12
IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated Motion Design on Lie Groups Alain Sarlette, Member, IEEE, Silvère Bonnabel, and Rodolphe Sepulchre, Senior Member, IEEE Abstract—The present paper proposes a unified geometric framework for coordinated motion on Lie groups. It first gives a general problem formulation and analyzes ensuing conditions for coordinated motion. Then, it introduces a precise method to design control laws in fully actuated and underactuated settings with simple integrator dynamics. It thereby shows that coordination can be studied in a systematic way once the Lie group geometry of the configuration space is well characterized. Applying the proposed general methodology to particular examples allows to retrieve control laws that have been proposed in the literature on intuitive grounds. A link with Brockett’s double bracket flows is also made. The concepts are illustrated on , and . Index Terms—Cooperative systems, distributed control, motion planning, Lie groups, geometric control. I. INTRODUCTION R ECENTLY, many efforts have been devoted to the design and analysis of control laws that coordinate swarms of identical autonomous agents—e.g., oscillator synchronization [1], [2], flocking mechanisms [3], [4], vehicle formations [5]–[9], spacecraft formations [10]–[15], mechanical system networks [16]–[18] and mobile sensor networks [19]–[23]. For systems on vector spaces, so-called consensus algorithms are shown to be efficient and robust [3], [24]–[28], and allow to address many relevant engineering issues and tasks [5], [24], [29]. However, in many applications, the agents to coordinate evolve on nonlinear manifolds: oscillators evolve on the circle , satellite attitudes on and vehicles move in or ; these particular manifolds share the geometric structure of a Lie group. Coordination on nonlinear manifolds is inherently more difficult than on vector spaces [46]. The goal of the present paper is to propose a unified geometric framework for coordinated motion on Lie groups, from a geometric definition of “coordination” to a geometric derivation of control laws for coordination like those proposed in [19]–[21], [30]–[32], in fully actuated and underactuated settings with simple integrator dynamics. The objective is to reach a state where the motion of the agents is coordinated, Manuscript received July 28, 2008; revised April 01, 2009. First published February 02, 2010; current version published May 12, 2010. This work was sup- ported by the Belgian Network Dynamical Systems, Control, and Optimization (DYSCO), funded by the Interuniversity Attraction Poles Programme, initiated by the Belgian State, Science Policy Office, and A. Sarlette is supported as an FNRS Fellow (Belgian Fund for Scientific Research). Recommended by Asso- ciate Editor Z. Qu. A. Sarlette and R. Sepulchre are with the Department of Electrical En- gineering and Computer Science, Université de Liège, Belgium (e-mail: [email protected]; [email protected]). S. Bonnabel was with Department of Electrical Engineering and Computer Science, Université de Liège, Belgium. He is now with Mines ParisTech, Paris, France (e-mail: [email protected]). Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identifier 10.1109/TAC.2010.2042003 while the values of their relative positions are a priori left arbitrary; definitions of “coordinated motion” and “relative positions” on a Lie group are the subject of Section II. Symmetries: The key point for the developments in this paper is invariance (or symmetry) in the behavior of the swarm of agents with respect to their absolute position on the Lie group: only relative positions (on the Lie group) matter. For instance, the configuration of a rigid body in the 3-D physical world is given by an orientation and a position vector in , whose com- bination corresponds to a position on Lie group . For rigid body coordination, it is then natural to write control laws that can be interpreted as internal forces in the swarm, rather than forces depending on an external reference frame which would privilegiate some arbitrary choice of orientation and origin. In- dependence with respect to reference frame corresponds to in- variance with respect to applying to all agents the same Lie group translation on . The symmetries determine how to define meaningful quanti- ties for the swarm, like “relative positions” on the Lie group, and what the dynamics of the coupled agents can be. Coordinated mo- tion—in short coordination—is defined as all situations where relative positions on the Lie group are fixed. Feedback control laws that asymptotically enforce coordination must be designed on the basis of error measurements involving appropriately in- variant quantities (e.g., relative agent positions on the Lie group). Previous Work: Results about synchronization (“reaching a common point”) and coordinated motion (“moving in an or- ganized way”) on vector spaces are becoming well established [24], [26]–[28]. Because a vector space can be identified with its tangent plane, both synchronization and coordinated motion can be seen as consensus problems on the same vector space: the former is a position consensus while the latter is a velocity con- sensus. Note that considering the motion of agents with the Lie group structure of implies that only position vectors in and associated translational motion are covered. In contrast, as soon as orientation/rotation of the vehicles or of the formation moving in a vector space is considered, the configuration space becomes the non-trivial Lie group . In general, when the configuration space is a Lie group, synchronization and coordi- nated motion are fundamentally different. The geometric view- point for dynamical systems on Lie groups is very well studied; see basic results in [33], [34] for simplified dynamics like those considered in the present paper, and [34]–[37] for a geometric theory of mechanical systems on Lie groups. General results for synchronization on compact Lie groups are proposed in [38], which points to related examples in the literature. But to the best of the authors’ knowledge, a unified geometric viewpoint for co- ordinated motion—in short coordination—on Lie groups is still lacking. Close to the present paper in its geometric flavor, [39] builds invariant observers for systems with Lie group symme- tries; observer design can be seen as two-agent leader-follower synchronization. 0018-9286/$26.00 © 2010 IEEE Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047

Coordinated Motion Design on Lie GroupsAlain Sarlette, Member, IEEE, Silvère Bonnabel, and Rodolphe Sepulchre, Senior Member, IEEE

Abstract—The present paper proposes a unified geometricframework for coordinated motion on Lie groups. It first gives ageneral problem formulation and analyzes ensuing conditions forcoordinated motion. Then, it introduces a precise method to designcontrol laws in fully actuated and underactuated settings withsimple integrator dynamics. It thereby shows that coordinationcan be studied in a systematic way once the Lie group geometryof the configuration space is well characterized. Applying theproposed general methodology to particular examples allows toretrieve control laws that have been proposed in the literature onintuitive grounds. A link with Brockett’s double bracket flows isalso made. The concepts are illustrated on ���, ��� and

���.

Index Terms—Cooperative systems, distributed control, motionplanning, Lie groups, geometric control.

I. INTRODUCTION

R ECENTLY, many efforts have been devoted to the designand analysis of control laws that coordinate swarms of

identical autonomous agents—e.g., oscillator synchronization[1], [2], flocking mechanisms [3], [4], vehicle formations[5]–[9], spacecraft formations [10]–[15], mechanical systemnetworks [16]–[18] and mobile sensor networks [19]–[23]. Forsystems on vector spaces, so-called consensus algorithms areshown to be efficient and robust [3], [24]–[28], and allow toaddress many relevant engineering issues and tasks [5], [24],[29]. However, in many applications, the agents to coordinateevolve on nonlinear manifolds: oscillators evolve on the circle

, satellite attitudes on and vehicles movein or ; these particular manifolds share thegeometric structure of a Lie group. Coordination on nonlinearmanifolds is inherently more difficult than on vector spaces[46]. The goal of the present paper is to propose a unifiedgeometric framework for coordinated motion on Lie groups,from a geometric definition of “coordination” to a geometricderivation of control laws for coordination like those proposedin [19]–[21], [30]–[32], in fully actuated and underactuatedsettings with simple integrator dynamics. The objective is toreach a state where the motion of the agents is coordinated,

Manuscript received July 28, 2008; revised April 01, 2009. First publishedFebruary 02, 2010; current version published May 12, 2010. This work was sup-ported by the Belgian Network Dynamical Systems, Control, and Optimization(DYSCO), funded by the Interuniversity Attraction Poles Programme, initiatedby the Belgian State, Science Policy Office, and A. Sarlette is supported as anFNRS Fellow (Belgian Fund for Scientific Research). Recommended by Asso-ciate Editor Z. Qu.

A. Sarlette and R. Sepulchre are with the Department of Electrical En-gineering and Computer Science, Université de Liège, Belgium (e-mail:[email protected]; [email protected]).

S. Bonnabel was with Department of Electrical Engineering and ComputerScience, Université de Liège, Belgium. He is now with Mines ParisTech, Paris,France (e-mail: [email protected]).

Color versions of one or more of the figures in this paper are available onlineat http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/TAC.2010.2042003

while the values of their relative positions are a priori leftarbitrary; definitions of “coordinated motion” and “relativepositions” on a Lie group are the subject of Section II.

Symmetries: The key point for the developments in thispaper is invariance (or symmetry) in the behavior of the swarmof agents with respect to their absolute position on the Lie group:only relative positions (on the Lie group) matter. For instance,the configuration of a rigid body in the 3-D physical world isgiven by an orientation and a position vector in , whose com-bination corresponds to a position on Lie group . For rigidbody coordination, it is then natural to write control laws thatcan be interpreted as internal forces in the swarm, rather thanforces depending on an external reference frame which wouldprivilegiate some arbitrary choice of orientation and origin. In-dependence with respect to reference frame corresponds to in-variance with respect to applying to all agents the same Liegroup translation on .

The symmetries determine how to define meaningful quanti-ties for the swarm, like “relative positions” on the Lie group, andwhat the dynamicsof thecoupledagents canbe. Coordinatedmo-tion—in short coordination—is defined as all situations whererelative positions on the Lie group are fixed. Feedback controllaws that asymptotically enforce coordination must be designedon the basis of error measurements involving appropriately in-variant quantities (e.g., relative agent positions on the Lie group).

Previous Work: Results about synchronization (“reachinga common point”) and coordinated motion (“moving in an or-ganized way”) on vector spaces are becoming well established[24], [26]–[28]. Because a vector space can be identified withits tangent plane, both synchronization and coordinated motioncan be seen as consensus problems on the same vector space: theformer is a position consensus while the latter is a velocity con-sensus. Note that considering the motion of agents with the Liegroup structure of implies that only position vectors inand associated translational motion are covered. In contrast, assoon as orientation/rotation of the vehicles or of the formationmoving in a vector space is considered, the configuration spacebecomes the non-trivial Lie group . In general, when theconfiguration space is a Lie group, synchronization and coordi-nated motion are fundamentally different. The geometric view-point for dynamical systems on Lie groups is very well studied;see basic results in [33], [34] for simplified dynamics like thoseconsidered in the present paper, and [34]–[37] for a geometrictheory of mechanical systems on Lie groups. General resultsfor synchronization on compact Lie groups are proposed in [38],which points to related examples in the literature. But to the bestof the authors’ knowledge, a unified geometric viewpoint for co-ordinated motion—in short coordination—on Lie groups is stilllacking. Close to the present paper in its geometric flavor, [39]builds invariant observers for systems with Lie group symme-tries; observer design can be seen as two-agent leader-followersynchronization.

0018-9286/$26.00 © 2010 IEEE

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 2: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1048 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

In applications, the ubiquitous example of motion on Liegroups is a rigid body in . When translational motion isdiscarded, the configuration space reduces to the compact Liegroup characterizing the body’s orientation; an elementof can be represented by the rotation matrixbetween a frame attached to the rigid body and a hypothet-ical fixed reference frame. The standard example of this typeis satellite attitude control, where synchronization, i.e., ob-taining equal orientations, has recently attracted much attention[10]–[13], [15], [18], [31], [40]–[43], with and without externalreference tracking; note that synchronization is a very specialcase of coordination. Considering rotations and translations,the configuration space of an -dimensional rigid body be-comes the non-compact Lie group .Recently, coordinated motion has been investigated on[8], [20], [21] and [9], [16], [17], [19] in the underactu-ated setting of steering control where the linear velocity is fixedin the body’s frame. Motion on with steering control isalso directly linked to the evolution of a Serret-Frenet framewith curvature control, as explained in [33]. Results taking intoaccount the full mechanical dynamics for rigid body motionare more difficult to obtain—see for instance applications ofthe framework of [35] for coordination on andin [18], [43] and [16], [17] respectively. Considering simplifieddynamics, as in the present paper, can be useful either tobuild a high-level planning controller or as a preliminary steptowards an integrated mechanical controller, as illustrated forsynchronization on in [31] and [32], [44] respectively.

Contributions: The main goal of the present paper is to pro-vide a unified geometric framework for coordinated motion onLie groups, proceeding as follows. (i)Coordinationon Liegroupsis defined from first principles of symmetry, distinguishing threevariants: left-invariant, right-invariant and biinvariant coordina-tion. (ii) Expressing the conditions for coordination in the associ-ated Lie algebra, a direct link is drawn between coordination onLie groups and consensus in vector spaces. (iii) It is investigatedhow biinvariant coordination restricts compatible relative posi-tions through a geometrically meaningful relation. These prop-erties are independent of the dynamics. Going over to controllaws, simplified first-order dynamics are assumed for individualagents, but underactuation is explicitly modeled; communica-tion among agents is restricted to a reduced set of links that canpossibly be directed and time-varying. (iv) Control laws basedon standard vector space consensus algorithms are given thatachieve the easier tasks of right-invariant coordination and fullyactuated left-invariant coordination for any initial condition ongeneral Lie groups. (v) A general method is proposed to designcontrol laws that achieve biinvariant coordination of fully actu-ated agents when communication links are undirected and fixed;extension to more general communication settings can be madealong the lines of [21]. Biinvariant coordination is a rather aca-demic problem, but (vi) the proposed design method is shown toapply to the practically most relevant problem of left-invariantcoordinationofunderactuatedagents.Theproposedcontrollerar-chitectureconsistsoftwosteps,addingtotheconsensusalgorithmapositioncontrollerderivedfromgeometricLyapunovfunctions.The position controllers are directly linked to the double bracketflows of [45] for gradient systems on adjoint orbits.

The power of the geometry is illustrated on ,and by analyzing the meaning of the geometric condi-tions for coordination, and by designing corresponding controllaws with the proposed general methodology. The obtained con-trollers have been previously proposed in the literature, but werederived on the basis of intuitive arguments for particular appli-cations. In that sense, the novelty of the present paper is not inthe expression of the control laws but in showing that they canbe derived in a unifying and systematic manner with the propergeometric setting.

The present paper focuses on the achievement of coordinatedmotion only, in the sense that the objective is for the swarm tomove and conserve relative positions on the Lie group; the ac-tual values of the relative positions on the Lie group, as long asthey are compatible with the coordinated motion, are not con-trolled. However, applications often require to stabilize partic-ular relative positions on the Lie group which are more efficientthan others e.g., for sensing, power consumption or at least col-lision avoidance. The focus of the present work—motion withfixed relative positions on the Lie group—can be viewed as “or-thogonal” to driving the agents towards particular relative posi-tions on the Lie group. Therefore it is expected that the results ofthe present work can be combined with appropriately invariantrelative position control algorithms on the Lie group (as e.g.,from [38]), in order to both reach a particular configuration ofrelative positions on the Lie group and stabilize a coordinatedmotion of the resulting configuration. A corresponding resultis proposed in [20] for steering control of planar vehicles (Liegroup ); remaining issues concerning a general theory forthis combination are discussed in [46].

Table of Contents: The paper is organized as follows. Sec-tion II examines the geometric properties of coordination on Liegroups (contributions (i), (ii), and (iii)). Section III presents thecontrol setting and basic control laws for right-invariant coordi-nationandfullyactuatedleft-invariantcoordination(contribution(iv)). Sections IV and V present control law design methods re-spectively for biinvariant coordination [contribution (v)] and forunderactuated left-invariant coordination [contribution (vi)]. Ex-amples are treated at the ends of Section II, Sections IV and V.

II. THE GEOMETRY OF COORDINATION

This section proposes definitions for coordination on Liegroups by starting from basic symmetry principles. It estab-lishes conditions on velocities for coordination and examinesimplications. Except that the symmetries must be compatible,these developments are independent of the dynamics consid-ered for the control problem. Notations are adapted from [34].

A. Relative Positions and Coordination

Consider “agents” evolving on a Lie group , withdenoting the position of agent at time . Let denote the

group inverse of , denote left multiplication,and right multiplication on .

Definition 1: The left-invariant relative position onof agent with respect to agent is . Theright-invariant relative position on of with respect to is

.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 3: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

SARLETTE et al.: COORDINATED MOTION DESIGN ON LIE GROUPS 1049

Indeed, (resp. ) is invariant under left (resp. right)multiplication: . Left-/right-in-variant relative positions are the joint invariants associated toleft-/right-invariant action of on ( copies).In the following, “relative positions” always refer to relative po-sitions on unless otherwise specified.

The two definitions of relative position lead to two types ofcoordination; a third type is defined by combining them.

Definition 2: Left-invariant coordination (LIC) means con-stant left-invariant relative positions —resp.right-invariant coordination (RIC) means constant right-in-variant relative positions —for all pairs of agents

. Biinvariant coordination (BIC) means simultaneous LICand RIC: and are constant for all .

The present paper thus associates coordination to fixed rela-tive positions. In contrast, synchronization is the situation whereall agents are at the same point on : ; thisis a very particular case of biinvariant coordination.

B. Velocities and Coordination

Denote by the Lie algebra of , i.e., its tangent space atidentity . This paper always considers endowed with the Eu-clidean metric. Denote by the Lie bracket on . Let

and be the maps on tan-gent spaces induced by and respectively. Let

denote the adjoint representation.Definition 4: Left-invariant velocity and right-in-

variant velocity of agent are defined byand .

Indeed, and (resp. ) have the same left-invariant (resp. right-invariant) velocity (resp. ), forany fixed . Note the important equality

(1)

The adjoint orbit of is set .Proposition 1: Left-invariant coordination corresponds to

equal right-invariant velocities . Right-invariantcoordination corresponds to equal left-invariant velocities

.Proof: For , .

But if , then .

Thus

. Since and

are invertible, is equivalent toor equivalently . The proof for right-invariant coordi-nation is strictly analogous.

Proposition 1 shows that coordination on the Lie groupis equivalent to consensus in the vector space . Consensusin vector spaces is well-studied, see [4], [24]–[28], [47]. Biin-variant coordination requires simultaneous consensus on and

; but the latter are not independent, they are linked through (1)which depends on the agents’ positions.

Proposition 2: Biinvariant coordination on a Lie group isequivalent to the following condition in the Lie algebra :

Proof: RIC requires ; denote the commonvalue of the by . Then LIC requires

. The proof with is similar.Proposition 2 shows that biinvariant coordination puts no con-

straints on the relative positions when the group is Abelian,since in this case. In contrast, on a generalLie group, biinvariant coordination with non-zero velocity canrestrict the set of possible relative positions as follows.

Proposition 3: Let .a. For every , is a subgroup of .b. The Lie algebra of is the kernel of , i.e.,

.Proof:

a. since is the identity operator.implies by simple inversion of the relation.Moreover, if and , then

. Thus satisfies all groupaxioms and must be a subgroup of .

b. Let with and . Thenthe tangent space to at . For constant ,

implies , with the basic Liegroup property . Thereforeis necessary. It is also sufficient since, for any such that

, the group exponential curvebelongs to .

and are called the isotropy subgroup and isotropyLie algebra of ; these are classical objects in group theory [35].From Propositions 2 and 3, one method to obtain a biinvariantlycoordinated motion on is to (1) choose in the vector spaceand set (2) position the agents on such that

for pairs corresponding to the edges of an undirectedtree graph; the Lie group property of then ensures that

for all pairs . The same can be done withand the . Note that a swarm at rest isalways biinvariantly coordinated.

Remark 1: In many applications involving coordinated mo-tion, reaching a particular configuration, i.e., specific values ofthe relative positions, is also relevant. Specific configurationsare defined as extrema of a cost function in [38]. Imposing rela-tive positions in the (intersection of) set(s) for some canbe another way to classify specific configurations; unlike [38],it works for non-compact Lie groups. For compact groups, thereseems to be no connection between configurations characterizedthrough and those defined by [38].

Remark 2: One can also first fix relative positions andthen characterize the set of velocities compatible with biin-variant coordination. For non-Abelian groups and a sufficientlylarge number of agents, this set generically reduces to .

C. Examples

The Lie group has trivial properties; it is presented toclarify the distinction with “motion of rigid bodies in ”,whose configuration space is the Lie group . Basicproperties for the special orthogonal groups and specialEuclidean groups , , can be found in e.g., [33].Left-invariant coordination for and was alreadyformulated in Lie group notation in [8], [9].

Example 1: : For , a point is denoted bya position vector .

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 4: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1050 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

Fig. 1. Coordinated motion on ���� � ��� � ����. (a) An initial situa-tion of coordinated motion; circles represent agent positions, arrows their linearvelocities. (b) The same swarm in coordinated motion with a different velocity.(c) A coordinated motion that the swarm of the first plot cannot reach withoutbreaking the coordinated motion to re-position the agents.

• Group multiplication corresponds to , inverseto , and identity to position vector 0. In partic-

ular, the group structure is decoupled in each coordinateand Abelian (i.e., group multiplication is commutative).Relative positions take the familiar form

.• The Lie algebra equals itself, operations and

reduce to the identity .• Adjoint operator for all and the

Lie bracket is identically zero.• LIC, RIC and BIC all collapse to the same and just require

identical linear velocities in ; in particular BIC impliesno restrictions on relative positions. Physically, coordinatedmotion means a rigid formation of points in moving witha fixed formation orientation. The direction of motion canchange when varying the velocity vector, as between Fig.1(a) and 1(b), but a rotation of the formation, as going fromFig. 1(a)–1(c), would require breaking coordination in .

Example 2: : The special orthogonal group de-scribes 3-D rotations. A point on is represented by amatrix with and .

• Group multiplication, inverse and identity are the corre-sponding matrix operations.

• The Lie algebra is the set of skew-symmetric 3 3matrices , operations and are representedby and , respectively. The invertible mapping

identifies with .• With this identification, and

(vector product).• In the standard interpretation of as rigid body orienta-

tion, and are the angular velocities expressed in bodyframe and in inertial frame respectively.

• LIC (equal ), RIC (equal ) and BIC have a clear phys-ical interpretation in this case.

• For BIC with , and. The dimension of (

of ) is 1. Agents in BIC rotate with the same angularvelocity in inertial space and have the same orientationup to a rotation around .

Example 3: : The special Euclidean group in the planedescribes planar rigid body motions (translations and

rotations). An element of can be writtenwhere is a position vector in the plane and is orien-

tation (or “heading”).

• Group multiplication whereis the rotation of angle . Identity and inverse

.• Lie algebra . Operations

and .• and

.• In the interpretation of rigid body motion, is the linear

velocity expressed in body frame, is the ro-tation rate. For , is not the body’s linear velocityexpressed in inertial frame; instead, is thecenter of the circle drawn by the rigid body moving with

. In [20], the intuitive argument to achieve co-ordination is to synchronize circle centers ; this actuallysynchronizes right-invariant velocities .

• In RIC, the agents move with the same velocity expressedin body frame (Fig. 2, ). In LIC, they move like a singlerigid body (or “formation”): relative orientations and rel-ative position vectors on the plane do not change (Fig. 2,

and ). Note that any combination of translation (as onFig. 2, ) and rotation (as on Fig. 2, ) of the formationcomposed by the agents is possible.

• In BIC, the swarm moves like a single rigid body andeach agent has the same velocity expressed in body frame.Propositions 2 and 3 characterize by

and by. This leads to three different cases:

(o) and .(i) , and

.(ii) , any .Define , the circle of radius containingthe origin, tangent to at the origin and such thatand imply rotation in the same direction. Then solving

for and making a few calculations shows thatand tangent to at .

This is consistent with an intuitive analysis of possibilitiesfor circular motion with unitary linear velocity and fixedrelative position vectors and orientations in the plane.

The dimension of ( of ) is (o) 3, (i) 2 or (ii)1. In case (o), the configuration is arbitrary but at rest. In case(i), the agents have the same orientation and move on parallelstraight lines (Fig. 2, ). In case (ii), they move on the samecircle and have the same orientation with respect to their localradius (Fig. 2, ). Unlike for LIC, combinations of translations

and rotations of the formation composed by the agentswould not correspond to BIC.

Example 4: : This group describes 3-D rigid body mo-tions (translations and rotations). An element of can bewritten , with a position vector in

and a rotation matrix describing orientation.• , identity and

inverse .• Lie algebra is

identified with with the same mapping asfor . Operations and

. As for , symbol“ ” denotes vector product.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 5: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

SARLETTE et al.: COORDINATED MOTION DESIGN ON LIE GROUPS 1051

Fig. 2. Coordinated swarms (light color: intermediate planar positions and ori-entations in time). �: RIC with varying velocity. � and � : LIC with � � �and � �� � respectively; note that any combination of translation �� � and ro-tation �� � of the formation composed by the agents still corresponds to LIC.� and � : BIC with � � � and � �� � respectively; note that combinationsof translations �� � and rotations �� � of the formation composed by the agentswould not correspond to BIC.

• and.

• In the interpretation of rigid body motion, left-invariant ve-locities and are the body’s linear and angular velocityrespectively, expressed in body frame; the right-invariant

is the angular velocity expressed in inertial frame; for, a physical interpretation for the right-invariant

is unclear.• Similarly to , the agents move in RIC with the same

velocity expressed in body frame; they move in LIC withfixed relative orientations and relative position vectors in

, like a single rigid body.• In BIC, the swarm moves like a single rigid body and

each agent has the same velocity expressed in body frame.Propositions 2 and 3 lead to three different cases charac-terizing

which requires

which requires

(o) and .(i) ,and characterizes rotation ofaxis .(ii) , any

and describingleft-invariant relative positions of agents that are on thesame cylinder of axis and radius , withorientations differing around axis by an angle exactlyequal to their relative angular position on the cylinder .

This is again obtained by solving for in andmaking several basic computations; it is less obvious than for

to find this result intuitively.The dimension of ( of ) is (o) 6, (i) 4 or (ii)

2. In case (o), the configuration is arbitrary but at rest. In case(i), the agents move on parallel straight lines and have the sameorientation up to rotation around their linear velocity vector. Incase (ii), for , the agents draw helices ofconstant pitch on the cylinder; the special case

gives circular trajectories (see figures in [9], [19]).In the degenerate situation , all agents areon the rotation axis.

III. COORDINATION BY CONSENSUS IN THE LIE ALGEBRA

A. Control Setting

Left-invariant1 systems on Lie groups appear naturally inmany physical systems, such as rigid bodies in space andcart-like vehicles. Motivated by examples like 2-axes attitudecontrol and steering control on or , this paperconsiders left-invariant dynamics with affine control

(2)

where the Lie algebra is identified with , is a con-stant drift velocity, has full column rank and speci-fies the range of the control term ; without loss of gen-erality, the column vectors of are assumed orthonormal. Theset of all assignable is denoted . Forfully actuated agents , (2) simplifies towithout loss of generality. The following always considers en-dowed with the Euclidean metric. Feedback control laws mustbe functions of variables which are compatible with the symme-tries of the problem setting, i.e., left-invariant. In terms of left-invariant variables, LIC corresponds to fixed (left-invariant) rel-ative positions, while RIC corresponds to equal (left-invariant)velocities.

In a realistic scalable setting, full communication between allagents cannot be assumed. The information flow among agentsis modeled by a restricted set of communication links;

denotes that sends information to . The communicationtopology is associated to a graph . is undirected if

1A right-invariant system is equivalent, simply by redefining the groupmultiplications.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 6: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1052 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

. is uniformly connected (see [24], [25]) if thereexist an agent and durations and such that, ,taking the union of the links appearing for at least in time span

, there is a directed path fromto every other agent .

B. Right-Invariant Coordination

Right-invariant coordination requires . In thesetting (2), this simply implies to agree on equal ; posi-tions can evolve arbitrarily. This problem is solved by theclassical vector space consensus algorithm [4], [25]–[28], [47]

(3)

Using (2), it translates into . It ex-ponentially achieves if is uniformly connected.Asymptotic RIC is then ensured for any initial and, of course,any relative positions which actually have no influence.Agent relies on the left-invariant velocity of .

For a time-invariant and undirected communication graph ,(3) is a gradient descent for the disagreement cost function

with the Euclidean metric in .

C. Left-Invariant Coordination

Left-invariant coordination requires , whichsuggests to use

(4)

Using (1), in terms of the left-invariant variables, (4) becomes

(5)

thanks to . To implement (4),agent must know the relative position and velocityof .

A priori, (5) converges as (3), ensuring global exponentialcoordination for uniformly connected . However, in contrastto (3), nothing guarantees that (5) can be implemented in anunderactuated setting. At equilibrium, (5) requires

(6)

which, for arbitrary relative positions of the agents, might admitno solution . This issue motivates the furtherstudy of underactuated LIC in Section V. Similarly, biinvariantcoordination requires simultaneous consensus on left- and right-invariant velocities. At equilibrium, this means that (6) musthold with equal controls , i.e.,

(7)

which also puts constraints on the relative positions of theagents. For this reason, biinvariant coordination is furtherstudied in Section IV.

The cost function asso-ciated to (4) is not left-invariant in general (it involves positions

), so (5) cannot be a left-invariant gradient of .Nevertheless, let be the subclass of compact groups

with unitary adjoint representation, i.e., satisfyingand (for instance ). It is

possible to define a biinvariant (that is, left- and right-invariant)Riemannian metric on if and only if [48]. Usingthe Euclidean metric on left-invariant velocities, as in thepresent paper, comes down to using a left-invariant metric, inaccordance with the left-invariant setting. If , then thismetric is biinvariant, andfor fixed undirected , (5) is a gradient descent for .

In the following, it is assumed that the agents are controllable.Obviously, controllability is sufficient for coordination as it al-lows the agents to reach any position from any initial condition.However, it is not always necessary, as long as positions com-patible with (6) or (7) are globally reachable; in particular, forAbelian groups, all positions satisfy (6) and (7); in that case,(underactuated) LIC and BIC become trivial.

IV. CONTROL DESIGN: FULLY ACTUATED

BIINVARIANT COORDINATION

A. Biinvariant Coordination on General Lie Groups

Biinvariant coordination requires to satisfy two objectives,LIC and RIC, simultaneously. In a first step, assume that theagents are given a reference right-invariant velocity , such thatLIC is ensured if each agent applies velocity . Itremains to simultaneously achieve RIC, which, as previouslyshown, involves controlling relative positions. Write a generalcontroller

(8)

where is a desired velocity and is necessary for relative po-sition control. Thus for the present, . The questionis how to design in order to achieve BIC. For fixed undirectedcommunication graph , inspired by the cost function for RIC,define

where denotes Euclidean norm. characterizes the dis-tance from RIC assuming that every agent has velocity

. Since , the timevariation of due to motion of is

(9)

where denotes the canonical scalar product in , defined withthe Euclidean metric. Thus if then ; a properchoice of should allow to decrease . Define2 the bracket

2In fact �� � expresses the effect of the Lie bracket on the dual space of .It is directly related to the coadjoint representation of �, commonly used formechanical systems; in general, �� � does not satisfy the Lie bracket properties.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 7: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

SARLETTE et al.: COORDINATED MOTION DESIGN ON LIE GROUPS 1053

Fig. 3. Biinvariant coordination as consensus on right-invariant velocity andLyapunov-based control to right-invariant coordination.

such that .Then (9) rewrites andthe choice

(10)

ensures that is non-increasing along the solutions:

To obtain an autonomous, left-invariant algorithm for biin-variant coordination, it remains to replace the reference velocity

by estimates on which the agents progressively agree. Sincethe goal is to define a common right-invariant velocity in , itis natural to proceed as in Section III-C and use the consensusalgorithm

(11)

which in terms of left-invariant velocities rewrites

(12)

Thus the overall controller is the cascade of a consensus al-gorithm to agree on a desired velocity for LIC, and a positioncontroller designed to decrease a natural distance to RIC. Toimplement the controller, agent must receive from communi-cating agents their relative positions and the valuesof their left-invariant auxiliary variables (see Fig. 3).

The following result characterizes the convergence propertiesof controller (8), (10), (12).

Theorem 1: Consider fully actuated agents communi-cating on a fixed, undirected graph and evolving on Lie group

according to with controller (8), (10), (12).(i) For any initial conditions , the

exponentially converge to .(ii) Define

All solutions converge to the critical set of . Inparticular, left-invariant coordination is asymptoticallyachieved for all initial conditions.

(iii) Biinvariant coordination is (at least locally) asymptoti-cally stable.

Proof: Regarding convergence, (12) is strictly equivalentto (11). Therefore, (i) simply restates a well-known convergenceresult for consensus algorithms in vector spaces on fixed undi-rected graphs [26].

Fig. 4. Biinvariant coordination as consensus on left-invariant velocity andLyapunov-based control to left-invariant coordination.

Since the converge, (8), (10) is an asymptotically au-tonomous system; the autonomous limit system is obtained byreplacing . From the derivation of in (10),the limit system is a gradient descent for ,which is smooth because the adjoint representation is smooth.According to [49], the -limit sets of an asymptotically au-tonomous system correspond to the chain recurrent sets of thelimit system. From [50] the chain recurrent set of a smoothgradient system is equal to its critical points. Therefore the

-limit set of (8), (10) is equal to the critical points of ,which proves (ii). Biinvariant coordination is locallyasymptotically stable as it is a local (and global) minimum of

, which proves (iii).Given , the region of attraction for BIC is a sublevel-set

where has 0 as only critical point (in practice, as only min-imum). Other local minima can involve e.g., the evenly dis-tributed on a circular with a ring graph (see [38]).

Extensions to varying and directed can be made with ad-ditional auxiliary variables along the lines of [19], [21], [51],[52]: at a first level, consensus algorithms define a desiredand a desired , which must be on the same adjoint orbit;at a second level, cost functions for individual agents ensurethat they asymptotically implement the desired velocities. Thedouble-consensus part is non-trivial to write in a fully left-in-variant setting, because and must belong to the same ad-joint orbit. The present paper proposes no explicit design of thisform. For fixed undirected , an advantage of algorithms with“double consensus” ( and ) would be that BIC becomes theonly locally stable equilibrium: interaction-related issues onlydepend on the performance of the consensus algorithm, for therest the agents behave individually. It is shown in [38] how aux-iliary variables can be used to build consensus algorithms thatavoid spurious local minima on various spaces.

B. Biinvariant Coordination on Lie Groups With a BiinvariantMetric

When , i.e., has a biinvariant metric, the cost func-tion can be used forleft-invariant control design.

A natural idea in this context would be to combine the costfunctions for LIC and RIC, writing , and derive agradient descent for of the form

. However, simulations of the resulting control law foralways converge to . A possible explanation

for this behavior is that the gradient controls velocities, not ex-plicitly positions, while it was shown in Section II that BIC atnon-zero velocity involves restrictions on compatible positions.

Nevertheless, the biinvariant metric allows to switch the rolesof LIC and RIC in the method of Section IV-A, using a con-sensus algorithm to define a common left-invariant velocity forRIC, and a cost function to drive positions to LIC (see Fig. 4).

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 8: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1054 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

The RIC consensus algorithm on auxiliary variables asymp-totically defines a common velocity by

(13)

Then defining the cost function

for LIC and proceeding as in the previous subsection, one ob-tains controller (8) with

(14)

Theorem 2: Consider fully actuated agents communi-cating on a connected, fixed, undirected graph and evolvingon according to with controller (8),(13), (14).

(i) For any initial conditions , the exponentiallyconverge to .

(ii) Define

. All solutions converge to the critical set of .In particular, right-invariant coordination is asymptoti-cally achieved.

(iii) Biinvariant coordination is (at least locally) asymptoti-cally stable.

Proof: The proof is omitted because it is similar to the oneof Theorem 1.

The region of attraction for BIC behaves as for Theorem 1.An advantage of Theorem 2 over Theorem 1 is that control

design is directly extended to underactuated agents. Indeed, (13)defines a valid consensus velocityfor underactuated agents provided that . The onlychange is that , instead of the exact gradient descent in (14),is its projection onto the control range of :

When is asymptotically defined with (13), the convergenceargument for asymptotically autonomous systems must be ex-tended to projections of gradient systems; a general proof ofthis technical issue is lacking in the present paper. It is the onlyreason to restrict Theorem 2 to fully actuated agents.

Brockett [45] has developed a general double-bracketform for gradient algorithms on adjoint orbits of compactsemi-simple groups, using the biinvariant Killing metric. Theconnection with the present paper is clear: once the consensusalgorithm has converged, the gradient control for agent posi-tions involves a cost function on the adjoint orbit of or .One example in [45] involves minimizing the distance towardsa subset of ; a similar objective will be pursued in Section V ofthe present paper (but with a different class of subsets). A maindifference of [45] is its focus on the evolution of variables in ,making abstraction of the underlying group, while the presentpaper actually controls positions of (possibly underactuated)

agents on . If is a compact group and the biinvariant Killingmetric coincides with the left-invariant metric of the presentpaper, then and control (10) for withfixed implies that follows the double bracket flow

(15)

This is the case among others for .

C. Example: Biinvariant Coordination in

Control laws for coordination in abound in the lit-erature—see among others papers about satellite attitude con-trol mentioned in the Introduction. Biinvariant coordination on

requires aligned rotation axes, and thus synchronizessatellite attitudes up to their phase around the rotation axis.

The compact group has a biinvariant metric, so Sec-tion IV-B applies. Algorithm (13) is used verbatim, with

the auxiliary variable associated to angular velocity . Asmentioned before (15), on . Thus in the fullyactuated case, (8), (14) lead to

(16)

Theorem 2 can be strengthened as follows for specific graphs.Proposition 4: If is a tree or complete graph, then BIC is

the only asymptotically stable limit set.Proof: According to Theorem 2, it remains to show that

BIC is the only local minimum of . Fixing ,critical points of correspond to

(17)

For the tree, start with the leaves . Thenwhere is the parent of . As a consequence, (17) for the

parent becomes where is the parentof . Using this argument up to the root, all must beparallel. If the agents are partitioned in two anti-aligned groups,then moving those groups towards each other decreases ; thus

is the only local minimum. For the complete graph,(17) becomes , where . Thisimplies either that all must be parallel or that .In the first case, further discussion is as for the tree. Rewriting

shows that corresponds to amaximum of .

Combining trees and cliques can yield more graphs with BICas only asymptotically stable limit set. For others, local minimamay exist. Classifying local minima of from graph propertiesis an open question.

It is straightforward to adapt (16) for underactuated agents;a popular underactuation on is to consider 2 orthogonalaxes of allowed rotations and , either controlling both ro-tation rates, i.e., , or imposing a fixed rotationrate around one axis, i.e., . Both cases are con-trollable [33], so the Jurdjevic-Quinn theorem [53] ensures localasymptotic stability of BIC, if is fixed in advanceor agreed on in finite time. A formal convergence proof for the

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 9: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

SARLETTE et al.: COORDINATED MOTION DESIGN ON LIE GROUPS 1055

asymptotically autonomous case where the follow (13) is cur-rently missing.

V. CONTROL DESIGN: UNDERACTUATED

LEFT-INVARIANT COORDINATION

Biinvariant coordination may appear as a rather academic ob-jective, whose motivation in applications is not clear. However,the methodology developed in Section IV for BIC control de-sign is instrumental to achieve left-invariant coordination of un-deractuated agents. The latter is well motivated by practical ap-plications. Here the role of the cost function is no longer to adda second level of coordination, but to fulfill the underactuationconstraints. Unlike the academic problem setting of BIC, thepresent section explicitly considers the most general setting ofpossibly directed and time-varying interconnection graph .

A. Left-Invariant Coordination of Underactuated Agents

The control design for underactuated LIC is decomposedin the two steps illustrated in Fig. 5. Analogously to thebiinvariant coordination design of Section IV-A, a feasibleright-invariant velocity is determined by a consensus algo-rithm. The corresponding left-invariant velocity is enforced bya Lyapunov-based feedback that decreases its distance from

.The consensus algorithm must enforce a feasible right-in-

variant velocity, that is a vector in the set

If is convex, then it is sufficient to initialize the consensusalgorithm (12) with . When is not convex, theconsensus algorithm must be adapted and the present paper hasno general method. Strategies inspired from [38] for compacthomogeneous manifolds may be helpful, as illustrated in theexample below.

Now assuming a known feasible right-invariant velocity ,the design of a Lyapunov based control to left-invariant coordi-nation proceeds similarly to Section IV-A.

Define to be the Euclidean distance in from tothe set . Let be the projection of on ; since isconvex, is the unique point in such that

. Following the same steps as inSection IV-A, define . Writing

(18)

the task is to design such that asymptotically, isdriven to a point where and converges to 0; this wouldasymptotically ensure LIC. For each individual agent , writethe cost function

where denotes Euclidean norm. characterizes the distancefrom to , that is the distance from LIC assuming that everyagent implements . The time variation ofdue to motion of is

(19)

Fig. 5. Underactuated left-invariant coordination as constrained consensuson right-invariant velocity and Lyapunov-based control to left-invariantcoordination.

where denotes the canonical scalar product in . To go on alongthe lines of Section IV-A, it must hold

; this condition on Lie algebra structure and con-trol setting is satisfied for examples below. Then (19) implies

, where

(20)

when identifying with , and a natural control is

(21)

Note that when , the position control is unnecessaryand vanishes, yielding simply .

The overall controller is the cascade of a consensus algorithmto agree on a desired velocity for LIC, and a position controllerdesigned from a natural Lyapunov function to reach positionscompatible with underactuation constraints and so to actuallyachieve LIC. To implement the controller, agent must get fromother agents their relative positions and the valuesof their left-invariant auxiliary variables . Since agents onlyinteract through the consensus algorithm, not through the costfunction, a connected fixed undirected graph is not required:

can be directed and time-varying, as long as it remains uni-formly connected (see Fig. 5).

A general characterization of the behavior of solutions of theclosed-loop system is more difficult here because the positioncontroller is not a gradient anymore. A crucial step for which thepresent paper proposes no explicit general solution is the designof an appropriate consensus algorithm on auxiliary variables.The other assumptions in the following result can be readilychecked for any particular case.

Theorem 3: Consider underactuated agents communi-cating on a uniformly connected graph and evolving on Liegroup according to with controller (18), (21)where is defined in (20), assuming that , it holds

. Assume that an appropriate con-sensus algorithm drives the arbitrarily initiated , ,such that they exponentially agree on ,independently of the agent motions .

(i) If the agents are controllable, then LIC is locally asymp-totically stable.

(ii) If, for any fixed , bounded implies bounded, and implies ,

then all agent trajectories on converge to the set where.

Proof: The overall system is a cascade of the exponen-tially stable consensus algorithm and position controller (18),(21) which is decoupled for the individual agents. Assumptions

and (21) exactly mean thatis non-increasing along the closed-loop solutions. Therefore, ifthe agents are controllable, Jurdjevic-Quinn theorem [53] im-plies local asymptotic stability of the local minimum

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 10: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1056 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

for the position controller. Then the overall system is the cas-cade of an exponentially stable system and a system for which

is locally asymptotically stable. Standard argu-ments on cascade systems (see e.g., [54], [55]) allow to concludethat is locally asymptotically stable for the overallsystem; this proves (i).

To prove (ii), first consider the case where constant. Then can only decrease, and since it is bounded from

below it tends to a limit; therefore is integrable in time for. For the same reason, is bounded, so according to

the assumption for (ii) is bounded as well; then ,which is a continuous function of , is bounded as well forthe closed-loop system, such that is uniformly continuousin time for . Barbalat’s Lemma implies that con-verges to 0, which implies that converges to 0, concludingthe proof. Now in fact varies, it exponentially converges tothe constant . This changes nothing to the fact that tendsto a finite limit and is bounded, so the same argumentapplies.

Condition is not always true when; however, it often holds in practice, as in the following

example on steering control of rigid bodies. For this example,Theorem 3 is improved by showing that LIC is the only stablelimit set. In general, possible improvements of the local stabilityresult depend on the geometry of and related consensus al-gorithms; particular settings of the literature feature fairly largeregions of attraction (at least in simulations).

B. Example: Steering Control on

Left-invariant coordination on under steering controlis studied in [19]. The present section shows how the algorithmsof [19] follow from the present general framework. Illustrationsof the algorithms by numerical simulation can also be found in[19].

Using the notations of Section II-C, the position and orien-tation of a rigid body in 3-D space is written ,which is an element of the Special Euclidean group ;group multiplication is the usual composition law for transla-tions and rotations, see Section II-C. Then requiring agents to“move in formation”, i.e., such that the relative position andheading of agent with respect to agent is fixed in the ref-erence frame of agent , , is equivalent to requiring left-in-variant coordination. Moreover, since linear and angular ve-locity in body frame correspond to the components of

, the problem of controlling each agent in its own frame withfeedback involving relative positions and orientations of otheragents only, fits the left-invariant problem setting described inSection III. The constraint of steering control—i.e., fixed linearvelocity in agent frame —implies (2) of the form

Steering controlled agents on are controllable [33].Following the method of Section V-A, write auxiliary vari-

ables ; then , cost functionand straightforward calculations show

that (19) becomes . This means thatand . Then (18),

(21) yield the controller

(22)

This is the same control law as derived in [19] from intuitive ar-guments. If an appropriate consensus algorithm is built, then allassumptions of Theorem 3 hold, implying local asymptotic sta-bility of 3-D “motion in formation” with steering control (22);in fact, [19] slightly improves Theorem 3 by also showing thatglobally, LIC is the only stable limit set.

It remains to design a consensus algorithm for the . Forthis, two cases are distinguished: linear motion and he-licoidal (of which a special case is circular) motion . Thefirst case (almost) never appears from a consensus algorithmwith arbitrary ; it can however be imposed by

, which will then remain true , in order to stabilize acoordinated motion in straight line.

• If (linear motion), then and. Agreement on

in the unit sphere can be achieved following [38], justachieving consensus in and normalizing; in fact nor-malizing is not even necessary, as it would just change thegain in (22). This leads to

(23)

for , again as in [19].• If , then and

, and. Designing a consensus algorithm, that

both achieves agreement on and can be writtenwith left-invariant variables, appears to be difficult. Sim-ilarly to the first case, suitable algorithms can be built ifthe overall dimension of the variables used for the con-sensus algorithm is enlarged with respect to the dimen-sion of the configuration space. The consensus algorithmproposed in [19] replaces by three components

, and associated with thevectors , , used to describe above; then

. The advantage of this em-bedding is that left-invariant consensusalgorithms can be decoupled for the , the and the .With the notations of the present paper, the correspondingconsensus algorithm proposed in [19] is

. Comparing left-invariant relative positionwith the terms and fac-

tors appearing in this consensus algorithm, one observesthat the latter is indeed left-invariant. It can be verified(see [19]) that this algorithm indeed synchronizes the

.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 11: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

SARLETTE et al.: COORDINATED MOTION DESIGN ON LIE GROUPS 1057

Remark 3: LIC with , under steering control re-quires to align vectors for all agents. This is in fact equiv-alent to BIC on with . The present sec-tion thus illustrates the method for BIC on for uniformlyconnected (instead of fixed undirected as in Section IV).

Remark 4: LIC under steering control on is treatedin [20], [21], where numerical simulations of the resulting al-gorithms can also be found. Like for , control algorithmsobtained intuitively, with several simplifications due to the lowerdimension, can be recovered with the general method of thepresent paper.

In fact, the group structure and control setting of steering con-trol on are such that and steering controls

, one has

with (24)

On explicitly,and , so

and . Then LIC automaticallyimplies equal , thus RIC, meaning that underactuated LIC isequivalent to BIC and imposes the same constraints on relativepositions . This is the case for any group and control settingsatisfying (24).

For steering control on , LIC is slightly different fromBIC because , so (24)would require which is not true ingeneral. Therefore, for LIC under steering control thecan differ by arbitrary rotations around , while BIC wouldrequire equal .

VI. CONCLUSION

This paper proposes a geometric framework for coordinationon general Lie groups and methods for the design of controllersdriving a swarm of underactuated, simple integrator agents to-wards coordination. It shows how the general framework pro-vides control laws for coordination of rigid bodies, on ,

and , and allows to easily handle different set-tings. Formal convergence results are local, but authors workingon particular applications have always observed fairly large re-gions of attraction (at least in simulations).

Following the numerous results about coordination on partic-ular Lie groups, various directions are still open to extend thegeneral framework of the present paper. A first case often en-countered in practice is to stabilize specific relative positions ofthe agents (“formation control”). In [20], [21] for instance, thesteering controlled agents on are not only coordinated ona circle, but regular distribution of the agents on the circle is alsostabilized; in the present paper, relative positions of the agentsare asymptotically fixed but arbitrary. The requirement of syn-chronization (most prominently, “attitude synchronization” on

) also fits in this category. A second important extensionwould be to consider more complex dynamics, like those of me-chanical systems.

REFERENCES

[1] S. Strogatz, Sync: The Emerging Science of Spontaneous Order. NewYork: Hyperion, 2003.

[2] S. Strogatz, “From Kuramoto to Crawford: Exploring the onset of syn-chronization in populations of coupled nonlinear oscillators,” PhysicaD, vol. 143, pp. 1–20, 2000.

[3] A. Jadbabaie, J. Lin, and A. Morse, “Coordination of groups of mobileautonomous agents using nearest neighbor rules,” IEEE Trans. Autom.Control, vol. 48, no. 6, pp. 988–1001, Jun. 2003.

[4] V. Blondel, J. Hendrickx, A. Olshevsky, and J. Tsitsiklis, “Conver-gence in multiagent coordination, consensus and flocking,” in Proc.44th IEEE Conf. Decision Control, 2005, pp. 2996–3000.

[5] J. Fax and R. Murray, “Information flow and cooperative control ofvehicle formations,” IEEE Trans. Autom. Control, vol. 49, no. 9, pp.1465–1476, Sep. 2004.

[6] J. Desai, J. Ostrowski, and V. Kumar, “Modeling and control of for-mations of nonholonomic mobile robots,” IEEE Trans. Robot. Autom.,vol. 17, no. 6, pp. 905–908, Jun. 2001.

[7] R. Olfati-Saber and R. Murray, “Graph rigidity and distributed forma-tion stabilization of multi-vehicle systems,” in Proc. 41st IEEE Conf.Decision Control, 2002, pp. 2965–2971.

[8] E. Justh and P. Krishnaprasad, “Equilibria and steering laws for planarformations,” Syst. Control Lett., vol. 52, pp. 25–38, 2004.

[9] E. Justh and P. Krishnaprasad, “Natural frames and interacting particlesin three dimensions,” in Proc. 44th IEEE Conf. Decision Control, 2005,pp. 2841–2846.

[10] M. Mesbahi and F. Hadaegh, “Formation flying control of multiplespacecraft via graphs, matrix inequalities, and switching,” J. Guid.,Control Dyn., vol. 24, no. 2, pp. 369–377, 2001.

[11] M. VanDyke and C. Hall, “Decentralized coordinated attitude controlof a formation of spacecraft,” J. Guid., Control Dyn., vol. 29, no. 5, pp.1101–1109, 2006.

[12] J. Lawton and R. Beard, “Synchronized multiple spacecraft rotations,”Automatica, vol. 38, pp. 1359–1364, 2002.

[13] T. Krogstad and J. Gravdahl, “Coordinated attitude control of satellitesin formation,” in Group Coordination and Cooperative Control, Ser.Lect. N. Control and Inform. Sci. New York: Springer, 2006, vol. 336,pp. 153–170.

[14] C. McInnes, “Distributed control for on-orbit assembly,” Adv. Astro-naut. Sci., vol. 90, pp. 2079–2092, 1996.

[15] W. Ren, “Distributed attitude consensus among multiple networkedspacecraft,” in Proc. Amer. Control Conf., 2006, [CD ROM].

[16] T. Smith, H. Hanssmann, and N. Leonard, “Orientation control of mul-tiple underwater vehicles with symmetry-breaking potentials,” in Proc.40th IEEE Conf. Decision Control, 2001, pp. 4598–4603.

[17] H. Hanssmann, N. Leonard, and T. Smith, “Symmetry and reductionfor coordinated rigid bodies,” Eur. J. Control, vol. 12, no. 2, pp.176–194, 2006.

[18] S. Nair, “Stabilization and Synchronization of Networked MechanicalSystems,” Ph.D. dissertation, Princeton Univ., Princeton, NJ, 2006.

[19] L. Scardovi, N. Leonard, and R. Sepulchre, “Stabilization of collectivemotion in three dimensions,” Commun. Inform. Syst. (Brockett LegacyIssue), vol. 8, no. 3, pp. 1521–1540, 2008.

[20] R. Sepulchre, D. Paley, and N. Leonard, “Stabilization of planar collec-tive motion with all-to-all communication,” IEEE Trans. Autom. Con-trol, vol. 52, no. 5, pp. 811–824, May 2007.

[21] R. Sepulchre, D. Paley, and N. Leonard, “Stabilization of planar collec-tive motion with limited communication,” IEEE Trans. Autom. Control,vol. 53, no. 3, pp. 706–719, 2008.

[22] N. Leonard, D. Paley, F. Lekien, R. Sepulchre, D. Frantantoni, and R.Davis, “Collective motion, sensor networks and ocean sampling,” Proc.IEEE, vol. 95, no. 1, pp. 48–74, Jan. 2007.

[23] D. Swain, N. Leonard, I. Couzin, A. Kao, and R. Sepulchre, “Alter-nating spatial patterns for coordinated group motion,” in Proc. 46thIEEE Conf. Decision Control, 2007, pp. 2925–2930.

[24] L. Moreau, “Stability of multi-agent systems with time-dependentcommunication links,” IEEE Trans. Autom. Control, vol. 50, no. 2, pp.169–182, Feb. 2005.

[25] L. Moreau, “Stability of continuous-time distributed consensus al-gorithms,” in Proc. 43rd IEEE Conf. Decision Control, 2004, pp.3998–4003.

[26] R. Olfati-Saber, J. Fax, and R. Murray, “Consensus and cooperationin networked multi-agent systems,” Proc. IEEE, vol. 95, no. 1, pp.215–233, Jan. 2007.

[27] R. Olfati-Saber and R. Murray, “Consensus problems in networks ofagents with switching topology and time delays,” IEEE Trans. Autom.Control, vol. 49, no. 9, pp. 1520–1533, Sep. 2004.

[28] J. Tsitsiklis, “Problems in Decentralized Decision Making and Com-putation,” Ph.D. dissertation, MIT, Cambridge, MA, 1984.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.

Page 12: IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, …asarlet/Data/sbs_tac2010.pdf · 2010-08-20 · IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010 1047 Coordinated

1058 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 55, NO. 5, MAY 2010

[29] J. Tsitsiklis, D. Bertsekas, and M. Athans, “Distributed asynchronousdeterministic and stochastic gradient optimization algorithms,” IEEETrans. Autom. Control, vol. AC-31, no. 9, pp. 803–812, Sep. 1986.

[30] H. Bai, M. Arcak, and J. Wen, “A decentralized design for group align-ment and synchronous rotation without inertial frame information,” inProc. 46th IEEE Conf. Decision Control, 2007, pp. 2552–2557.

[31] A. Sarlette, R. Sepulchre, and N. Leonard, “Cooperative attitude syn-chronization in satellite swarms: A consensus approach,” in Proc. 17thIFAC Symp. Autom. Control Aerosp., 2007, [CD ROM].

[32] A. Sarlette, R. Sepulchre, and N. Leonard, “Autonomous rigid bodyattitude synchronization,” in Proc. 46th IEEE Conf. Decision Control,2007, pp. 2566–2571.

[33] V. Jurdjevic, Geometric Control Theory. Cambridge, U.K.: Cam-bridge Univ. Press, 1997.

[34] V. Arnold, Mathematical Methods of Classical Mechanics. NewYork: Springer, 1989.

[35] J. Marsden and T. Ratiu, Introduction to Mechanics and Symmetry.New York: Springer, 1994.

[36] N. Leonard, “Averaging and Motion Control of Systems on LieGroups,” Ph.D. dissertation, Univ. Maryland, College Park, 1994.

[37] F. Bullo, “Nonlinear Control of Mechanical Systems: A RiemannianGeometry Approach,” Ph.D. dissertation, CalTech, Pasadena, CA,1998.

[38] A. Sarlette and R. Sepulchre, “Consensus optimization on manifolds,”SIAM J. Control Optim., vol. 48, no. 1, pp. 56–76, 2009.

[39] S. Bonnabel, P. Martin, and P. Rouchon, “Symmetry-preserving ob-servers,” IEEE Trans. Autom. Control, vol. 53, no. 11, pp. 2514–2526,2008.

[40] A. Bondhus, K. Pettersen, and J. Gravdahl, “Leader/follower synchro-nization of satellite attitude without angular velocity measurements,”in Proc. 44th IEEE Conf. Decision Control, 2005, pp. 7270–7277.

[41] C. Beugnon, E. Buvat, M. Kersten, and S. Boulade, “GNC designfor the DARWIN spaceborne interferometer,” in Proc. 6th ESA Conf.Guid., Navig. Control Syst., 2005, [CD ROM].

[42] S. Nair and N. Leonard, “Stabilization of a coordinated network of ro-tating rigid bodies,” in Proc. 43rd IEEE Conf. Decision Control, 2004,pp. 4690–4695.

[43] S. Nair and N. Leonard, “Stable synchronization of mechanical systemnetworks,” SIAM J. Control Optim., vol. 47, no. 2, pp. 661–683, 2008.

[44] A. Sarlette, R. Sepulchre, and N.E. Leonard, “Autonomous rigid bodyattitude synchronization,” Automatica, vol. 45, no. 2, pp. 572–577,2009.

[45] R. Brockett, “Differential geometry and the design of gradient algo-rithms,” in Proc. Symp. Pure Math., AMS, 1993, vol. 54, pp. 69–92.

[46] A. Sarlette, “Geometry and Symmetries in Coordination Control,”Ph.D. dissertation, Univ. Liège, Liège, Belgium, 2009.

[47] A. Olshevsky and J. Tsitsiklis, “Convergence rates in distributed con-sensus and averaging,” in Proc. 45th IEEE Conf. Decision Control,2006, pp. 3387–3392.

[48] V. Arsigny, X. Pennec, and N. Ayache, “Bi-invariant means in Liegroups. Application to left-invariant polyaffine transformations,”INRIA Res. Rep., 2006, vol. 5885.

[49] K. Mischaikow, H. Smith, and H. Thieme, “Asymptotically au-tonomous semi-flows, chain recurrence and Lyapunov functions,”AMS Trans., vol. 347, no. 5, pp. 1669–1685, 1995.

[50] M. Hurley, “Chain recurrence, semiflows, and gradients,” J. Dynam.Diff. Eq., vol. 7, no. 3, pp. 437–456, 1995.

[51] L. Scardovi and R. Sepulchre, “Collective optimization over averagequantities,” in Proc. 45th IEEE Conf. Decision Control, 2006, pp.3369–3374.

[52] L. Scardovi, A. Sarlette, and R. Sepulchre, “Synchronization and bal-ancing on the� -torus,” Syst. Control Lett., vol. 56, no. 5, pp. 335–341,2007.

[53] V. Jurdjevic and J. Quinn, “Controllability and stability,” J. Diff. Eq.,vol. 28, no. 3, pp. 381–389, 1978.

[54] E. D. Sontag, “Remarks on stabilization and input-to-state stability,” inProc. 28th IEEE Conf. Decision Control, 1989, pp. 1376–1378.

[55] R. Sepulchre, M. Jankovic, and P. Kokotovic, Constructive NonlinearControl. New York: Springer, 1997.

Alain Sarlette (M’09) received the M.S. degreein physical and space engineering and the Ph.D.degree in applied mathematics and control from theUniversity of Liège, Belgium, in 2005 and 2009,respectively.

He is currently a Postdoctoral Research Fellowat the Belgian FNRS, visiting at Mines Paris-Tech (2009–2010). His research interests includedistributed systems and swarm control, quantumcontrol, and space applications.

Dr. Sarlette received the 2009 EECI Ph.D. Awardon Embedded and Networked Control.

Silvère Bonnabel was born in Marseille, France, in1981. He received the M.S. degree in engineering andthe Ph.D. degree in mathematics and control fromMines ParisTech, Paris, France, in 2004 and 2007,respectively.

He was a Postdoctoral Fellow at the University ofLiège, Belgium, in 2008. He is currently a Researcherat Mines ParisTech. His research interests include ge-ometrical methods for filtering and control. He hasworked on industrial applications (chemical reactor).

Rodolphe Sepulchre (SM’09) received the M.S.degree in engineering and the Ph.D. degree in ap-plied mathematics from the University of Louvain,Belgium, in 1990 and 1994, respectively.

He is a Professor in the Department ElectricalEngineering and Computer Science, Universitéde Liège, Institut Montefiore B28, B-4000 Liège,Belgium. He held research or teaching positions atthe University of California, Santa Barbara (1994 to1996), Princeton University, Princeton, NJ (2002 to2003), and the University of Louvain (2001 to 2009).

His research interests include nonlinear control systems, dynamical systems,and optimization on manifolds.

Dr. Sepulchre received the 2008 IEEE Antonio Ruberti Prize.

Authorized licensed use limited to: ECOLE DES MINES PARIS. Downloaded on May 07,2010 at 12:09:59 UTC from IEEE Xplore. Restrictions apply.