introduction to feedback control of underactuated vtol ... · introduction to feedback control of...

13
Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin, Claude Samson Abstract—The paper is an introduction to feedback control design for a family of robotic aerial vehicles with Vertical Take- Off and Landing (VTOL) capabilities such as quadrotors, ducted- fan tail-sitters, helicopters, etc. Potential applications for such devices, like surveillance, monitoring, or mapping, are varied and numerous. For these applications to emerge, motion control algorithms that guarantee a good amount of robustness against state measurement/estimation errors and unmodeled dynamics like, e.g., aerodynamic perturbations, are needed. The feedback control methods here considered range from basic linear control schemes to more elaborate nonlinear control solutions. The modeling of the dynamics of these systems is first recalled and discussed. Then, several control algorithms are presented and commented upon in relation to implementation issues and various operating modes encountered in practice, from teleoperated to fully autonomous flight. Particular attention is paid to the incorporation of integral-like control actions, often overlooked in nonlinear control studies despite their practical importance to render the control performance more robust with respect to unmodeled or poorly estimated additive perturbations. I. I NTRODUCTION The growing interest for the robotics research community in Vertical Take-Off and Landing (VTOL) vehicles as ex- emplified by quadrotors [10], [19], [21], [23], [65], [82], ducted-fan tail-sitters [43], [58], [63], and helicopters [16], [32], [39], is partly due to the numerous applications that can be addressed with such systems like surveillance, in- spection, or mapping. Recent technological advances in sen- sors, batteries and processing cards, that allow embarking on small vehicles all components necessary for autonomous flights at a reasonable cost, also constitute a favorable factor. The development of these applications raises several issues. First, from a mechanical viewpoint, the vehicle must have good flying capabilities and provide with enough payload for embarked sensors and energy. Then, real-time pose es- timation, necessary for feedback loops implementation, is often challenging. Since in a non-instrumented environment no single sensor can produce direct measurements of the pose, a combination of sensors has to be used. Data fusion algorithms must be designed to cope with sensor limitations Corresponding author: Minh-Duc Hua, Institut des Syst` emes Intelligents et de Robotiques (ISIR), Universit´ e Pierre et Marie Curie, Boite courrier 173, 4 place Jussieu, 75252 Paris cedex 05, France. Tel:+33144276362. Minh-Duc Hua is with ISIR CNRS-UPMC (France) and JVN-VNUHCM (Vietnam). E-mail: hua@isir.upmc.f r. Tarek Hamel is with I3S UNS-CRNS (France). E-mail: thamel@i3s.unice.f r. Pascal Morin is with ISIR-UPMC (France). E-mail: morin@isir.upmc.f r. Claude Samson is with INRIA Sophia Antipolis-M´ editerran´ ee (France). E- mail: Claude.Samson@inria.f r. associated with, for instance, unreliable magnetometers, low- precision GPS devices, and low-frequency visual processing of camera images. Finally, feedback control laws must achieve fast response and robustness to aerodynamic perturbations (i.e., wind gusts). All these difficulties are amplified on small-scale vehicles due to the limited payload, the high sensitivity to aerodynamic perturbations and the complexity of aerodynamic effects at this scale. Many autonomous flight experiments have been reported in recent years. In particular, aggressive maneuvers have been achieved with quadrotors, like spins and flips [45], dance in the air [75], or high-speed flight through narrow environments [54]. The control algorithms used for these extreme maneuvers much rely on feedforward action in order to account for and anticipate strong dynamic effects. This type of performance has been obtained in indoor environments by taking advantage of full state measurements acquired at high-frequency rates with an external 3D tracking system. Reported results for aerial robotic vehicles using onboard sensors only, such as GPS, laser range finders, cameras, or acoustic sensors, address far less aggressive maneuvers [29], [9], [19], [23], [62], [77] because the system then suffers from a heavier payload (sensors and data processing cards) which reduces its maneuverability, and also from low-frequency measurement rates and poor quality state estimates little compatible with the monitoring of aggressive maneuvers. The present paper focuses essentially on feedback control aspects, knowing that mechanical design and state estima- tion aspects are equally important. The main objective is to provide the reader with an introduction to feedback control methods, linear and nonlinear, developed for translational velocity and/or position stabilization and trajectory tracking applications. Whereas linearization of the system’s dynamics yields well-known control analysis and design tools, that are local in scope, the nonlinear approach exploits the physics of the motion control problem at hand in a more natural way. This aspect reflects in algorithms which are not more complex, and a computational effort not more important than that of linear controllers. Under certain conditions, nonlinear controllers can also be endowed with provable global stability and convergence properties. As for robustness, although no robustness analysis per say is reported, it is a permanent preoccupation whose related issues are discussed at various places in the paper. For VTOL vehicles, these issues are all the more important that: precise aerodynamic models, valid in a large range of operating conditions, are extremely difficult to obtain; wind gusts are unpredictable, and embarked power is not always sufficient to counterbalance the destabilizing

Upload: others

Post on 24-Jan-2020

28 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

Introduction to Feedback Control of UnderactuatedVTOL Vehicles

Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin, Claude Samson

Abstract—The paper is an introduction to feedback controldesign for a family of robotic aerial vehicles with Vertical Take-Off and Landing (VTOL) capabilities such as quadrotors, ducted-fan tail-sitters, helicopters, etc. Potential applications for suchdevices, like surveillance, monitoring, or mapping, are variedand numerous. For these applications to emerge, motion controlalgorithms that guarantee a good amount of robustness againststate measurement/estimation errors and unmodeled dynamicslike, e.g., aerodynamic perturbations, are needed. The feedbackcontrol methods here considered range from basic linear controlschemes to more elaborate nonlinear control solutions. Themodeling of the dynamics of these systems is first recalled anddiscussed. Then, several control algorithms are presented andcommented upon in relation to implementation issues and variousoperating modes encountered in practice, from teleoperatedto fully autonomous flight. Particular attention is paid to theincorporation of integral-like control actions, often overlookedin nonlinear control studies despite their practical importanceto render the control performance more robust with respect tounmodeled or poorly estimated additive perturbations.

I. INTRODUCTION

The growing interest for the robotics research communityin Vertical Take-Off and Landing (VTOL) vehicles as ex-emplified by quadrotors [10], [19], [21], [23], [65], [82],ducted-fan tail-sitters [43], [58], [63], and helicopters [16],[32], [39], is partly due to the numerous applications thatcan be addressed with such systems like surveillance, in-spection, or mapping. Recent technological advances in sen-sors, batteries and processing cards, that allow embarkingon small vehicles all components necessary for autonomousflights at a reasonable cost, also constitute a favorable factor.The development of these applications raises several issues.First, from a mechanical viewpoint, the vehicle must havegood flying capabilities and provide with enough payloadfor embarked sensors and energy. Then, real-time pose es-timation, necessary for feedback loops implementation, isoften challenging. Since in a non-instrumented environmentno single sensor can produce direct measurements of thepose, a combination of sensors has to be used. Data fusionalgorithms must be designed to cope with sensor limitations

Corresponding author: Minh-Duc Hua, Institut des Systemes Intelligents etde Robotiques (ISIR), Universite Pierre et Marie Curie, Boite courrier 173, 4place Jussieu, 75252 Paris cedex 05, France. Tel:+33144276362.

Minh-Duc Hua is with ISIR CNRS-UPMC (France) and JVN-VNUHCM(Vietnam). E-mail: [email protected].

Tarek Hamel is with I3S UNS-CRNS (France). E-mail:[email protected].

Pascal Morin is with ISIR-UPMC (France). E-mail:[email protected].

Claude Samson is with INRIA Sophia Antipolis-Mediterranee (France). E-mail: [email protected].

associated with, for instance, unreliable magnetometers, low-precision GPS devices, and low-frequency visual processingof camera images. Finally, feedback control laws must achievefast response and robustness to aerodynamic perturbations (i.e.,wind gusts). All these difficulties are amplified on small-scalevehicles due to the limited payload, the high sensitivity toaerodynamic perturbations and the complexity of aerodynamiceffects at this scale.

Many autonomous flight experiments have been reported inrecent years. In particular, aggressive maneuvers have beenachieved with quadrotors, like spins and flips [45], dance inthe air [75], or high-speed flight through narrow environments[54]. The control algorithms used for these extreme maneuversmuch rely on feedforward action in order to account for andanticipate strong dynamic effects. This type of performancehas been obtained in indoor environments by taking advantageof full state measurements acquired at high-frequency rateswith an external 3D tracking system. Reported results for aerialrobotic vehicles using onboard sensors only, such as GPS, laserrange finders, cameras, or acoustic sensors, address far lessaggressive maneuvers [29], [9], [19], [23], [62], [77] becausethe system then suffers from a heavier payload (sensors anddata processing cards) which reduces its maneuverability,and also from low-frequency measurement rates and poorquality state estimates little compatible with the monitoringof aggressive maneuvers.

The present paper focuses essentially on feedback controlaspects, knowing that mechanical design and state estima-tion aspects are equally important. The main objective is toprovide the reader with an introduction to feedback controlmethods, linear and nonlinear, developed for translationalvelocity and/or position stabilization and trajectory trackingapplications. Whereas linearization of the system’s dynamicsyields well-known control analysis and design tools, that arelocal in scope, the nonlinear approach exploits the physicsof the motion control problem at hand in a more naturalway. This aspect reflects in algorithms which are not morecomplex, and a computational effort not more important thanthat of linear controllers. Under certain conditions, nonlinearcontrollers can also be endowed with provable global stabilityand convergence properties. As for robustness, although norobustness analysis per say is reported, it is a permanentpreoccupation whose related issues are discussed at variousplaces in the paper. For VTOL vehicles, these issues are allthe more important that:• precise aerodynamic models, valid in a large range of

operating conditions, are extremely difficult to obtain;• wind gusts are unpredictable, and embarked power is

not always sufficient to counterbalance the destabilizing

Page 2: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

2

effects of these perturbations;• pose measurement/estimation errors, in relation to the size

and inertia of the vehicle, can be significant.The paper is organized as follows. Control models are

reviewed in Section II, with complementary comments onthe contribution of aerodynamic forces and torques. SectionsIII and IV contain an overview on existing feedback controltechniques developed for the class of underactuated VTOLvehicles. First, in Section III linear techniques based onlinearized control models are reviewed. The classical poleplacement method invoked for the determination of the controlgains can be complemented with an optimization policy (LQR,LQG, H2, H∞, etc.) and also extended to slowly time-varyingoperating conditions via gain-scheduling techniques. In Sec-tion IV nonlinear techniques (exact input-output feedbacklinearization, backstepping, hierarchical control, etc.) are pre-sented. This section contains also a more detailed expositionof the control approach proposed in [26]. Finally, concludingremarks and perspectives are given in Section V. Suggestionsof references for complementary reading are pointed out alongthe paper.

II. SYSTEM MODELING

In the first approximation, VTOL vehicles can be modeledas rigid bodies immersed in a fluid and moving in 3D-space.These vehicles are usually controlled via a thrust force ~T alonga body-fixed direction ~k to create translational motion and atorque vector Γ ∈ R3 for attitude monitoring. In practice,the torque actuation is typically generated via propellers (e.g.,multirotors), rudders or flaps (e.g., ducted-fan tail-sitters), ora swashplate mechanism (e.g., helicopters).

Fig. 1. Inertial frame I and body frame B.

We assume that the thrust ~T = −T~k, with T ∈ R, appliesat a point that lies on, or close to, the axis {G;~k}, with Gthe vehicle’s center of mass (CoM) so that it does not createan important torque at G. All external forces acting on thevehicle (gravity and buoyancy forces, added-mass forces, anddissipative aerodynamic reaction forces, etc.) are summed upin a vector ~Fe. Due to aerodynamic reaction forces, this vectorgenerally depends on the airflow’s velocity and directionrelatively to the vehicle, as well as on its translational andangular accelerations (via added-mass effects). Applying theNewton-Euler formalism, one obtains the following equations

of motion of the vehicle (see, e.g., [17, Ch. 2], [78, Ch. 1],[26]):

mξ = −TRe3 + Fe(ξ, ξ, R, ω, ω, d(t)) +RΣRΓ (1)

R = RS(ω) (2)

Iω = −S(ω)Iω+Γ+Γe(ξ, ξ, R, ω, ω, d(t)) + τg+ΣTTe3

(3)

where the following notation is used:• m is the vehicle’s mass, I ∈ R3×3 its inertia matrix.• I={O;~ıo,~o,~ko} is a fixed (inertial or Galilean) frame

with respect to which the vehicle’s absolute pose (position+ orientation) is measured. This frame is typically chosenas the NED frame (North-East-Down) with ~ıo pointingto the North, ~o pointing to the East, and ~ko pointing tothe center of the earth. B={G;~ı,~,~k} is a frame attachedto the body. The vector ~k is parallel to the thrust forceaxis. This leaves two possible and opposite directionsfor this vector. The direction here chosen (~k pointingdownward nominally) is consistent with the conventionused for VTOL vehicles (see Fig. 1).

• ξ = (ξ1, ξ2, ξ3)> ∈ R3 is the vector of coordinates of thevehicle’s CoM position expressed in the inertial frame I.

• R ∈ SO(3) is the rotation matrix representing theorientation of the body-fixed frame B with respect to theinertial frame I. The column vectors of R correspondto the vectors of coordinates of ~ı, ~, ~k expressed in thebasis of I.

• ω = (ω1, ω2, ω3)> ∈ R3 is the angular velocity vectorof the body-fixed frame B relative to the inertial frame Iand expressed in B.

• Γe is the external torque vector induced by all externalforces.

• S(·) is the skew-symmetric matrix associated with thecross product (i.e., S(u)v = u× v,∀u, v ∈ R3).

• d(t) represents external disturbances, including wind ef-fect, which do not depend on the vehicle’s position andmotion.

• τg is the gyroscopic torque associated with rotor crafts.• e3 = (0, 0, 1)> is the third vector of the canonical basis

of R3 and also the vector of coordinates in B of the thrustdirection vector ~k.

• ΣT and ΣR denote 3× 3 (approximately) constant cou-pling matrices.

To be complete, the dynamical model should also include amodeling of the generation of force and torque control inputs,since these inputs are typically produced via actuators whichhave their own dynamics. Experience shows that the actuators’dynamics of a well-designed VTOL system are sufficientlyfast with respect to the vehicle’s dynamics so that they can beneglected, at least in the first approximation. Another reasonfor considering this approximation is that it allows for thedecoupling of the generic aspects of the control problem fromthe specifics attached to each particular vehicle. Nevertheless,the value of this decoupling assumption needs in practiceto be re-assessed and, eventually, re-considered in order tobring adequate modifications and refinements to the controlalgorithms. For instance, in the case of vehicles with propellers

Page 3: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

3

and deflecting surfaces, the way actuation thrust force andtorques depend on the airflow needs to be looked at closely.Advanced specialized issues like this one, addressed in [29]for instance, call for complementary studies which are beyondthe scope of the present paper.

A. About the coupling between thrust and actuation torques

The constant matrices ΣT and ΣR in Eqs. (1) and (3) rep-resent the coupling between the thrust force and the actuatedtorques. For instance, the term ΣTTe3 in Eq. (3) accountsfor the fact that the thrust force vector ~T may not applyexactly at the vehicle’s CoM. For most VTOL vehicles thisterm is relatively small so that it can be compensated for bycontrol torque actions. By contrast, the influence of the torquecontrol inputs on the translational dynamics via the couplingmatrix ΣR in Eq. (1), whose expression mainly depends on thevehicle’s torque generation mechanism, is more involved. Thecoupling term RΣRΓ is often referred to as small body forcesin the literature. This coupling is negligible (i.e., ΣR ≈ 0) forquadrotors [21], [65], [10] such as the CEA-List X4-flyer (seeFig. 2), but it can be significant for helicopters, due to theswashplate mechanism [24, Ch.1], [15], [39], [46], [59, Ch.5], and for ducted-fan tail-sitters, due to the rudder system[60, Ch. 3], [63]. For instance, the coupling matrix ΣR of aconventional helicopter is approximately given by (see, e.g.,[24, Ch.1], [59, Ch.5], [39])

ΣR =

0 ε1 0ε2 ε3 ε4

0 0 0

where the εi’s depend on the position of the main and tailrotor disks. Referring to the model of ducted-fan tail-sitters,such as the HoverEye (see Fig. 3), the coupling matrix ΣR isgiven by (see, e.g., [63], [60, Ch. 3])

ΣR = − 1L

S(e3), (4)

with L the distance between the plane of controlled fins andthe vehicle’s CoM. More details on the difficulties associatedwith this coupling term and control solutions proposed in theliterature will be provided further on.

Fig. 2. X4-flyer: quadrotor developed by CEA-List.

Fig. 3. Ducted-fan tail-sitter HoverEye of Bertin Technologies.

B. About external forces and torques

The modeling of the external force Fe and torque Γe actingon the vehicle is a major problem due to the complexity offluid dynamics and of the interactions between the vehicle, theactuators, and the surrounding fluid (for more details about thisissue, see, e.g., [65], [29], [10] for quadrotors, [35], [38], [60,Ch. 3], [61] for ducted-fan tail-sitters, and [56], [67], [83] forhelicopters). In particular, when the vehicle is designed so as totake advantage of strong aerodynamic lift forces, as in the caseof airplanes, modeling the dependence of lift and drag forcesupon the vehicle’s orientation relatively to the airflow direction(i.e., the so-called angle of attack and side-slip angle) is not asimple matter. To our knowledge, no analytical representationof these forces over the vehicle’s entire operating envelop hasever been worked out. Existing models are often based ona superposition principle which consists in i) decomposingthe vehicle in distinct rigid parts (propellers, rudders, duct,etc.), ii) computing a force/torque acting on each of these partsby neglecting the aerodynamic effects created by the otherparts, and iii) summing up all these forces/torques to obtain theresultant force/torque. For ducted-fan tail-sitters for example,the vehicle is decomposed into the duct structure, the propeller,and rudders. Aerodynamics forces include lift and drag forces(and resp. momentum drag force) generated by the airflowcirculating outside (resp. through) the duct structure, as well aslift and drag forces on the propeller and rudders generated bythe airflow inside the duct. Evaluating these forces separatelyis hardly justified since, rigorously, the vehicle moves in a fluidwhich exerts, by the rotation of the propeller and by airflowdeflection from the rudders, a force on the fluid which in turnapplies a force on the duct structure.

Good understanding of aerodynamic effects and fine mod-eling of these effects are crucial to design the vehicle’s ge-ometrical and mechanical characteristics in order to optimizethe system’s maneuverability and energetic efficiency. Theyare also useful for simulation purposes in order to evaluatethe performance, robustness, and also the limitations of acontroller. For control design, however, the knowledge of aprecise and well-tuned model is not as important. A classicalreason is that a well-designed feedback control is expectedto grant robustness in the sense of performance insensitivitywith respect to model inaccuracies. In this respect, energydissipation due to friction also tends to be a favorable factor.

Page 4: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

4

In practice, it is possible to use an approximate model ora real-time estimation of the terms Fe and Γe. The latterpossibility is preferred when adequate and accurate on-boardsensors allow for it. Moreover, experiments show that in mostcases an explicit model of Γe is not necessary. The possibilityof obtaining a good on-line estimation of Γe, via dynamicinversion or high-gain observer, mainly depends on the qualityand frequency acquisition of the vehicle’s attitude estimationand of the measurement of its angular velocity.

C. A simplified control model

The control model (1)–(3) is very general, but it involvesterms –aerodynamic forces and torques in particular– whichare very complex. For many VTOL vehicles, simplified modelsof these terms can be sufficiently accurate for feedback controldesign purposes. To keep the exposition simple without impair-ing its generality too much, feedback control will be discussedon the basis of the following simplified control model:

mξ = −TRe3 + Fe(ξ, d(t)) +RΣRΓ (5)

R = RS(ω) (6)Iω = −S(ω)Iω+Γ (7)

Assumptions under which the reduction of system (1)–(3) tothe above form is justified are the following:• Added mass effects are negligible (i.e., Fe and Γe do not

depend on ξ and ω);• Aerodynamic forces do not depend on the vehicle’s

orientation and they apply close to the vehicle’s CoM(i.e., Fe and Γe depend neither on R nor on ω). Thisassumption corresponds ideally to the case of a sphericalvehicle with a perfectly smooth surface, for which theresultant of drag forces applies at the CoM;

• Both τg and ΣT are either negligible or known so thatthey can be directly compensated for by the control torqueaction.

These assumptions, although restrictive, hold with a reason-ably good degree of accuracy for many VTOL vehicles forwhich the resultant aerodynamic reaction force is, in nominalconditions, dominated by the drag component, as in the caseof quadrotors, ducted fans, helicopters, etc.

System (5)–(7) shows full actuation of the rotational dy-namics, via the 3-dimensional torque control vector Γ, andunderactuation of the translational dynamics, via the mono-dimensional thrust intensity T . At the translational level, in thecase where ΣR = 0, controllability results from the nonlinearcoupling term TRe3 between the control input T and thevehicle’s orientation characterized by the rotation matrix R.Feedback control thus has to exploit this coupling in orderto stabilize the vehicle’s translational dynamics. The nextsections gives an overview of feedback methods, starting withlinear control schemes.

III. BASIC CONTROL DESIGN FOR VTOL VEHICLES:LINEAR CONTROL SCHEMES

Linear control techniques are based on linear approxima-tions of the system’s dynamics about desired feasible state

trajectories. They are widely used in airplanes autopilots. Forthese vehicles, the angle of attack over the nominal flightenvelop is small and the linearization of aerodynamics models,that are precisely tuned by using extensive wind tunnel mea-surements over a restricted set of nominal cruising velocities,then yields accurate linearized models. For VTOL vehicles,the variations of the vehicle’s attitude with respect to theairflow direction can be much more important. Nevertheless, asimilar control design approach can be adopted. This approachbasically consists in considering a set of reference trajectoriesassociated with constant translational velocities, in developingcontrol laws for each of them, and in combining these lawsvia interpolation to obtain a practical control algorithm. Tosimplify the exposition, the focus is hereafter put on the caseof quasi-stationary flight in the absence of wind, which canbe seen as an extension to the problem of stabilizing a fixedposition in space with a reference velocity equal to zero.The reviewed approaches can be extended to other trajectoriesand flight conditions, such as constant velocity cruising withconstant wind, provided that aerodynamic forces are eithermeasured or estimated on-line so that the equilibrium aboutwhich linearization is done can be calculated.

In the case of hovering and in the absence of wind, aerody-namic forces arising from relative vehicle/wind velocity can beneglected in front of the gravity force, since their intensities areproportional to the square of this velocity. As a consequence,the external forces are essentially reduced to Fe = mge3

(with g the gravity acceleration). The equilibrium orientationmatrix Re associated with hovering at a constant position ξand zero yaw angle (with ξ = 0 and ω = 0) is the identitymatrix. A first-order approximation of the matrix R about thisequilibrium is R ≈ I3 + S(Θ), with Θ ∈ R3 any minimalparametrization of SO(3) around the identity matrix (e.g., thevector of Euler angles φ: roll, θ: pitch, ψ: yaw). This yieldsthe following linear approximation of system (5)–(7) at theequilibrium (ξ = 0, ξ = 0,Θ = 0, ω = 0, T = mg,Γ = 0):

mξ = −mgS(Θ)e3 − T e3 + ΣRΓ (8)

Θ = ω (9)Iω = Γ (10)

with T = T − mg. Without loss of generality, let usassume that the vehicle’s inertia matrix is diagonal (i.e.,I = diag(I1, I2, I3)). For most VTOL vehicles and especiallyfor tail-sitters and helicopters, the coupling matrix ΣR requiresclose attention. This coupling, which induces marginally stableor unstable zero dynamics depending on the vehicle’s actuationconfiguration, has motivated various research studies [24,App.A], [39], [61], [65]. As a case study, let us considerthe class of ducted-fan tail-sitters for which the couplingmatrix ΣR is given by Eq. (4). In this case, system (8)–(10)can be decomposed into four independent Single-Input-Single-Output (SISO) linear systems. The first two systems concernthe vehicle’s altitude and Euler yaw angle dynamics. Theirequations are

mξ3 = −TI3ψ = Γ3

Page 5: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

5

The origins (ξ3, ξ3) = (0, 0) and (ψ, ω3) = (0, 0) of thesedouble integrators are exponentially stabilized by classical PD(Proportional Derivative) controllers which can be comple-mented with an integral action (yielding a PID controllers).One remarks that the third torque control variable Γ3 is onlyused for the control of the vehicle’s yaw angle, but is of nouse for the control of the vehicle’s position ξ. The other twoindependent SISO linear systems are{

mξ1 = −mgθ + Γ2/L

I2θ = Γ2(11)

{mξ2 = mgφ− Γ1/L

I1φ = Γ1(12)

By an adequate change of coordinates, each of these twosystems can be rewritten as

x1 = x2

x2 = x3 + εu

x3 = x4

x4 = u

(13a)(13b)(13c)(13d)

with (x1, x2, x3, x4) = (ξ1, ξ1,−gθ,−gθ), u = −(g/I2)Γ2,and ε = −I2/(mgL) for system (11); and (x1, x2, x3, x4) =(ξ2, ξ2, gφ, gφ), u = (g/I1)Γ1, and ε = −I1/(mgL) forsystem (12). The linear system (13a)–(13d) is controllable butthe stabilization of its origin x = 0, although obvious, deservesclose attention. A certain number of contributions on nonlinearcontrol of VTOL vehicles turn out to have a close kinship withthe linear control approaches summarized below:1) The first approach focuses on the exponential stabilizationof (x1, x2) = (0, 0) by using the control input εu when ε 6= 0.Setting v := x3 + εu, one obtains the double integrator{

x1 = x2

x2 = v(14)

whose origin is exponentially stabilized by the PD controller

v = −k0x1 − k1x2, with k0,1 > 0.

The closed-loop exponential convergence of v to zero alsofollows. Once this convergence is achieved (i.e., when x1 =x2 = 0), the dynamics of the variables x3 and x4 are givenby {

x3 = x4

x4 = −x3/ε(15)

It appears from these equations that the origin of system (15) isat best marginally stable. It is so when ε > 0, but it is unstablewhen ε < 0. Besides, even when it is marginally stable as inthe case of a helicopter for instance, it is oscillatory at a highfrequency, since ε is often very small, and thus highly sensitiveto the slightest perturbation. For these reasons this approach,which looks simple and attractive in the first place because itexploits the coupling terms associated with a non-zero matrixΣR, is not commendable.2) The second approach consists in designing a full statefeedback controller

u = −k0x1 − k1x2 − k2x3 − k3x4, with k0,1,2,3 > 0,

yielding the exponential stabilization of the origin of system(13a)–(13d). In this case, when ε is small the term εu affectsthe control performance marginally only. In fact, it sufficesto choose the control gains ki (i = 0, . . . , 4) so that thecharacteristic polynomial of the closed-loop system, given by

p4 + (k3 + εk1)p3 + (k2 + εk0)p2 + k1p+ k0,

is Hurwitz. For example, by choosing k2 and k3 large enoughwith respect to εk0, the precise knowledge of ε does not matter.Although this approach looks simple enough, it is seldom usedin practice. A probable reason is that practitioners prefer todecompose the fourth-order system (13a)–(13d) into a cascadeof subsystems of order two at most so as to facilitate controlimplementation (with on-the-shelf PID controllers) and failurediagnosis. In particular, a separation principle between thecontrol of x1 (in a so-called Guidance loop) and the controlof x3 (in a so-called Control loop) is often adopted. It ispartly justified by the different control and measurement ratesassociated with the two levels: low frequency for the Guidanceloop and high frequency for the Control loop. This yields thehierarchical control described next.3) Classical hierarchical control essentially neglects the cou-pling term εu in the translational dynamics (i.e., by settingε = 0 or equivalently ΣR = 0) and consists in considering x3

as a control input for subsystem (13a)–(13b). Then, the desiredvalue of x3, here denoted as xd3, is used as a reference valuefor subsystem (13c)–(13d). Either backstepping or a high-gaintechnique can be applied. For instance, backstepping leads torewrite the system as

x1 = x2

x2 = xd3 + x3

˙x3 = x4

˙x4 = −xd3 + u

with x3 = x3 − xd3 and x4 = x4 − xd3. With this approachxd3 represents a desired control input for the asymptoticstabilization of the origin of subsystem (13a)–(13b). Witha PD controller of the form xd3 = −k0x1 − k1x2, withk0,1 > 0, closing the loop yields xd3 = −k0x2 − k1x3 andxd3 = −k0x3−k1x4. It then remains to design a control inputu which asymptotically stabilizes the “errors” x3 and x4 atzero. One can take, for instance

u = xd3 − k2x3 − k3x4, with k2,3 > 0. (16)

Since the characteristic polynomial of the closed-loop systemis

(p2 + k1p+ k0)(p2 + k3p+ k2),

the positivity of all gains ki (i = 0, . . . , 3) indeed ensures, bya straightforward application of the Routh-Hurwitz criterion,the exponential stability of the origin of the initial fourth-ordersystem. A variation of the backstepping technique consists inusing a high-gain controller for the second subsystem (i.e.,subsystem (13c)–(13d)) while neglecting the dynamics of thereference variable xd3 (i.e., by setting xd3 = xd3 = 0 in theexpression (16) of the control input u). This yields the simpler

Page 6: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

6

controller u = −k2x3−k3x4 and the closed-loop characteristicpolynomial

p4 + k3p3 + k2p

2 + k1k2p+ k0k2

which is Hurwitz if and only if k3 > k1 and k2 >k23k0

k1(k3−k1) >0, by application of the Routh-Hurwitz criterion. The choiceof high gains k2 and k3 is consistent with these inequalitiesand is legitimate in practice when the attitude dynamics canbe rendered fast compared to the translational dynamics, andhigh-frequency measurements of the vehicle’s attitude andangular velocities are available. With recent advances onMEMS gyrometers, precise high-frequency angular velocitymeasurements from an embedded Inertial Measurement Unit(IMU) are now available at a modest cost. However, estimatingthe vehicle’s orientation is much more difficult. The useof accelerometers to recover roll and pitch angles is ofteninvoked, but the subject remains controversial [53]. Alternativeapproaches have been proposed lately based on complemen-tary measurements (issued from a GPS, for instance). Thisstill constitutes an active research topic [13], [25], [47], [52],[72]. A separation of the system’s dynamics which better takesthe above-mentioned measurement rates and limitations intoaccount consists in moving the orientation variables to the slowGuidance loops, together with the position and translationalvelocity variables. This yields the hierarchical control schemedescribed next.4) Another hierarchical control strategy consists in using x4

(i.e., the variable associated with roll and pitch angular veloc-ities) as a virtual control input for the third-order subsystem(13a)–(13c), with the coupling term εu being neglected. Thedesired value of x4, denoted as xd4, is then used as a referencefor the second subsystem which, in this case, is reducedto the first-order equation (13d). For instance, applying abackstepping technique yields the following system:

x1 = x2

x2 = x3

˙x3 = xd4 + x4

˙x4 = −xd4 + u

with x4 = x4 − xd4 and xd4 typically chosen of the form xd4 =−k0x1−k1x2−k2x3 with k0,1,2 > 0. Then, the control inputu is designed in order to exponentially stabilize x4 = 0. Forinstance, the feedback controller

u = xd4 − k3x4, with k3 > 0, (17)

yields the following closed-loop characteristic polynomial

(p3 + k2p2 + k1p+ k0)(p+ k3)

which is Hurwitz provided that all gains ki (i = 0, . . . , 3)are positive and k1k2 > k0. A simplified solution consists inusing a high-gain controller for the second subsystem (i.e.,subsystem (13d)) while neglecting the dynamics of xd4 (i.e.,by setting xd4 = 0 in the expression (17) of u). Then, theclosed-loop characteristic polynomial is

p4 + k3(p3 + k2p2 + k1p+ k0)

and it is Hurwitz if and only if k1k2 > k0 and k3 >k21

k1k2−k0 .The latter inequality is consistent with the choice of a highgain k3 for the control of the second subsystem.

Beyond classical pole placement techniques, modern LinearControl Theory is concerned with the determination of opti-mal control gains associated with optimal control problems.Typical underlying preoccupations are energy efficiency androbustness with respect to model uncertainties and/or aerody-namic disturbances. Examples of published studies along thisdirection are:• [7], where a Linear-Quadratic-Regulator (LQR) controller

is proposed for the control of the rotational dynamics ofa quadrotor helicopter and compared to a PID controller.The LQR control technique has also been applied to thestabilization of the lateral position and Euler roll angle ofa quadrotor helicopter [11] and to velocity stabilizationof a ducted-fan tail-sitter [79].

• [3], where a Linear-Quadratic-Gaussian (LQG) controller,incorporating a quadratic criterion for weighting themodel sensibility with respect to parametric variations,is applied to a model of quadrotor helicopter.

• [12], [44], [48], [66], [80], where robust H2 and H∞control techniques are applied to helicopter models.

The main limitation of linear control techniques is the localnature of the control design and analysis. More precisely,the following issues should be kept in mind when applyingthese techniques, especially for cruising flight (with non-zeroreference velocity):• In the case of reference trajectories with time-varying

velocities, or in the case of wind gusts, the equilibriumorientation varies with time. The linearized system asso-ciated with the error dynamics is then also time-varying,and ensuring the stability of the origin of this system be-comes much more involved. Such is the case, for instance,of the “gain-scheduling” approach used for helicopters(e.g., [36]) and airplanes autopilots (e.g., [30]), which isbased on interpolating local linear controllers calculatedfor a set of nominal operating conditions. By essence, thisapproach does not account for rapidly changing operatingconditions that render the stability analysis untractableeven in the ideal case where all forces are measuredexactly.

• Due to aerodynamic disturbances, the vehicle may atsome point depart from its nominal operating domain andeven get far from it. In this case, the linearized dynamicsare no longer representative of the real dynamics.

• A more purely technical issue, here mentioned essentiallyto point out useless complications, is that linearization in-volves a minimal parametrization of the vehicle’s attitude.This yields representation singularities which artificiallylimit the domain of application of the controller, unless aswitching procedure between several local minimal rep-resentations, specifically tailored to solve this problem, isadded to the controller. In this respect, some parametriza-tions (e.g., Rodrigues parametrization) are known to bebetter than others because their domain of validity islarger. For instance, despite their historical importance

Page 7: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

7

and frequent use, Euler angles are not best appropriatefor the control of small VTOL vehicles.

IV. NONLINEAR CONTROL

Compared to control design methods based on linearization,nonlinear methods can yield controllers with a significantly en-larged domain of stability and enhanced robustness, especiallyin the case of highly nonlinear dynamics, input saturations,or fast time-varying perturbations. To simplify the expositionnonlinear feedback control laws for underactuated VTOLvehicles are here classified into two families, depending onthe design approach which is considered.

A. Controllers based on dynamic extensions

The approach essentially consists in considering TRe3 as asystem state variable and in differentiating it until three inde-pendent control variables allowing for the exact linearizationof the translational dynamics are obtained. Early work [22],based on exact input-output linearization [31], has focused onthe control of a planar vertical takeoff and landing (PVTOL)aircraft. Extensions to the 3D case for a nonlinear model of aconventional helicopter are reported in [39]. To provide thereader with a more precise idea of the approach, withoutcomplicating the exposition too much, let us again assumethat the external forces acting on the vehicle are reduced to thevehicle’s weight (i.e., Fe = mge3) and that the coupling termΣRΓ in the translational dynamics can be neglected. Then, onederives the following relations from the equations of system(5)–(7):

X1 = X2/m

X2 = X3

X3 = X4

X4 = U

(18)

with

X = (X1, X2, X3, X4)>

= (ξ,mξ,−TRe3 +mge3,−Rδ)>,δ = (Tω2,−Tω1, T )>,

U = R(−T e3 + TS(e3)I−1Γ + 2TS(e3)ω − TS(ω)2e3

− TS(e3)I−1S(ω)Iω).

Provided that T 6= 0, the application (T ,Γ) −→ U issurjective, thus allowing one to view U as a new control vector.Since system (18) is linear and controllable, the asymptoticstabilization of a given reference trajectory for ξ = X1 is thena simple matter. Although this approach looks attractive at firstglance, its implementation raises a few issues:• For rotary-wing VTOL vehicles, such as ducted-fans and

quadrotors, the thrust force is generated by propeller(s)and T is a function of the associated motor’s angularvelocity, a quantity commonly used as a reference inputby motor manufacturers. If, instead of this velocity, Tis used as a control variable this means that the motor’sjerk becomes a control input which has to be monitored.This in turn induces serious complications along the

necessity to incorporate the motor’s dynamics into thecontrol design.

• In view of the expression of U , the calculation of thecontrol inputs T and Γ involves the knowledge of Tand T . For most platforms, these two variables are notavailable to measurement. Theoretically, these variablescould be estimated on-line from measurements of thevehicle’s velocity. In practice, however, the low frequencyrate of these measurements (approximately 10Hz with aGPS) does not allow for accurate estimations.

• The control is well-defined as long as T 6= 0. But thenon-crossing of zero by T , as a state variable, is notguaranteed.

B. Hierarchical controllers

Many nonlinear control laws of this type for underactuatedVTOL vehicles have been proposed. Problems addressed withthis class of methods include:• way-point navigation [62];• visual servoing [40];• control with partial state measurement [1], [6];• robustness with respect to unmodeled dynamics like, e.g.,

variations of the vehicle’s mass or gravitational field [21],[41], aerodynamic disturbances [26], [62], [63], [73],parametric uncertainties [33], [49], [50];

• robustness with respect to measurement errors [14];• actuators saturation [20], [50], [51], [74], [86].

Most design methods are based on “classical” Lyapunov-typeapproaches, but other techniques have also been used such assliding-mode control [8], [42], [85] or predictive control [5],[34], [37].

A common denominator of hierarchical nonlinear con-trollers is their focus on the determination of the thrust vectorTRe3 to control the vehicle’s translational dynamics. Theyare reminiscent of the hierarchical linear controllers describedin Section III by their two-stages architecture composed of a“low-level” fast inner-loop and a “high-level” slow outer-loop.The principles of hierarchical control for the stabilization ofreference translational velocities are presented next. Indica-tions of how these principles can be extended to other controlobjectives, such as the stabilization of reference positiontrajectories and the compensation of unmodeled dynamics andperturbations via the use of an integral action, are then given.

Let us first further simplify the control model (5)–(7)by neglecting the coupling term ΣRΓ in Eq. (5) and byusing ω as a control variable. Both simplifications are oftenmade. While the validity of the former should be assessedon a case-by-case basis, the latter corresponds to the linearhierarchical strategy considered last in the previous section,with x4 taken as a virtual control input. It is justified by thefull actuation of the vehicle’s orientation and by the high-frequency rate measurement of the angular velocity vector ωfrom an embedded set of gyrometers, which allows for the fastconvergence of ω to a desired reference value ωd. The controlmodel is then reduced to the following equations:{

mξ = −TRe3 + Fe(ξ, d(t))R = RS(ω)

(19)

Page 8: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

8

with T and ω as control inputs.1) Velocity control:

Let ξr denote a reference translational velocity expressed inthe inertial frame I, ξr its time-derivative, and ˙

ξ := ξ− ξr thevelocity error expressed in the same frame. One obtains thefollowing error model:{

m¨ξ = −TRe3 + F (ξ, t)

R = RS(ω)

(20a)

(20b)

withF (ξ, t) := Fe(ξ, d(t))−mξr(t) (21)

and with either ξ :=∫ t

0˙ξ(s)ds the integral of the velocity

error, or ξ := ξ − ξr the position tracking error when areference trajectory ξr is specified.a) Equilibrium analysis

In order to asymptotically stabilize zero for the velocityerror ˙

ξ := ξ − ξr, ˙ξ ≡ 0 must be an equilibrium of Eq. (20a).

This in turn implies that ¨ξ ≡ 0 must be satisfied and thus, in

view of Eq. (20a), that

Tη = F (ξr, t), (22)

with η := Re3 the unit vector characterizing the thrust forcedirection. The underlying geometrical interpretation is clear: atthe equilibrium the thrust direction is aligned with the directionof the external force (corrected by the reference acceleration)and the thrust intensity counterbalances the external force in-tensity. The above basic relation calls for simple but importantremarks:• As long as F (ξr, t) is different from zero, there exist only

two solutions (Tr, ηr) to Eq. (22), given by

(Tr, ηr) =

(±|F (ξr, t)|,±

F (ξr, t)|F (ξr, t)|

). (23)

In practice, the chosen solution is most of the timeimposed by physical considerations, such as the positivityof the thrust intensity.

• As already mentioned in the linear control section, moni-toring the translational motion of a flying vehicle does notimpose by itself any constraint upon the yaw angle (i.e.,the modification of which does not change the thrust di-rection). This angle can thus be controlled independentlyto achieve a complementary objective. This also means,as one could logically expect, that only three independentactuation inputs corresponding to three degrees of free-dom are needed to control the three translational velocity(or position) coordinates of a vehicle.

• When F (ξr, t) = 0, any unit vector η satisfies Eq.(22) with T = 0. This singular case corresponds tospecific operating conditions which are not commonlyencountered in practice with VTOL vehicles. This is dueto the gravity term involved in F (ξr, t) which keepsthis force away from zero when aerodynamic forces andtranslational accelerations are not too strong. However,intense wind gusts or very aggressive reference trajecto-ries may provoke this situation whose severity, in terms of

control, can be appreciated by the fact that the lineariza-tion of the error system (20) at any equilibrium point( ˙ξ,R) = (0, R∗) satisfying the equilibrium conditionF (ξr, .) ≡ 0 is not controllable, even though the systemitself remains theoretically (mathematically) controllable.Stabilization of such equilibria is a challenging problemwhose solution involves very specific nonlinear controltechniques (see, e.g., [55]), the exposition of which isbeyond the scope of the present paper.

• When F (ξr, t) 6= 0, well-posedness of the solutions(23) to Eq. (22) is much related to the dependence ofFe upon the vehicle’s orientation (i.e. upon the rotationmatrix R). More precisely, when Fe and, thus, F do notdepend on R, as in the hovering case for which Fe isessentially reduced to the vehicle’s weight, it is clear thatthe equation has a unique solution ηr once the sign ofTr has been chosen. Another favorable case is when theaerodynamic force Fe is dominated by its drag (vs. lift)component whose direction is (by definition) given bythe airflow, and is thus essentially independent of thevehicle’s orientation. Otherwise, the equation becomesimplicit in ηr and existence and uniqueness of the so-lution (given the sign of Tr) are no longer systematic.Several sub-cases can then be considered in order toattempt determining sufficient conditions under whichthese properties are satisfied. We will come back to thisissue further on.

b) Orientation controlIn a first stage, let us consider the problem of stabilizing a

given desired direction characterized by a unit vector ηd, in-dependently of other possible control objectives. This desireddirection may equally be specified by the orientation of a user’sjoystick, or be the (locally) unique solution ηr previouslyevoked for velocity control purposes. For instance, in the caseof hovering, one can take ηd = e3 and nonlinear solutionsto the control problem can be compared to the linear onespresented in the previous section. Defining ω1,2 := (ω1, ω2)>,a possible nonlinear control solution, among many otherpossibilities, is [26]

ω1,2 =(R>(k0

η × ηd(1+η>ηd)2

− S(η)2(ηd × ηd)))

1,2

(24)

with k0 > 0. This control ensures the exponential stabilityof the equilibrium point η = ηd provided that the initialconditions η(0) and ηd(0) are not opposite to each other. It isobtained by considering the storage function V := 1 − η>ηdwhose non-positive time-derivative, along any solution to thecontrolled system, satisfies

V = −k0|η × ηd|2

(1+η>ηd)2.

Two remarks are in order at this point. The first one concernsthe time-derivative ηd which is present in the right-hand sideof Eq. (24). This term can be problematic when the desireddirection ηd depends on the vehicle’s orientation R because itstime-derivative then involves ω. Relation (24) then becomesan implicit equation in ω which may not have a (bounded)solution. Note that, in the case where ηd is taken equal to ηr,

Page 9: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

9

this difficulty may arise even when the well-posedness of ηris established. In this case, another control solution must besought. The second remark concerns the calculation of ηr as areference direction. In practice, and in general, the expressionof aerodynamic forces along arbitrary reference trajectoriesis not known. As a consequence, neither F (ξr, t) nor ηr isknown. On the other hand, it is often possible to obtain anon-line estimation of F at the current point (i.e., F (ξ, t))via, for example, accelerometer measurements or high-gainobservers (see [27], [24, Ch. 2] for more details). For thisreason, control approaches based on this latter informationare often preferable. The velocity control schemes discussednext take this route, knowing that stabilization of the referencethrust direction is not by itself sufficient to ensure velocity sta-bilization and that complementary terms involving the velocityerror ˙

ξ must also be introduced in the control law.c) Classical control design

This approach (see, e.g., [62],[63]) is an extension of theprevious orientation control design to translational velocitycontrol. First, the velocity error dynamics is rewritten as(compare with Eqs. (20a)-(21))

m¨ξ = −β( ˙

ξ)− Tη + F (ξ, t), (25)

withF (ξ, t) := β( ˙

ξ) + F ( ˙ξ, t) (26)

and β(.) a function chosen so as to make ˙ξ = 0 a globally

asymptotically stable equilibrium of the equation m¨ξ =

−β( ˙ξ). A simple example is the linear function defined by

β(x) := kβx, with kβ > 0. It then only remains to ensure theasymptotic stability of F −Tη = 0 via an adequate choice ofthe thrust intensity T and of the control of the thrust directionη. Assuming that F does not vanish, a possibility consistsin defining T and a desired thrust direction ηd according to(compare with Eq. (23))

(T, ηd) =

(±|F (ξ, t)|,± F (ξ, t)

|F (ξ, t)|

)and in making η converge to ηd. To this purpose, if ηd doesnot (or little) depend(s) on the vehicle’s orientation, one canuse the orientation control ω1,2 := (ω1, ω2)> defined by Eq.(24).

Let us briefly comment upon the choice of the feedbackterm β( ˙

ξ). Beside the infinite number of possibilities for thechoice of the function β, it matters to limit the risk of avanishing F . To this end a function uniformly smaller in normthan the vehicle’s weight (corresponding to the norm of Fewhen hovering in the absence of wind) seems preferable tothe linear function used in linear controllers. One can consider,for instance

β( ˙ξ) := sat∆(Kβ

˙ξ),

with Kβ a diagonal positive gain matrix, ∆ < mg, andsat∆(.) either the classical saturation function defined bysat∆(x) := xmin(1,∆/|x|) or a differentiable approximationsuch as sat∆(x) := ∆x/(1 + |x|).d) Other control design

A nonlinear control design, recently proposed by the authors[26] and with a structure slightly different from the classi-cal design presented previously, involves the “current” forceF (ξ, t) of relation (21) whose intensity Tc and direction ηcare given by

(Tc, ηc) =

(±|F (ξ, t)|,± F (ξ, t)

|F (ξ, t)|

).

The proposed controller, whose angular velocity componentsare reminiscent of these defined by Eq. (24) with ηd replacedby ηc, is

T = Tc + Tc(ηT ηc − 1 + k2ηT ˙ξ)

= TcηT (ηc + k2

˙ξ)

ω1,2 = R>(k0

η × ηc(1+η>ηc)2

− S(η)2(ηc × ηc))

1,2

+(k1TcR

>(η × ˙ξ))

1,2

(27a)

(27b)

It is derived by considering the storage function

L :=mk1

2| ˙ξ|2 + (1− η>ηc)

whose time-derivative

L = −k0|η × ηd|2

(1+η>ηd)2− k1k2Tc(η>

˙ξ)2

along any solution to the controlled system is non-positive.It ensures the asymptotic stability of the equilibrium ( ˙

ξ, η) =(0, ηc) and the convergence to this equilibrium of all system’ssolutions with initial conditions such that η(0) and ηc(0) arenot opposite to each other.

The main difference with the classical control design is theuse of F rather than F in the definition of the “desired” thrustdirection ηc, with the advantage of not adding the contributionof the feedback term β( ˙

ξ) to the risk of encountering thesingular case associated with the vanishing of F . Being lessprone to ill-posedness, this control solution potentially appliesto a larger operating domain.e) Complementary remarks• Several studies (see, e.g., [18], [50], [51], [59, Ch. 5])

propose to define a desired vehicle’s orientation Rd ∈SO(3) instead of a desired thrust direction ηd (= Rde3).For example, Rd may be computed from ηd and anothernon-collinear direction which defines a desired value forthe yaw angle. Stabilization of R to Rd can then beachieved with any of the numerous control solutionsproposed in the literature (see, e.g., [50], [51], [33], [73],[81], [82], [84], [18], [59, Ch. 5], [64]). However, theapproach presented here shows that there is no necessityin the first place to determine the vehicle’s orientationcompletely when the objective is just to control itstranslational velocity (or its position). Controlling theyaw angle corresponds to a complementary decoupledobjective which is achieved via the determination of theremaining angular velocity component ω3, provided, ofcourse, that this degree of freedom is itself associated

Page 10: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

10

with an independent means of actuation. For instance, inthe case of helicopters and airplanes in cruising mode itusually matters to keep the slide-slip angle, i.e., the lateralangle between the airflow and the vehicle’s longitudinalaxis, small. This is achieved via the monitoring of theyaw angle (called the roll angle in the case of airplanes,because the longitudinal axis and thrust direction arequasi horizontal) either by relying on passive aerody-namic forces associated with the geometry and shape ofthe vehicle (helicopter’s tail, airplane’s dihedral angle), orby using complementary actuators (helicopter’s tail rotor,airplanes ailerons) that allow for a more precise controlof ω3 and the yaw (resp. roll) angle in a large operatingrange.

• As pointed out before, well-posedness of the nonlinearvelocity control solutions here considered partly relieson the assumption of a force F (ξ, t) or F (ξ, t) alwaysdifferent from zero. Particularly harsh unpredictable fly-ing conditions capable of inducing large tracking er-rors can clearly invalidate this property, no matter howgood the control system is. However, the fact that theseforces essentially reduce to the resultant external forceFe(ξr, d(t)) at the equilibrium (when the reference ac-celeration is small) suggests to select reference velocitiesfor which the magnitude of this force remains largerthan a certain “security” threshold. Furthermore, usingthe dissipativity (passivity) of aerodynamic forces, it hasbeen shown (see [26] for details) that properties of quasi-global stability can be obtained under the sole conditionthat F (ξr, d(t)) 6= 0 for all t.

• Another potential cause, also pointed out before, for ill-posedness of the controllers is the strong dependence ofthe external force Fe upon the vehicle’s orientation R. Inparticular, a detailed analysis of equilibrium thrust direc-tions in the case where the lift component is importanthas revealed that, given a reference translational veloc-ity, there may exist several such equilibrium directionsand, furthermore, that their number may vary with thereference velocity [68]. This in turn implies that someof these equilibria may disappear when the referencevelocity changes, in which case the asymptotic stabi-lization of such an equilibrium is clearly meaningless.This difficulty, intimately related to the appearance of theso-called “stall” phenomenon, calls for complementaryinvestigations.

• The control expressions (24) and (27b) involve the time-derivative of the desired (resp. current) thrust direction ηd(resp. ηc). Since these directions themselves depend onthe vector of external forces Fe, an estimation of the time-derivative of this vector is needed to calculate the controlinputs. Obtaining a good estimation of this derivativefor arbitrary flight conditions is not easy. However, forfavorable flight conditions, such as hovering or constantvelocity cruising with constant wind, this derivative canbe taken equal to zero without much affecting the controlperformance. More details on the issue can be found in[4] in the particular hovering case. This simplification isalso commonly made in practice for way-point navigation

and visual servo-control.• In relation to the previous remark, not only can Fe

vary quickly in some cases, because of either rapidlychanging environmental conditions or important refer-ence accelerations required by the execution of extrememaneuvers, the possible dependence of this force uponthe vehicle’s orientation makes the time-derivative ofthis force a function depending linearly on the angularvelocity ω, thus rendering the control expressions (24)and (27b) implicit in the control inputs ω1,2. As longas the corresponding “gains” are small, this dependencecan be neglected in the calculation of the control inputs.Otherwise, it has to be evaluated via a model of Fe andeventually used in a modified expression of the control.This issue particularly concerns vehicles with large wingsurfaces that produce strong lift forces depending on theirorientation with respect to the airflow. Exploring it furtheris of utmost importance to extend and generalize thenonlinear control methods presented here.

Basic extensions of the previous nonlinear velocity con-trollers to the more demanding control objectives of velocitycontrol with integral correction and of position control arepresented next. For the sake of brevity these extensions areonly specified for the classical design version, but they applysimilarly to the other design (see [26] for details).

2) Velocity control with integral correction:Robustness against model uncertainties is discussed next. Thecontrol laws presented so far involve the knowledge of theexternal force vector Fe. In practice, whatever the method usedfor estimating Fe (e.g., functional modeling, on-line estimationfrom embarked sensors data, etc.) the knowledge of this term isnever totally accurate. To counterbalance the negative effectsresulting from imprecise knowledge of this force and fromother unmodeled terms involved in the system’s dynamics, anintegral action can be incorporated into the control law. Forexample, as suggested in [62], [63], [73], one can modify thefeedback term β in Eq. (25) so as to introduce an integralcorrection term, i.e.,

β(∫ ˙ξ,

˙ξ) = Kβ2

∫˙ξ + sat∆(Kβ1

˙ξ), (28)

with ∆ a positive number and (Kβ1 ,Kβ2) a pair of positivediagonal gain matrices. Note, in view of Eq. (26), that theexpression of F is then also modified.

It is common knowledge that integral correction action (i.e.,the term

∫ ˙ξ) constitutes an effective means to compensate

for model uncertainties. But it is also well-known that thistype of correction may generate instability and wind-up prob-lems (like, e.g., large overshoot, slow desaturation, actuatorsaturation, etc.). For instance, even when ˙

ξ stays close tozero its integral can grow arbitrarily large. As a consequence,there is an increased risk for F to cross zero and for thecontrol expression to become singular. On the other hand, theuse of a saturation function sat∆I

(∫ ˙ξ) in place of

∫ ˙ξ in the

expression of β(·) does not prevent this integral term fromgrowing arbitrarily large, leading to slow desaturation andsluggish dynamics. A possible remedy consists in using an

Page 11: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

11

anti-windup integrator technique (see, e.g., [26], [28], [76] foranti-windup solutions). For instance, one may replace the term∫ ˙ξ in Eq. (28) by the “bounded” integral term z calculated as

follows (see, e.g., [76])

z = kz(−z + sat∆z (z + ˙ξ)), |z(0)| < ∆z ,

with kz > 0 and ∆z > 0. Using the classical saturationfunction, this relation yields a pure integrator z = kz

˙ξ when

|z + ˙ξ| ≤ ∆z , and also |z(t)| ≤ ∆z and |z(t)| ≤ 2kz∆z ,

∀t ≥ 0.3) Position control:

As in the case of velocity control with integral correction,position trajectory tracking can be achieved via a suitablemodification of the feedback term β. A first solution is

β(ξ, ˙ξ) = sat∆2(Kβ2 ξ) + sat∆1(Kβ1

˙ξ),

with (∆1,∆2) a pair of positive numbers and (Kβ1 ,Kβ2) apair of positive diagonal gain matrices. The position feedbackterm sat∆2(Kβ2 ξ) provides with the correction necessary tothe convergence of ξ to zero, while saturation functions areagain used to reduce the risk of a vanishing F .

To further add robustness with respect to unmodeled dynam-ics, a bounded nonlinear integrator with anti-windup propertiescan be also incorporated into the controller. The solutionproposed in [26] involves new tracking error variables defined

by ¯ξ := ξ + z,

˙ξ := ˙

ξ + z, with z a “bounded integral”of ξ, twice differentiable and with a second time-derivativeuniformly bounded in norm by an arbitrary number specifiedby the user. Examples of calculation of such a term are given in[26] and [28]. In the latter reference z is calculated as follows

z = −kvz z + satzmax/2(kpz(−z + sat∆z(z + ξ))),

z(0) = 0, z(0) = 0,(29)

with the classical saturation function, and kvz, kpz,∆z, zmaxdenoting positive numbers. This relation yields the followingupper-bounds: |z| < ∆z + zmax/(2k2

vz), |z| < zmax/(2kvz),

|z| < zmax. Using the definitions of ¯ξ and

˙ξ in the equations

of system (20) yieldsm ¨ξ = −TRe3 + Fz(ξ, z, t)

R = RS(ω)

(30a)

(30b)

with Fz(ξ, z, t) := F (ξ, t)+mz. Since Eqs. (30a)–(30b) are ofthe same form as those of the original system (20), one can usethe position controller evoked previously, modulo the change

of (ξ, ˙ξ, F ) to ( ¯

ξ,˙ξ, Fz), to obtain the asymptotic stability of

( ¯ξ,

˙ξ) = (0, 0). If z were the exact (unbounded) integral of ξ,

the convergence of ¯ξ to zero would clearly imply the one of

ξ to zero. It is not difficult to show that the same conclusionholds when z is calculated according to Eq. (29).

V. CONCLUSION

Basic control design ideas and principles for VTOL vehicleshave been reviewed. One of the messages that we have tried topass on is that the nonlinear approach to the control problems

has definite assets with respect to the well established anduniversally used linearization approach. It is founded on thephysics of flight, it is respectful of the geometry of motion inspace, and it allows for a large operating domain. The controlschemes derived from it are not more complex and may in factbe seen as natural extensions of locally approximating linearcontrol schemes. The interest of this latter point of view is thatit reconciles the two approaches by also pointing out that linearcontrol gain optimization techniques are useful for the tuningof a certain number of parameters involved in nonlinear controlschemes. The paper is an introduction to control principles onwhich it is possible to elaborate fully operational automaticcontrol flight systems. As such it does not pretend to coverall aspects of flight control. Nor does it address all VTOLvehicles. In this respect, the assumption of independence ofaerodynamic forces with respect to the vehicle’s orientationis a severe limitation. Many problems, a certain number ofwhich have been pointed out along the paper, are still unsolvedand directions for further exploration are numerous (sensoryfusion for state estimation, coupling of state estimation withcontrol, vision-based control, to name a few). Let us just againmention here the importance of taking aerodynamic forces intoaccount in the control design if the objective is to fly smalldevices in adverse wind conditions, and/or when lift becomesthe dominant aerodynamic force component. Convertible aerialvehicles capable of hovering and of economy-cruising by usinglift are particularly concerned by this issue and the adjacentproblem of transition between these two modes [69], [70],[57]. The spectrum of vehicles of this type, from blimps toflapping-wing devices [2], [71], is very large. Beyond thecase of flying vehicles, many other “thrust-propelled” vehicles,such as marine ships, hovercrafts, and submarines, can alsobe addressed with similar ideas and techniques by properlytaking the specific nature and properties of external forcesinto account. All works in this direction participate in thedevelopment of a unified control theory for the ever growingfamily of underactuated vehicles.

ACKNOWLEDGMENTS

This work has been supported by the “Chaire d’excellenceen robotique RTE-UPMC”.

REFERENCES

[1] A. Abdessameud and A. Tayebi. Global trajectory tracking controlof VTOL-UAVs without linear velocity measurements. Automatica,46(6):1053–1059, 2010.

[2] J. R. Azinheira, A. Moutinho, and E. C. De Paiva. Airship hoverstabilization using a backstepping control approach. AIAA Journal ofGuidance, Control, and Dynamics, 29(4):903–914, 2006.

[3] A. Benallegue, A. Belaidi, and A. Mokhtari. Polynomial linear quadraticgaussian and sliding mode observer for a quadrotor unmanned aerialvehicle. Journal of Robotics and Mechatronics, 17(3):483–495, 2006.

[4] S. Bertrand, N. Guenard, T. Hamel, H. Piet-Lahanier, and L. Eck. Ahierarchical controller for miniature VTOL UAVs: Design and stabilityanalysis using singular perturbation theory. Control Engineering Prac-tice, pages 1099–1108, 2011.

[5] S. Bertrand, T. Hamel, and H Piet-Lahanier. Performance Improvementof an Adaptive Controller using Model Predictive Control : Applicationto an UAV Model. In 4th IFAC Symposium on Mechatronic Systems,pages 770–775, 2006.

Page 12: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

12

[6] S. Bertrand, T. Hamel, and H. Piet-Lahanier. Stabilization of a SmallUnmanned Aerial Vehicle Model without Velocity Measurement. InIEEE International Conference on Robotics and Automation, pages 724–729, 2007.

[7] S. Bouabdallah, A. Noth, and R. Siegwart. PID vs LQ ControlTechniques Applied to an Indoor Micro Quadrotor. In Intelligent Robotsand Systems, pages 2451–2456, 2004.

[8] S. Bouabdallah and R. Siegwart. Backstepping and Sliding-mode Tech-niques Applied to an Indoor Micro Quadrotor. In IEEE InternationalConference on Robotics and Automation, pages 2247–2252, 2005.

[9] P.-J. Bristeau, F. Callou, D. Vissire, and N. Petit. The Navigation andControl technology inside the AR.Drone micro UAV. In IFAC WorldCongress, pages 1477–1484, 2011.

[10] P.J. Bristeau, P. Martin, and E. Salaun. The role of propeller aerodynam-ics in the model of a quadrotor UAV. In European Control Conference,pages 683–688, 2009.

[11] P. Castillo, R. Lozano, and A. Dzul. Stabilization of a Mini Rotorcraftwith Four Rotors. IEEE Control Systems Magazine, pages 45–55, 2005.

[12] M. L. Civita, G. Papageorgiou, W. C. Messner, and T. Kanade. Designand flight testing of a gain-scheduled H∞ loop shaping controllerfor wide-envelope flight of a robotic helicopter. In American ControlConference, pages 4195–4200, 2003.

[13] J.L. Crassidis, F.L. Markley, and Y. Cheng. Survey of nonlinear attitudeestimation methods. AIAA Journal of Guidance, Control, and Dynamics,30(1):12–28, 2007.

[14] H. de Plinval, P. Morin, and P. Mouyon. Nonlinear control of under-actuated vehicles with uncertain position measurements and applicationto visual servoing. In American Control Conference, pages 3253–3259,2012.

[15] A. Dzul, T. Hamel, and R. Lozano. Modeling and nonlinear control fora coaxial helicopter. In IEEE International Conference on Systems, Manand Cybernetics, volume 6, 2002.

[16] P. Fabiani, V. Fuertes, A. Piquereau, R. Mampey, and F. Teichteil-Knigsbuch. Autonomous flight and navigation of VTOL UAVs: fromautonomy demonstrations to out-of-sight flights. Aerospace Science andTechnology, 11:183–193, 2007.

[17] T. I. Fossen. Guidance and control of ocean vehicles. John Wiley andSons, 1994.

[18] E. Frazzoli, M. A. Dahleh, and E. Feron. Trajectory tracking controldesign for autonomous helicopters using a backstepping algorithm. InAmerican Control Conference, pages 4102–4107, 2000.

[19] N. Guenard, T. Hamel, and R. Mahony. A Practical Visual Servo Controlfor an Unmanned Aerial Vehicle. IEEE Transactions on Robotics,24(2):331–340, 2008.

[20] J.F. Guerrero-Castellanos, N. Marchand, A. Hably, S. Lesecq, andJ. Delamare. Bounded attitude control of rigid bodies: Real timeexperimentation to a quadrotor mini-helicopter. Control EngineeringPractice, 19:790–797, 2011.

[21] T. Hamel, R. Mahony, R. Lozano, and J. Ostrowski. Dynamic modellingand configuration stabilization for an X4-flyer. In IFAC World Congress,pages 200–212, 2002.

[22] J. Hauser, S. Sastry, and G. Meyer. Nonlinear control design for slightlynon-minimum phase systems: Application to V/STOL. Automatica,28:651–670, 1992.

[23] G. M. Hoffmann, H. Huang, S. L. Waslander, and C. J. Tomlin. Quadro-tor Helicopter Flight Dynamics and Control: Theory and Experiment.In AIAA Guidance, Navigation and Control Conference and Exhibit,number 2007-6461, 2007.

[24] M.-D. Hua. Contributions to the Automatic Control of Aerial Vehicles.PhD thesis, University of Nice-Sophia Antipolis, 2009.

[25] M.-D. Hua. Attitude estimation for accelerated vehicles using GPS/INSmeasurements. Control Engineering Practice, pages 723–732, 2010.

[26] M.-D. Hua, T. Hamel, P. Morin, and C. Samson. A control approachfor thrust-propelled underactuated vehicles and its application to VTOLdrones. IEEE Transactions on Automatic Control, 54(8):1837–1853,2009.

[27] M.-D. Hua, P. Morin, and C. Samson. Balanced-force-control for un-deractuated thrust-propelled vehicles. In IEEE Conference on Decisionand Control, pages 6435–6441, 2007.

[28] M.-D. Hua and C. Samson. Time sub-optimal nonlinear PI and PIDcontrollers applied to Longitudinal Headway Car Control. InternationalJournal of Control, (10):1717–1728, 2011.

[29] H. Huang, G. M. Hoffmann, S. L. Waslander, and C. J. Tomlin.Aerodynamics and Control of Autonomous Quadrotor Helicopters inAggressive Maneuvering. In IEEE International Conference on Roboticsand Automation, pages 3277–3282, 2009.

[30] R. A. Hyde and K. Glover. The Application of Scheduled Controllersto a VSTOL Aircraft. IEEE Transactions on Automatic Control,38(7):1021–1039, 1993.

[31] A. Isidori. Nonlinear Control Systems. Springer Verlag, third edition,1995.

[32] A. Isidori, L. Marconi, and A. Serrani. Robust autonomous guidance :an internal model approach. Advances in industrial control. SpringerVerlag, 2003.

[33] A. Isidori, L. Marconi, and A. Serrani. Robust Nonlinear Motion Controlof a Helicopter. IEEE Transactions on Automatic Control, 48(3):413–426, 2003.

[34] A. Jadbabaie, J. Yu, and J. Hauser. Receding horizon control of theCaltech ducted fan : A control Lyapunov function approach. In IEEEConference on Control Applications, pages 51–56, 1999.

[35] E. N. Johnson and M. A. Turbe. Modeling, Control, and Flight Testingof a Small Ducted Fan Aircraft. AIAA Journal of Guidance, Control,and Dynamics, 29(4):769–779, 2006.

[36] B. Kadmiry and D. Driankov. A fuzzy gain-scheduler for the attitudecontrol of an unmanned helicopter. IEEE Transactions on Fuzzy Systems,12(4):502–515, 2004.

[37] H. J. Kim, D. H. Shim, and S. Sastry. Nonlinear Model PredictiveTracking Control for Rotorcraft-based Unmanned Aerial Vehicles. InAmerican Control Conference, pages 3576–3581, 2002.

[38] A. Ko, O. J. Ohanian, and P. Gelhausen. Ducted fan UAV modeling andsimulation in preliminary design. In AIAA Modeling and SimulationTechnologies Conference and Exhibit, number 2007–6375, 2007.

[39] T. J. Koo and S. Sastry. Output tracking control design for a helicoptermodel based on approximate linearization. In IEEE Conference onDecision and Control, pages 3635–3640, 1998.

[40] F. Le-Bras, R. Mahony, T. Hamel, and P. Binetti. Adaptive filtering andimage based visual servo control of a ducted fan flying robot. In IEEEConference on Decision and Control, pages 1751–1757, 2006.

[41] D. Lee, T. C. Burg, D. M. Dawson, D. Shu, B. Xian, and E. Tatlicioglu.Robust Tracking Control of an Underactuated Quadrotor Aerial-RobotBased on a Parametric Uncertain Model. In Conference on System, Man,and Cybernetics, pages 3187–3192, 2009.

[42] D. Lee, H. J. Kim, and S. Sastry. Feedback Linearization vs. AdaptiveSliding Mode Control for a Quadrotor Helicopter. International Journalof Control, Automation, and Systems, 7(3):419–428, 2009.

[43] L. Lipera, J. Colbourne, M. Tischler, M. Mansur, M. Rotkowitz, andP. Patangui. The micro craft istar micro-air vehicle: Control systemdesign and testing. In Annual Forum of the American Helicopter Society,pages 1–11, 2001.

[44] C.-C. Luo, R.-F. Liu, C.-D. Yang, and Y.-H. Chang. Helicopter H∞control design with robust flying quality. Aerospace Science andTechnology, 7(2):159–169, 2003.

[45] S. Lupashin, A. Schoellig, M. Sherback, and R. D’Andrea. A simplelearning strategy for high-speed quadrocopter multi-flips. In IEEEInternational Conference on Robotics and Automation, pages 642–1648,2010.

[46] R. Mahony, T. Hamel, and A. Dzul. Hover control via an approximateLyapunov control for a model helicopter. In IEEE Conference onDecision and Control, pages 3490–3495, 1999.

[47] R. Mahony, T. Hamel, and J.-M. Pflimlin. Nonlinear complementaryfilters on the special orthogonal group. IEEE Transactions on AutomaticControl, 53(5):1203–1218, 2008.

[48] S. Mammar and G. Duc. Loop shaping H∞ design applied to therobust stabilization of an helicopter. In IEEE Conference on ControlApplications, pages 806–811, 1992.

[49] L. Marconi, A. Isidori, and A. Serrani. Autonomous vertical landing onan oscillating platform: an internal-model based approach. Automatica,38:21–32, 2002.

[50] L. Marconi and R. Naldi. Robust full degree-of-freedom tracking controlof a helicopter. Automatica, 43:1909–1920, 2007.

[51] L. Marconi and R. Naldi. Aggressive control of helicopters in presenceof parametric and dynamical uncertainties. Mechatronics, 18:381–389,2008.

[52] P. Martin and E. Salaun. An invariant observer for Earth-Velocity-Aidedattitude heading reference systems. In IFAC World Congress, pages9857–9864, 2008.

[53] P. Martin and E. Salaun. The true role of accelerometer feedback inquadrotor control. In IEEE International Conference on Robotics andAutomation, pages 1623–1629, 2010.

[54] D. Mellinger, N. Michael, and V. Kumar. Trajectory generation and con-trol for precise aggressive maneuvers with quadrotors. The InternationalJournal of Robotics Research, pages 1–11, 2012.

Page 13: Introduction to Feedback Control of Underactuated VTOL ... · Introduction to Feedback Control of Underactuated VTOL Vehicles Minh-Duc Hua, Tarek Hamel (Member, IEEE), Pascal Morin,

13

[55] P. Morin and C. Samson. Control with transverse functions and a singlegenerator of underactuated mechanical systems. In IEEE Conference onDecision and Control, pages 6110–6115, 2006.

[56] R. Naldi. Prototyping, Modeling and Control of a Class of VTOL AerialRobots. PhD thesis, University of Bologna, 2008.

[57] R. Naldi and L. Marconi. Optimal transition maneuvers for a class ofV/STOL aircraft. Automatica, 47:870–879, 2011.

[58] R. Naldi, L. Marconi, and A. Sala. Modelling and Control of a MiniatureDucted-Fan in Fast Forward Flight. In American Control Conference,pages 2552–2557, 2008.

[59] R. Olfati-Saber. Nonlinear Control of Underactuated MechanicalSystems with Application to Robotics and Aerospace vehicles. PhDthesis, Massachusetts Institute of Technology, 2001.

[60] J.-M. Pflimlin. Commande d’un minidrone a helice carenee: de lastabilisation dans le vent a la navigation autonome (in French). PhDthesis, Ecole Doctorale Systemes de Toulouse, 2006.

[61] J.-M. Pflimlin, P. Binetti, P. Soueres, T. Hamel, and D. Trouchet.Modeling and attitude control analysis of a ducted-fan micro aerialvehicle. Control Engineering Practice, pages 209–218, 2010.

[62] J.-M. Pflimlin, T. Hamel, P. Soueres, and R. Mahony. A HierarchicalControl Strategy for the Autonomous Navigation of a Ducted Fan FlyingRobot. In IEEE International Conference on Robotics and Automation,pages 2491–2496, 2006.

[63] J.-M. Pflimlin, P. Soueres, and T. Hamel. Hovering flight stabilizationin wind gusts for ducted fan UAV. In IEEE Conference on Decisionand Control, pages 3491–3496, 2004.

[64] P. Pounds, T. Hamel, and R. Mahony. Attitude control of rigid bodydynamics from biased IMU measurements. In IEEE Conference onDecision and Control, pages 4620–4625, 2007.

[65] P. Pounds, R. Mahony, and P. Corke. Modelling and control of a largequadrotor robot. Control Engineering Practice, pages 691–699, 2010.

[66] E. Prempain and I. Postlethwaite. Static H∞ loop shaping control of afly-by-wire helicopter. Automatica, 41:1517–1528, 2005.

[67] R. W. Prouty. Helicopter Performance, Stability, and Control. Krieger,1986.

[68] D. Pucci. Flight dynamics and control in relation to stall. In AmericanControl Conference, 2012.

[69] D. Pucci, T. Hamel, P. Morin, and C. Samson. Nonlinear Control ofPVTOL Vehicles subjected to Drag and Lift. In IEEE Conference onDecision and Control, pages 6177–6183, 2011.

[70] D. Pucci, T. Hamel, P. Morin, and C. Samson. Modeling and Controlof Bisymmetric Aerial Vehicles Subjected to Drag and Lift. TechnicalReport hal-00685827, INRIA, 2012.

[71] H. Rifai, N. Marchand, and G. Poulin. Path Tracking Control of aFlapping Micro Aerial Vehicle (MAV). In IFAC World Congress, pages5359–5364, 2008.

[72] A. Robert and A. Tayebi. On the Attitude Estimation of AcceleratingRigid-Bodies Using GPS and IMU Measurements. In IEEE Conferenceon Decision and Control, pages 8088–8093, 2011.

[73] A. Roberts and A. Tayebi. Adaptive position tracking of vtol uavs. IEEETransactions on Robotics, 27(1):129–142, 2011.

[74] A. Sanchez, P. Garcia, P. Castillo, and R. Lozano. Simple Real-timestabilization of a VTOL aircraft with Bounded Signals. AIAA Journalof Guidance, Control, and Dynamics, 31(4):1166–1176, 2008.

[75] A. Schoellig, F. Augugliaro, S. Lupashin, and R. D’Andrea. Synchro-nizing the motion of a quadrocopter to music. In IEEE InternationalConference on Robotics and Automation, pages 3355–3360, 2010.

[76] S. Seshagiri and H. Khalil. Robust output feed back regulation ofminimum-phase nonlinear systems using conditional integrators. Au-tomatica, 41:43–54, 2005.

[77] O. Shakernia, Y. Ma, T. Koo, and S. Sastry. Landing an Unmanned AirVehicle: Vision based motion estimation and nonlinear control. AsianJournal of Control, 1(3):128–145, 1999.

[78] B. L. Stevens and F. L. Lewis. Aircraft control and simulation. JohnWiley and Sons, 1992.

[79] R. H. Stone. Control architecture for a Tail-Sitter Unmanned Air Vehicle.In 5th Asian Control Conference, pages 736–744, 2004.

[80] M. D. Takahashi. Systhesis and evaluation of an H2 control law for ahovering helicopter. AIAA Journal of Guidance, Control, and Dynamics,16(3):579–584, 1993.

[81] A. Tayebi. Unit quaternion-based output feedback for the attitudetracking problem. IEEE Transactions on Automatic Control, 53(6):1516–1520, 2008.

[82] A. Tayebi and S. McGilvray. Attitude Stabilization of a VTOL QuadrotorAircraft. IEEE Transactions on Control Systems Technology, 14(3):562–571, 2006.

[83] J. C. A. Vilchis, B. Brogliato, A. Dzul, and R. Lozano. Nonlinearmodelling and control of helicopters. Automatica, 39:1583–1596, 2003.

[84] J. T.-Y. Wen and K. Kreutz-Delgado. The attitude control problem. IEEETransactions on Automatic Control, 36(10):1148–1162, 1991.

[85] R. Xu and U. Ozguner. Sliding mode control of a class of underactuatedsystems. Automatica, 44:233–241, 2008.

[86] A. Zavala-Rio, I. Fantoni, and R. Lozano. Global stabilization of aPVTOL aircraft model with bounded inputs. International Journal ofControl, 76(18):1833–1844, 2003.

Minh-Duc Hua graduated from Ecole Polytech-nique, France, in 2006, and received his Ph.D. fromthe University of Nice–Sophia Antipolis, France,in 2009. He spent two years as a postdoctoralresearcher at I3S UNS-CNRS, France. He is cur-rently Charge de Recherche of the Centre Nationalde la Recherche Scientifique (CNRS) at the ISIRlaboratory of the University Pierre et Marie Curie(UPMC), France. He is also Affiliate Scientist ofthe John Von Neumann Institute, Vietnam NationalUniversity HCM City. His research interests include

nonlinear control theory, estimation and teleoperation with applications toautonomous mobile robots such as UAVs.

Tarek Hamel is Professor at the University ofNice Sophia Antipolis since 2003. He conductedhis Ph.D. research at the University of Technolo-gie of Compiegne (UTC), France, and receivedhis doctorate degree in Robotics from the UTCin 1996. After two years as a research assistantat the (UTC), he joined the Centre d’Etudes deMcanique d’Iles de France in 1997 as an associateprofessor. His research interests include nonlinearcontrol theory, estimation and vision-based controlwith applications to Unmanned Aerial Vehicles. He

is currently Associate Editor for IEEE Transactions on Robotics and forControl Engineering Practice.

Pascal Morin received the Maıtrise degree fromUniversite Paris-Dauphine in 1990, and the Diplomed’Ingenieur and Ph.D. degrees from Ecole des Minesde Paris in 1992 and 1996 respectively. He spentone year as a post-doctoral fellow in the Control andDynamical Systems Department at the California In-stitute of Technology. He was Charge de Rechercheat INRIA, France, from 1997 to 2011. He is cur-rently in charge of the “Chaire RTE-UPMC Mini-drones autonomes” at the ISIR lab of UniversityPierre et Marie Curie (UPMC) in Paris. His research

interests include stabilization and estimation problems for nonlinear systems,and applications to mechanical systems such as nonholonomic vehicles orUAVs.

Claude Samson graduated from the EcoleSuperieure d’Electricite in 1977, and receivedhis Docteur-Ingenieur and Docteur d’Etat degreesfrom the University of Rennes, in 1980 and 1983,respectively. He joined INRIA in 1981, where heis presently Directeur de Recherche. His researchinterests are in control theory and its applicationsto the control of mechanical systems. Dr. Samsonis the coauthor, with M. Leborgne and B. Espiau,of the book Robot Control. The Task-FunctionApproach (Oxford University Press, 1991).