTION
A quadrotor unmanned aerial vehicle (UAV) consists of
two pairs of counter-rotating rotors and propellers, located
at the vertices of a square frame. It is capable of vertical
take-off and landing (VTOL), but it does not require com-
plex mechanical linkages, such as swash plates or teeter
hinges, that commonly appear in typical helicopters. Due
to its simple mechanical structure, it has been envisaged for
various applications such as surveillance or mobile sensor
networks as well as for educational purposes. There are
several university-level projects [1], [2], [3], [4], and com-
mercial products [5], [6], [7] related to the development and
application of quadrotor UAVs.
Despite the substantial interest in quadrotor UAVs, little
attention has been paid to constructing nonlinear control
systems for them, particularly to designing nonlinear tracking
controllers. Linear control systems such as proportional-
derivative controllers or linear quadratic regulators are widely
used to enhance the stability properties of an equilibrium [1],
[3], [4], [8], [9]. A nonlinear controller is developed for
the linearized dynamics of a quadrotor UAV with saturated
positions in [10]. Backstepping and sliding mode techniques
are applied in [11]. Since these controllers are based on
Euler angles, they exhibit singularities when representing
complex rotational maneuvers of a quadrotor UAV, thereby
fundamentally restricting their ability to track nontrivial
trajectories.
Taeyoung Lee, Mechanical and Aerospace Engineering, Florida Institute
of Technology, Melbourne, FL 39201 taeyoung@fit.edu
Melvin Leok, Mathematics, University of California at San Diego, La
Jolla, CA 92093 mleok@math.ucsd.edu
N. Harris McClamroch, Aerospace Engineering, University of Michigan,
Ann Arbor, MI 48109 nhm@umich.edu
This research has been supported in part by NSF under grants CMMI-
1029551.
This research has been supported in part by NSF under grants DMS-
0726263, DMS-1001521, DMS-1010687, and CMMI-1029445.
Geometric control is concerned with the development of
control systems for dynamic systems evolving on nonlinear
manifolds that cannot be globally identified with Euclidean
spaces [12], [13], [14]. By characterizing geometric proper-
ties of nonlinear manifolds intrinsically, geometric control
techniques provide unique insights to control theory that
cannot be obtained from dynamic models represented using
local coordinates [15]. This approach has been applied to
fully actuated rigid body dynamics on Lie groups to achieve
almost global asymptotic stability [14], [16], [17], [18].
In this paper, we develop a geometric controller for
a quadrotor UAV. The dynamics of a quadrotor UAV is
expressed globally on the configuration manifold of the
special Euclidean group SE(3). We construct a tracking
controller to follow prescribed trajectories for the center of
mass and heading direction. It is shown that this controller
exhibits almost global exponential attractiveness to the zero
equilibrium of tracking errors. Since this is a coordinate-
free control approach, it completely avoids singularities and
complexities that arise when using local coordinates.
Compared to other geometric control approaches for rigid
body dynamics, this is distinct in the sense that it controls
an underactuated quadrotor UAV to stabilize six translational
and rotational degrees of freedom using four thrust inputs,
while asymptotically tracking four outputs consisting of its
position and heading direction. We demonstrate that this
controller is particularly useful for complex, acrobatic ma-
neuvers of a quadrotor UAV, such as recovering from being
initially upside down.
II. QUADROTOR DYNAMICS MODEL
Consider a quadrotor vehicle model illustrated in Fig.
1. This is a system of four identical rotors and propellers
located at the vertices of a square, which generate a thrust
and torque normal to the plane of this square.
We choose an inertial reference frame {
~
i
1
,
~
i
2
,
~
i
3
} and a
body-fixed frame {
~
b
1
,
~
b
2
,
~
b
3
}. The origin of the body-fixed
frame is located at the center of mass of this vehicle. The
first and the second axes of the body fixed frame,
~
b
1
,
~
b
2
,
lie in the plane defined by the centers of the four rotors, as
illustrated in Fig. 1. The third body fixed axis
~
b
3
is normal
to this plane, and points downward, opposite to the direction
of the total thrust. Define
m R the total mass
J R
3×3
the inertia matrix with respect to the body-
fixed frame
R SO(3) the rotation matrix from the body-fixed frame
to the inertial frame
R
3
the angular velocity in the body-fixed frame
49th IEEE Conference on Decision and Control
December 15-17, 2010
Hilton Atlanta Hotel, Atlanta, GA, USA
978-1-4244-7744-9/10/$26.00 ©2010 IEEE 5420

~
i
1
~
i
2
~
i
3
~
b
1
~
b
2
~
b
3
f
1
f
2
f
3
f
4
x
R
Fig. 1. Quadrotor model
x R
3
the location of the center of mass in the inertial
frame
v R
3
the velocity of the center of mass in the inertial
frame
d R the distance from the center of mass to the
center of each rotor in the
~
b
1
,
~
b
2
plane
f
i
R the thrust generated by the i-th propeller along
the
~
b
3
axis
τ
i
R the torque generated by the i-th propeller about
the
~
b
3
axis
f R the total thrust, i.e., f =
P
4
i=1
f
i
M R
3
the total moment in the body-fixed frame
The configuration of this quadrotor UAV is defined by the
location of the center of mass and the attitude with respect to
the inertial frame. Therefore, the configuration manifold is
the special Euclidean group SE(3), which is the semidirect
product of R
3
and the special orthogonal group SO(3) =
{R R
3×3
| R
T
R = I, det R = 1}.
We assume that the thrust of each propeller is directly
controlled, i.e., we do not consider the dynamics of rotors
and propellers, and the direction of the thrust is normal to
the quadrotor plane. The total thrust is f =
P
4
i=1
f
i
, which
acts along the direction of
~
b
3
. According to the definition
of the rotation matrix R SO(3), the direction of the i-th
body fixed axis
~
b
i
is given by Re
i
in the inertial frame, where
e
1
= [1; 0; 0], e
2
= [0; 1; 0], e
3
= [0; 0; 1] R
3
. Therefore,
the total trust is given by f Re
3
R
3
in the inertial frame.
We also assume that the torque generated by each propeller
is directly proportional to its thrust. Since it is assumed that
the first and the third propellers rotate clockwise, and the
second and the fourth propellers rotate counterclockwise,
when they are generating a positive thrust f
i
along
~
b
3
, the
torque generated by the i-th propeller can be written as τ
i
=
(1)
i
c
τ f
f
i
for a fixed constant c
τ f
. All of these assumptions
are common [19], [4], and the presented control system can
readily be extended to include linear rotor dynamics studied
in [11]. Under these assumptions, the total thrust f and the
total moment M can be written as
f
M
1
M
2
M
3
=
1 1 1 1
0 d 0 d
d 0 d 0
c
τ f
c
τ f
c
τ f
c
τ f
f
1
f
2
f
3
f
4
. (1)
The determinant of the above 4 × 4 matrix is 8c
τ f
d
2
, so it
is invertible when d 6= 0 and c
τ f
6= 0. Therefore, for a given
f, M, the thrust of each rotor f
i
can be obtained from (1).
Using this equation, the total thrust f R and the moment
M R
3
are considered as control inputs in this paper.
The equations of motion of this quadrotor UAV can be
written as
˙x = v, (2)
m ˙v = mge
3
fRe
3
, (3)
˙
R = R
ˆ
, (4)
J
˙
+ × J = M, (5)
where the hat map
ˆ
· : R
3
so(3) is defined by the condition
that ˆxy = x × y for all x, y R
3
.
III. GEOMETRIC TRACKING CONTROL ON SE(3)
We develop a controller to follow a prescribed trajectory
of the location of the center of mass, x
d
(t), and the desired
direction of the first body-fixed axis,
~
b
1
d
(t).
The overall controller structure is as follows. The trans-
lational dynamics of a quadrotor UAV is controlled by the
total thrust fRe
3
, where the magnitude of the total thrust
f is directly controlled, and the direction of the total thrust
Re
3
is along the third body-fixed axis
~
b
3
. For a given
translational command x
d
(t), we select the total thrust f,
and the desired direction of the third body-fixed axis
~
b
3
d
to
stabilize the translational dynamics.
Once the desired direction of the third body-fixed frame
~
b
3
d
is chosen, there is one remaining degree of freedom in
selecting the desired attitude, R
d
SO(3). This corresponds
to the heading direction of the quadrotor UAV in the plane
normal to
~
b
3
d
, and it is determined by the desired direction
of the first body-fixed axis,
~
b
1
d
(t). More explicitly, we
assume that
~
b
1
d
is not parallel to
~
b
3
d
, and we project it
onto the plane normal to
~
b
3
d
to obtain the complete desired
attitude: R
d
= [
~
b
2
d
×
~
b
3
d
,
~
b
2
d
,
~
b
3
d
] SO(3), where
~
b
2
d
=
(
~
b
3
d
×
~
b
1
d
)/k
~
b
3
d
×
~
b
1
d
k. The control moment M is selected
to follow this desired attitude.
In short, four-dimensional control inputs are designed to
follow a three-dimensional translational command, and a
one-dimensional heading direction, i.e., we guarantee that
x(t) x
d
(t) and Proj[
~
b
1
(t)] Proj[
~
b
1
d
(t)] as t ,
where Proj[·] denotes the normalized projection onto the
plane orthogonal to
~
b
3
d
. The overall controller structure and
the definition of the heading direction are illustrated in Fig.
2 and 3.
Trajectory
tracking
Attitude
tracking
-
-
-
-
-
Quadrotor
Dynamics
-
f
M
~
b
3
d
x
d
~
b
1
d
x, v, R,
6
-q
q
Controller
Fig. 2. Controller structure
5421

~
b
3
d
~
b
1
d
~
b
2
d
=
~
b
3
d
×
~
b
1
d
k
~
b
3
d
×
~
b
1
d
k
Proj[
~
b
1
d
] =
~
b
2
d
×
~
b
3
d
Plane
normal to
~
b
3
d
Fig. 3. Definition of the desired heading direction
We develop this controller directly on the nonlinear con-
figuration Lie group and thereby avoid any singularities and
complexities that arise in local coordinates. As a result, we
are able to achieve almost global exponential attractiveness
to the zero equilibrium of tracking errors. Due to page limit
constraint, all of the proofs are relegated to [20].
A. Tracking Errors
We define the tracking errors for x, v, R, as follows. The
tracking errors for the position and the velocity are given by:t
e
x
= x x
d
, (6)
e
v
= v v
d
. (7)
The attitude and angular velocity tracking error should be
carefully chosen as they evolve on the tangent bundle of
the nonlinear space SO(3). The error function on SO(3) is
chosen to be
Ψ(R, R
d
) =
1
2
tr
I R
T
d
R
. (8)
This is locally positive-definite about R = R
d
within the
region where the rotation angle between R and R
d
is
less than 180
[14]. This set can be represented by the
sublevel set of Ψ where Ψ < 2, namely L
2
= {R
d
, R
SO(3) | Ψ(R, R
d
) < 2}, which almost covers SO(3). When
the variation of the rotation matrix is expressed as δR = Rˆη
for η R
3
, the derivative of the error function is given by
D
R
Ψ(R, R
d
) · Rˆη =
1
2
tr
R
T
d
Rˆη
=
1
2
(R
T
d
R R
T
R
d
)
· η, (9)
where the vee map
: so(3) R
3
is the inverse of the
hat map. We used the fact that
1
2
tr[ˆxˆy] = x
T
y for any
x, y R
3
. From this, the attitude tracking error e
R
is chosen
to be
e
R
=
1
2
(R
T
d
R R
T
R
d
)
. (10)
The tangent vectors
˙
R T
R
SO(3) and
˙
R
d
T
R
d
SO(3)
cannot be directly compared since they lie in different
tangent spaces. We transform
˙
R
d
into a vector in T
R
SO(3),
and we compare it with
˙
R as follows:
˙
R
˙
R
d
(R
T
d
R) = R
ˆ
R
d
ˆ
d
R
T
d
R = R(Ω R
T
R
d
d
)
.
We choose the tracking error for the angular velocity as
follows:
e
= R
T
R
d
d
. (11)
We can show that e
is the angular velocity of the rotation
matrix R
T
d
R, represented in the body-fixed frame, since
d
dt
(R
T
d
R) = (R
T
d
R) ˆe
.
B. Tracking Controller
For given smooth tracking commands x
d
(t),
~
b
1
d
(t), and
some positive constants k
x
, k
v
, k
R
, k
, we define
~
b
3
d
=
k
x
e
x
k
v
e
v
mge
3
+ m¨x
d
k−k
x
e
x
k
v
e
v
mge
3
+ m¨x
d
k
, (12)
where we assume that
k−k
x
e
x
k
v
e
v
mge
3
+ m¨x
d
k 6= 0. (13)
We also assume that
~
b
1
d
is not parallel to
~
b
3
d
. The desired
attitude is given by R
d
= [
~
b
2
d
×
~
b
3
d
,
~
b
2
d
,
~
b
3
d
] SO(3), where
~
b
2
d
= (
~
b
3
d
×
~
b
1
d
)/k
~
b
3
d
×
~
b
1
d
k. The desired trajectory satisfies
k mge
3
+ m¨x
d
k < B (14)
for a given positive constant B. The control inputs f, M are
chosen as follows:
f = (k
x
e
x
k
v
e
v
mge
3
+ m¨x
d
) · Re
3
, (15)
M = k
R
e
R
k
e
+ × J
J(
ˆ
R
T
R
d
d
R
T
R
d
˙
d
). (16)
The control moment M given in (16) corresponds to a
tracking controller on SO(3). For the attitude dynamics of a
rigid body described by (4), (5), this controller exponentially
stabilizes the zero equilibrium of the attitude tracking errors.
Similarly, the expression in the parentheses in (15) corre-
sponds to a tracking controller for the translational dynamics
on R
3
. The total thrust f and the desired direction
~
b
3
d
of
the third body-fixed axis are chosen such that if there is
no attitude tracking error, the thrust vector term fRe
3
in the translational dynamics of (3) becomes this tracking
controller in R
3
. Therefore, the translational tracking error
will converge to zero provided that the attitude tracking error
is identically zero.
Certainly, the attitude tracking error may not be zero at any
instant. As the attitude tracking error increases, the direction
of the control input term fRe
3
of the translational dynamics
deviates from the desired direction R
d
e
3
. This may cause
instability in the complete dynamics. In (15), we carefully
design the magnitude of the total thrust f so that it is reduced
when there is a large attitude tracking error. The expression
of f includes the dot product of the desired third body-fixed
axis
~
b
3
d
= R
d
e
3
and the current third body-fixed axis
~
b
3
=
Re
3
. Therefore, the magnitude of f is reduced when the
angle between those two axes becomes larger. These effects
are carefully analyzed in the stability proof for the complete
dynamics in [20].
In short, this control system is designed so that the
position tracking error converges to zero when there is no
attitude tracking error, and it is properly adjusted for non-
zero attitude tracking errors to achieve asymptotic stability
of the complete dynamics.
C. Exponential Asymptotic Stability
We first show exponential stability of the attitude dynamics
in the sublevel set L
2
= {R
d
, R SO(3) | Ψ(R, R
d
) <
2}, and based on this result, we show exponential stability
5422

of the complete dynamics in the smaller sublevel set L
1
=
{R
d
, R SO(3) | Ψ(R, R
d
) < 1}
Proposition 1: (Exponential Stability of Attitude Dynam-
ics) Consider the control moment M defined in (16) for any
positive constants k
R
, k
. Suppose that the initial condition
satisfies
Ψ(R(0), R
d
(0)) < 2, (17)
ke
(0)k
2
<
2
λ
max
(J)
k
R
(2 Ψ(R(0), R
d
(0))), (18)
where λ
min
(J) denotes the minimum eigenvalue of the
inertia matrix J. Then, the zero equilibrium of the attitude
tracking error e
R
, e
is exponentially stable. Furthermore,
there exist constants α
2
, β
2
> 0 such that
Ψ(R(t), R
d
(t)) min
2, α
2
e
β
2
t
. (19)
Proof: See [20].
In this proposition, (17), (18) represent a region of attrac-
tion for the attitude dynamics. This requires that the initial
attitude error should be less than 180
. Therefore, the region
of attraction for the attitude almost covers SO(3), and the
region of attraction for the angular velocity can be increased
by choosing a larger controller gain k
R
in (18).
Since the direction of the total thrust is the third body-
fixed axis, the stability of the translational dynamics depends
on the attitude tracking error. More precisely, the position
tracking performance is affected by the difference between
~
b
3
= Re
3
and
~
b
3
d
= R
d
e
3
. In the proceeding stability
analysis, it turns out that for the stability of the complete
translational and rotational dynamics, the attitude error func-
tion Ψ should be less than 1, which states that the initial
attitude error should be less than 90
. For the stability of
the complete system, we restrict the initial attitude error to
obtain the following proposition.
Proposition 2: (Exponential Stability of the Complete Dy-
namics) Consider the control force f and moment M defined
in (15), (16). Suppose that the initial condition satisfies
Ψ(R(0), R
d
(0)) ψ
1
< 1 (20)
for a fixed constant ψ
1
. Define W
1
, W
12
, W
2
R
2×2
to be
W
1
=
c
1
k
x
m
c
1
k
v
2m
(1 + α)
c
1
k
v
2m
(1 + α) k
v
(1 α) c
1
, (21)
W
12
=
k
x
e
v
max
+
c
1
m
B 0
B 0
, (22)
W
2
=
"
c
2
k
R
λ
max
(J)
c
2
k
2λ
min
(J)
c
2
k
2λ
min
(J)
k
c
2
#
, (23)
where α =
p
ψ
1
(2 ψ
1
), e
v
max
= max{ke
v
(0)k,
B
k
v
(1α)
},
c
1
, c
2
R. For any positive constants k
x
, k
v
, we choose
positive constants c
1
, c
2
, k
R
, k
such that
c
1
< min
k
v
(1 α),
4mk
x
k
v
(1 α)
k
2
v
(1 + α)
2
+ 4mk
x
,
p
k
x
m
,
(24)
c
2
< min
k
,
4k
k
R
λ
min
(J)
2
k
2
λ
max
(J) + 4k
R
λ
min
(J)
2
,
p
k
R
λ
min
(J),
r
2
2 ψ
1
k
R
λ
max
(J)
,
(25)
λ
min
(W
2
) >
4kW
12
k
2
λ
min
(W
1
)
. (26)
Then, the zero equilibrium of the tracking errors of the com-
plete system is exponentially stable. The region of attraction
is characterized by (20) and
ke
(0)k
2
<
2
λ
max
(J)
k
R
(1 Ψ(R(0), R
d
(0))). (27)
Proof: See [20].
D. Almost Global Exponential Attractiveness
Proposition 2 requires that the initial attitude error is
less than 90
to achieve exponential stability of the com-
plete dynamics. Suppose that this is not satisfied, i.e. 1
Ψ(R(0), R
d
(0)) < 2. From Proposition 1, we are guaranteed
that the attitude error function Ψ exponentially decreases, and
therefore, it enters the region of attraction of Proposition
2 in a finite time. Therefore, by combining the results of
Proposition 1 and 2, we can show almost global exponential
attractiveness.
Definition 1: (Exponential Attractiveness [21]) An equi-
librium point z = 0 of a dynamic systems is exponentially
attractive if, for some δ > 0, there exists a constant α(δ) > 0
and β > 0 such that kz(0)k < δ kz(t)k α(δ)e
βt
for
any t > 0.
This should be distinguished from the stronger notion of
exponential stability, in which the constant α(δ) in the above
bound is replaced by α(δ) kz(0)k.
Proposition 3: (Almost Global Exponential Attractiveness
of the Complete Dynamics) Consider a control system de-
signed according to Proposition 2. Suppose that the initial
condition satisfies
1 Ψ(R(0), R
d
(0)) < 2, (28)
ke
(0)k
2
<
2
λ
max
(J)
k
R
(2 Ψ(R(0), R
d
(0))). (29)
Then, the zero equilibrium of the tracking errors of the
complete dynamics is exponentially attractive.
Proof: See [20].
Since the region of attraction given by (28) for the
attitude almost covers SO(3), this is referred to as almost
global exponential attractiveness in this paper. The region
of attraction for the angular velocity can be expanded by
choosing a large gain k
R
in (29).
E. Properties and Extensions
One of the unique properties of the presented controller
is that it is directly developed on SE(3) using rotation ma-
trices. Therefore, it avoids the complexities and singularities
associated with local coordinates of SO(3), such as Euler
angles. It also avoids the ambiguities that arise when using
quaternions to represent the attitude dynamics. As the three-
sphere S
3
double covers SO(3), any attitude feedback con-
troller designed in terms of quaternions could yield different
5423

control inputs depending on the choice of quaternion vectors.
The corresponding stability analysis would need to carefully
consider the fact that convergence to a single attitude implies
convergence to either of the two disconnected, antipodal
points on S
3
[22]. This requires a continuous selection of the
sign of quaternions or a discontinuous control system, which
are shown to be sensitive to small measurement noise [23].
Without these considerations, a quaternion-based controller
can exhibit an unwinding phenomenon, where the controller
unnecessarily rotates the attitude through large angles [15]. In
this paper, the use of rotation matrices in the controller design
and stability analysis completely eliminates these difficulties.
Another novelty of the presented controller is the choice
of the total thrust in (15). This is designed to follow position
tracking commands, but it is also carefully designed to
guarantee the overall stability of the complete dynamics by
feedback control of the direction of the third body-fixed axis.
This consideration is natural as each column of a rotation
matrix represents the direction of each body-fixed axis.
Therefore, another advantage of using rotation matrices is
that the controller has a well-defined physical interpretation.
In Propositions 1 and 3, exponential stability and exponen-
tial attractiveness are guaranteed for almost all initial attitude
errors, respectively. The attitude error function defined in (8)
has the following critical points: the identity matrix, and
rotation matrices that can be written as exp(πˆv) for any
v S
2
. These non-identity critical points of the attitude
error function lie outside of the region of attraction. As it is
a two-dimensional subspace of the three-dimensional SO(3),
we claim that the presented controller exhibits almost global
properties in SO(3). It is impossible to construct a smooth
controller on SO(3) that has global asymptotic stability. The
two-dimensional family of non-identity critical points can be
reduced to four points by modifying the error function to be
1
2
tr[G(I R
T
d
R)] for a matrix G 6= I R
3×3
. The presented
controller can be modified accordingly.
IV. NUMERICAL EXAMPLE
The parameters of the quadrotor UAV are chosen accord-
ing to a quadrotor UAV developed in [2].
J = [0.0820, 0.0845, 0.1377] kgm
2
, m = 4.34 kg
d = 0.315 m, c
τ f
= 8.004 × 10
4
m.
The controller parameters are chosen as follows:
k
x
= 16m, k
v
= 5.6m, k
R
= 8.81, k
= 2.54.
We consider the following two cases.
(I) This maneuver follows an elliptical helix while rotating
the heading direction at a fixed rate. Initial conditions
are chosen as
x(0) = [0, 0, 0], v(0) = [0, 0, 0],
R(0) = I, Ω(0) = [0, 0, 0].
The desired trajectory is as follows.
x
d
(t) = [0.4t, 0.4 sin πt, 0.6 cos πt],
~
b
1
d
(t) = [cos πt, sin πt, 0].
(II) This maneuver recovers from being initially upside
down. Initial conditions are chosen as
x(0) = [0, 0, 0], v(0) = [0, 0, 0],
R(0) =
1 0 0
0 0.9995 0.0314
0 0.0314 0.9995
, Ω(0) = [0, 0, 0].
The desired trajectory is as follows.
x
d
(t) = [0, 0, 0],
~
b
1
d
(t) = [1, 0, 0].
Simulation results are presented in Figures 4 and 5. For
Case (I), the initial value of the attitude error function Ψ(0) is
less than 0.15. This satisfies the conditions for Proposition 2,
and exponential asymptotic stability is guaranteed. As shown
in Figure 4, the tracking errors exponentially converge to
zero. This example illustrates that the proposed controlled
quadrotor UAV can follow a complex trajectory that involve
large angle rotations and nontrivial translations accurately.
In Case (II), the initial attitude error is 178
, which yields
the initial attitude error function Ψ(0) = 1.995 > 1. This
corresponds to Proposition 3, which implies almost global
exponential attractiveness. In Figure 5(b), the attitude error
function Ψ decreases, and it becomes less than 1 at t =
0.88 seconds. After that instant, the position tracking error
and the angular velocity error converge to zero as shown
in Figures 5(c) and 5(d). The region of attraction of the
proposed control system almost covers SO(3), so that the
corresponding controlled quadrotor UAV can recover from
being initially upside down.
V. CONCLUSION
We presented a global dynamic model for a quadrotor
UAV, and we developed a geometric tracking controller
directly on the special Euclidean group that is intrinsic and
coordinate-free, thereby avoiding the singularities of Euler
angles and the ambiguities of quaternions in representing
attitude. It exhibits exponential stability when the initial
attitude error is less than 90
, and it yields almost global
exponentially attractiveness when the initial attitude error is
less than 180
. These are illustrated by numerical examples.
This controller can be extended as follows. In this paper,
four input degrees of freedom are used to track a three-
dimensional position, and a one-dimensional heading direc-
tion. But, without changing the controller structure, they
can be used to follow arbitrary three-dimensional attitude
commands. The remaining one input degree of freedom can
be used to maintain the altitude as much as possible. By
constructing a hybrid controller based on these two tracking
modes, we can generate complicated acrobatic maneuvers of
a quadrotor UAV.
REFERENCES
[1] M. Valenti, B. Bethke, G. Fiore, and J. How, “Indoor multi-vehicle
flight testbed for fault detection, indoor multi-vehicle flight testbed for
fault detection, isolation, and recovery, in Proceedings of the AIAA
Guidance, Navigation and Control Conference, 2006.
5424

Citations
More filters
Proceedings ArticleDOI
09 May 2011
TL;DR: An algorithm is developed that enables the real-time generation of optimal trajectories through a sequence of 3-D positions and yaw angles, while ensuring safe passage through specified corridors and satisfying constraints on velocities, accelerations and inputs.
Abstract: We address the controller design and the trajectory generation for a quadrotor maneuvering in three dimensions in a tightly constrained setting typical of indoor environments. In such settings, it is necessary to allow for significant excursions of the attitude from the hover state and small angle approximations cannot be justified for the roll and pitch. We develop an algorithm that enables the real-time generation of optimal trajectories through a sequence of 3-D positions and yaw angles, while ensuring safe passage through specified corridors and satisfying constraints on velocities, accelerations and inputs. A nonlinear controller ensures the faithful tracking of these trajectories. Experimental results illustrate the application of the method to fast motion (5–10 body lengths/second) in three-dimensional slalom courses.

1,875 citations


Cites background from "Geometric tracking control of a qua..."

  • ...Note that spatial scaling also applies to the problem with multiple keyframes but the property is less useful as the positions of all keyframes must be scaled by the same factor....

    [...]

Book ChapterDOI
01 Apr 2016
TL;DR: This work presents a method of jointly optimizing polynomial path segments in an unconstrained quadratic program that is numerically stable for high-order polynomials and large numbers of segments, and is easily formulated for efficient sparse computation.
Abstract: We explore the challenges of planning trajectories for quadrotors through cluttered indoor environments. We extend the existing work on polynomial trajectory generation by presenting a method of jointly optimizing polynomial path segments in an unconstrained quadratic program that is numerically stable for high-order polynomials and large numbers of segments, and is easily formulated for efficient sparse computation. We also present a technique for automatically selecting the amount of time allocated to each segment, and hence the quadrotor speeds along the path, as a function of a single parameter determining aggressiveness, subject to actuator constraints. The use of polynomial trajectories, coupled with the differentially flat representation of the quadrotor, eliminates the need for computationally intensive sampling and simulation in the high dimensional state space of the vehicle during motion planning. Our approach generates high-quality trajecrtories much faster than purely sampling-based optimal kinodynamic planning methods, but sacrifices the guarantee of asymptotic convergence to the global optimum that those methods provide. We demonstrate the performance of our algorithm by efficiently generating trajectories through challenging indoor spaces and successfully traversing them at speeds up to 8 m/s. A demonstration of our algorithm and flight performance is available at: http://groups.csail.mit.edu/rrg/quad_polynomial_trajectory_planning.

578 citations


Cites methods from "Geometric tracking control of a qua..."

  • ...The nonlinear controller employed to follow differentiable trajectories was developed in [18], and consists of independent calculations for thrust and moments:...

    [...]

Book ChapterDOI
01 Jan 2016
TL;DR: This chapter presents a modular Micro Aerial Vehicle (MAV) simulation framework, which enables a quick start to perform research on MAVs, and is a good starting point to tackle higher level tasks, such as collision avoidance, path planning, and vision based problems, like Simultaneous Localization and Mapping (SLAM), on MAV.
Abstract: In this chapter we present a modular Micro Aerial Vehicle (MAV) simulation framework, which enables a quick start to perform research on MAVs. After reading this chapter, the reader will have a ready to use MAV simulator, including control and state estimation. The simulator was designed in a modular way, such that different controllers and state estimators can be used interchangeably, while incorporating new MAVs is reduced to a few steps. The provided controllers can be adapted to a custom vehicle by only changing a parameter file. Different controllers and state estimators can be compared with the provided evaluation framework. The simulation framework is a good starting point to tackle higher level tasks, such as collision avoidance, path planning, and vision based problems, like Simultaneous Localization and Mapping (SLAM), on MAVs. All components were designed to be analogous to its real world counterparts. This allows the usage of the same controllers and state estimators, including their parameters, in the simulation as on the real MAV.

421 citations

Journal ArticleDOI
TL;DR: The recent work on micro unmanned aerial vehicles (UAVs), a fast-growing field in robotics, is surveyed, outlining the opportunities for research and applications, along with the scientific and technological challenges.
Abstract: We survey the recent work on micro unmanned aerial vehicles (UAVs), a fast-growing field in robotics, outlining the opportunities for research and applications, along with the scientific and technological challenges. Micro-UAVs can operate in three-dimensional environments, explore and map multi-story buildings, manipulate and transport objects, and even perform such tasks as assembly. While fixed-base industrial robots were the main focus in the first two decades of robotics, and mobile robots enabled most of the significant advances during the next two decades, it is likely that UAVs, and particularly micro-UAVs, will provide a major impetus for the next phase of education, research, and development.

409 citations


Cites background or result from "Geometric tracking control of a qua..."

  • ...Lee et al. (2010) showed that the two controllers result in a nonlinear controller that explicitly track trajectories in SE( 3)....

    [...]

  • ...A similar result with rotation matrices is available in Lee et al. (2010)....

    [...]

  • ...…SO( 3) and compensating for the nonlinear inertial terms gives us u2 = J ( −kReR − k e ) + × J −J ( ̂RT Rdes des − RT Rdeṡdes ) (5) If we do not consider constraints on the state or the inputs, Equations (4)–(5) achieve asymptotic convergence to specified trajectories in SE( 3) (Lee et al., 2010)....

    [...]

  • ...As shown by Lee et al. (2010), in the absence of constraints on inputs, the maximum permissible error on the angular velocity for stable hover is proportional to the inverse of the square root of the largest moment of inertia....

    [...]

  • ...If we do not consider constraints on the state or the inputs, Equations (4)–(5) achieve asymptotic convergence to specified trajectories in SE( 3) (Lee et al., 2010)....

    [...]

Journal ArticleDOI
02 Feb 2017
TL;DR: This work proposes a method to formulate trajectory generation as a quadratic program (QP) using the concept of a Safe Flight Corridor (SFC), a collection of convex overlapping polyhedra that models free space and provides a connected path from the robot to the goal position.
Abstract: There is extensive literature on using convex optimization to derive piece-wise polynomial trajectories for controlling differential flat systems with applications to three-dimensional flight for Micro Aerial Vehicles. In this work, we propose a method to formulate trajectory generation as a quadratic program (QP) using the concept of a Safe Flight Corridor (SFC). The SFC is a collection of convex overlapping polyhedra that models free space and provides a connected path from the robot to the goal position. We derive an efficient convex decomposition method that builds the SFC from a piece-wise linear skeleton obtained using a fast graph search technique. The SFC provides a set of linear inequality constraints in the QP allowing real-time motion planning. Because the range and field of view of the robot's sensors are limited, we develop a framework of Receding Horizon Planning , which plans trajectories within a finite footprint in the local map, continuously updating the trajectory through a re-planning process. The re-planning process takes between 50 to 300 ms for a large and cluttered map. We show the feasibility of our approach, its completeness and performance, with applications to high-speed flight in both simulated and physical experiments using quadrotors.

344 citations


Cites background from "Geometric tracking control of a qua..."

  • ...Thus, Φi(t), d dt Φi(t), d2 dt2 Φi(t), d3 dt3 Φi(t) indicate the desired position, velocity, acceleration and jerk at time t, which are the input for the non-linear controller [15] to calculate desired force and momentum for controlling the quadrotor....

    [...]

  • ...We assume the robot is able to follow our generated trajectories through a non-linear controller [15]....

    [...]

References
More filters
Book
01 Jan 2003
TL;DR: In this paper, the authors propose energy-based methods for stabilizing nonholonomic systems using non-holonomic control theory based on geometric properties of the system's properties. But they do not discuss the energy-independent methods of stabilisation.
Abstract: Introduction.- Mathematical Preliminaries.- Basic Concepts in Geometric Mechanics.- Introduction to Aspects of Geometric Control Theory.- Nonholonomic Mechanics.- Control of Mechanical and Nonholonomic Systems.- Optimal Control.- Stability of Nonholonomic Systems.- Energy-Based Methods for Stabilization.- References.- Index.

1,328 citations


"Geometric tracking control of a qua..." refers background in this paper

  • ...Geometric control is concerned with the development of control systems for dynamic systems evolving on nonlinear manifolds that cannot be globally identified with Euclidean spaces [12], [ 13 ], [14]....

    [...]

Proceedings ArticleDOI
20 Aug 2007
TL;DR: In this paper, a theoretical development is presented, and validated through both thrust test stand measurements and vehicle flight tests using the Stanford Testbed of Autonomous Rotorcraft for Multi-Agent Control (STARMAC) quadrotor helicopter.
Abstract: Quadrotor helicopters are emerging as a popular platform for unmanned aerial vehicle (UAV) research, due to the simplicity of their construction and maintenance, their ability to hover, and their vertical take o and landing (VTOL) capability. Current designs have often considered only nominal operating conditions for vehicle control design. This work seeks to address issues that arise when deviating significantly from the hover flight regime. Aided by well established research for helicopter flight control, three separate aerodynamic eects are investigated as they pertain to quadrotor flight, due to vehicular velocity, angle of attack, and airframe design. They cause moments that aect attitude control, and thrust variation that aects altitude control. Where possible, a theoretical development is first presented, and is then validated through both thrust test stand measurements and vehicle flight tests using the Stanford Testbed of Autonomous Rotorcraft for Multi-Agent Control (STARMAC) quadrotor helicopter. The results enabled improved controller performance.

1,074 citations


"Geometric tracking control of a qua..." refers background in this paper

  • ...There are several university-level projects [1], [2], [3], [4], and commercial products [5], [6], [7] related to the development and application of quadrotor UAVs....

    [...]

  • ...Linear control systems such as proportionalderivative controllers or linear quadratic regulators are widely used to enhance the stability properties of an equilibrium [1], [3], [4], [8], [9]....

    [...]

Book
01 Dec 1996
TL;DR: In this paper, the authors present a generalization of the Hamilton-Jacobi theory for systems on Lie groups and homogenous spaces, including linear and polynomial control systems with quadratic costs.
Abstract: Introduction Acknowledgments Part I. Reachable Sets and Controllability: 1. Basic formalism and typical problems 2. Orbits of families of vector fields 3. Reachable sets of Lie-determined systems 4. Control affine systems 5. Linear and polynomial control systems 6. Systems on Lie groups and homogenous spaces Part II. Optimal Control Theory: 7. Linear systems with quadratic costs 8. The Riccati equation and quadratic systems 9. Singular linear quadratic problems 10. Time-optimal problems and Fuller's phenomenon 11. The maximum principle 12. Optimal problems on Lie groups 13. Symmetry, integrability and the Hamilton-Jacobi theory 14. Integrable Hamiltonian systems on Lie groups: the elastic problem, its non-Euclidean analogues and the rolling-sphere problem References Index.

1,066 citations

Proceedings ArticleDOI
18 Apr 2005
TL;DR: The results of two nonlinear control techniques applied to an autonomous micro helicopter called Quadrotor are presented, a backstepping and a sliding-mode techniques.
Abstract: The latest technological progress in sensors, actuators and energy storage devices enables the developments of miniature VTOL systems. In this paper we present the results of two nonlinear control techniques applied to an autonomous micro helicopter called Quadrotor. A backstepping and a sliding-mode techniques. We performed various simulations in open and closed loop and implemented several experiments on the test-bench to validate the control laws. Finally, we discuss the results of each approach. These developments are part of the OS4 project in our lab.

1,010 citations


"Geometric tracking control of a qua..." refers methods in this paper

  • ...Backstepping and sliding mode technique s are applied in [11]....

    [...]

Journal ArticleDOI
TL;DR: A new quaternion-based feedback control scheme for exponential attitude stabilization of a four-rotor vertical takeoff and landing aerial robot known as a quadrotor aircraft is proposed and the model-independent PD controller, without compensation of the Coriolis and gyroscopic torques, provides asymptotic stability for the problem.
Abstract: In this paper, we propose a new quaternion-based feedback control scheme for exponential attitude stabilization of a four-rotor vertical takeoff and landing aerial robot known as a quadrotor aircraft. The proposed controller is based upon the compensation of the Coriolis and gyroscopic torques and the use of a PD/sup 2/ feedback structure, where the proportional action is in terms of the vector quaternion and the two derivative actions are in terms of the airframe angular velocity and the vector quaternion velocity. We also show that the model-independent PD controller, where the proportional action is in terms of the vector-quaternion and the derivative action is in terms of the airframe angular velocity, without compensation of the Coriolis and gyroscopic torques, provides asymptotic stability for our problem. The proposed controller as well as some other controllers have been tested experimentally on a small-scale quadrotor aircraft.

982 citations


"Geometric tracking control of a qua..." refers background in this paper

  • ...All of these assumptions are common [19], [4], and the presented control system can readily be extended to include linear rotor dynamics studied in [11]....

    [...]

Frequently Asked Questions (20)
Q1. What are the contributions in "Geometric tracking control of a quadrotor uav on se(3)" ?

This paper provides new results for the tracking control of a quadrotor unmanned aerial vehicle ( UAV ). A globally defined model of the quadrotor UAV rigid body dynamics is introduced as a basis for the analysis. 

Since the direction of the total thrust is the third bodyfixed axis, the stability of the translational dynamics depends on the attitude tracking error. 

Since it is assumed that the first and the third propellers rotate clockwise, and the second and the fourth propellers rotate counterclockwise, when they are generating a positive thrust fi along −~b3, the torque generated by the i-th propeller can be written as τi = (−1)icτffi for a fixed constant cτf . 

In short, this control system is designed so that the position tracking error converges to zero when there is no attitude tracking error, and it is properly adjusted for nonzero attitude tracking errors to achieve asymptotic stability of the complete dynamics. 

Since the region of attraction given by (28) for the attitude almost covers SO(3), this is referred to as almost global exponential attractiveness in this paper. 

The total thrust f and the desired direction ~b3d of the third body-fixed axis are chosen such that if there is no attitude tracking error, the thrust vector term −fRe3 in the translational dynamics of (3) becomes this tracking controller in R3. 

For any positive constants kx, kv , the authors choose positive constants c1, c2, kR, kΩ such thatc1 < min { kv(1− α),4mkxkv(1− α) k2v(1 + α) 2 + 4mkx , √ kxm } ,(24)c2 < min { kΩ,4kΩkRλmin(J) 2k2Ωλmax(J) + 4kRλmin(J) 2 ,√ kRλmin(J),√ 22− ψ1 kRλmax(J)} , (25)λmin(W2) > 4‖W12‖2λmin(W1) . (26)Then, the zero equilibrium of the tracking errors of the complete system is exponentially stable. 

The translational dynamics of a quadrotor UAV is controlled by the total thrust −fRe3, where the magnitude of the total thrust f is directly controlled, and the direction of the total thrust −Re3 is along the third body-fixed axis −~b3. 

As the attitude tracking error increases, the direction of the control input term fRe3 of the translational dynamics deviates from the desired direction Rde3. 

This requires a continuous selection of the sign of quaternions or a discontinuous control system, which are shown to be sensitive to small measurement noise [23]. 

For a given translational command xd(t), the authors select the total thrust f , and the desired direction of the third body-fixed axis ~b3d to stabilize the translational dynamics. 

The tracking errors for the position and the velocity are given by:tex = x− xd, (6) ev = v − vd. (7)The attitude and angular velocity tracking error should be carefully chosen as they evolve on the tangent bundle of the nonlinear space SO(3). 

The configuration of this quadrotor UAV is defined by the location of the center of mass and the attitude with respect to the inertial frame. 

According to the definition of the rotation matrix R ∈ SO(3), the direction of the i-th body fixed axis~bi is given by Rei in the inertial frame, where e1 = [1; 0; 0], e2 = [0; 1; 0], e3 = [0; 0; 1] ∈ R3. 

The region of attraction of the proposed control system almost covers SO(3), so that the corresponding controlled quadrotor UAV can recover from being initially upside down. 

(8)This is locally positive-definite about R = Rd within the region where the rotation angle between R and Rd is less than 180◦ [14]. 

Define m ∈ R the total mass J ∈ R3×3 the inertia matrix with respect to the body-fixed frame R ∈ SO(3) the rotation matrix from the body-fixed frame to the inertial frame Ω ∈ R3 the angular velocity in the body-fixed frame978-1-4244-7744-9/10/$26.00 ©2010 IEEE 5420x ∈ R3 the location of the center of mass in the inertial frame v ∈ R3 the velocity of the center of mass in the inertial frame d ∈ 

the region of attraction for the attitude almost covers SO(3), and the region of attraction for the angular velocity can be increased by choosing a larger controller gain kR in (18). 

The equations of motion of this quadrotor UAV can be written asẋ = v, (2) mv̇ = mge3 − fRe3, (3)Ṙ = RΩ̂, (4)JΩ̇ + Ω× JΩ = M, (5)where the hat map ·̂ : R3 → so(3) is defined by the condition that x̂y = x× y for all x, y ∈ R3.III. 

R the distance from the center of mass to the center of each rotor in the ~b1,~b2 plane fi ∈ R the thrust generated by the i-th propeller along the −~b3 axis τi ∈ R the torque generated by the i-th propeller about the ~b3 axisf ∈ R the total thrust, i.e., f = ∑4 i=1 fiM ∈ R3 the total moment in the body-fixed frame