This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Geometric sliding mode control of mechanical systems on Lie groups

Eduardo Espindola eespindola@comunidad.unam.mx    Yu Tang tang@unam.mx Ningbo Institute of Technology, Zhejiang University, Ningbo, CHINA.
Abstract

This paper presents a generalization of conventional sliding mode control designs for systems in Euclidean spaces to fully-actuated simple mechanical systems whose configuration space is a Lie group for the trajectory-tracking problem. A generic kinematic control is first devised in the underlying Lie algebra, which enables the construction of a Lie group on the tangent bundle where the system state evolves. A sliding subgroup is then proposed on the tangent bundle with the desired sliding properties, and a control law is designed for the error dynamics trajectories to reach the sliding subgroup globally exponentially. Tracking control is then composed of the reaching law and sliding mode, and is applied for attitude tracking on the special orthogonal group SO(3)SO(3) and the unit sphere 𝒮3\mathcal{S}^{3}. Numerical simulations show the performance of the proposed geometric sliding-mode controller (GSMC) in contrast with two control schemes of the literature.

keywords:
Geometric control; Lie groups; Mechanical systems; Sliding subgroups.
thanks: This paper was not presented at any IFAC meeting. Corresponding author Yu Tang, on leave from the National Autonomous University of Mexico, Mexico City, MEXICO.

,

1 Introduction

Sliding mode control (SMC) (Utkin, 1977) has been proven to be a very powerful control design method for systems evolving in Euclidean spaces. Its design usually consists of two stages: the reaching stage where the controller drives the system trajectories to a sliding surface, a subspace embedded in the Euclidean space designed to convoy some specific characteristics (e.g., convergence time, actuator saturation) in accordance with the given control objectives, and a sliding stage where the system trajectories converge to the origin according to the reduced-order dynamics constrained in the sliding surface, achieving the control objectives. In the sliding stage, the reduced-order dynamics is independent of the system dynamics, and therefore, this control design method ensures its robustness against a certain class of disturbances and has achieved great success in a wide range of applications.

When this method is extended to mechanical systems whose configuration space is a general Lie group, care must be taken in the design of the sliding surface. Unlike the Euclidean case, when the system configuration space is a Lie group GG, its time rate of change belongs to the tangent space TgGT_{g}G at the configuration gg. Therefore, the state space is composed of the tangent bundle G×TgGG\times T_{g}G. The topological structure and the underlying properties of the configuration space and the tangent space are very different. Without taking this into account in the SMC design, the sliding surface may not belong to the tangent bundle, and therefore no guarantee is offered to ensure that the system trajectories reach the sliding surface and the sliding mode may not exist at all (Gómez et al., 2019). The main problem is thus how to devise a group operation such that the tangent bundle is a Lie group and that the sliding subgroup is immersed in the tangent bundle so that the salient features of SMC in the Euclidean space mentioned above may be inherited by a general Lie group.

We present in this paper a general method of designing a sliding mode control, a geometric sliding mode control (GSMC), for fully-actuated mechanical systems whose configuration space is a Lie group. A generic kinematic control is first devised in the underlying Lie algebra (the tangent space at the group identity with a bilinear map), which enables us to build a Lie group on the tangent bundle where the system state evolves. Then a sliding subgroup is proposed on the tangent bundle, and the sliding mode is guaranteed to exist. The sliding subgroup is designed to convoy control objectives, in particular, the almost global asymptotic convergence of the trajectories of the reduced-order dynamics to the identity of the tangent bundle is considered, which is the strongest convergence that may be achieved by continuous time-invariant feedback in a smooth Lie group (Bhat & Bernstein, 2000). The reaching control law is then designed to drive the trajectories to the sliding subgroup globally exponentially. Tracking is then composed of the reaching law and sliding mode, as in the Euclidean case.

1.1 Related work

The geometric approach to control designs has achieved significant advances for mechanical systems on nonlinear manifolds, for recent developments in this topic, see, for instance, Bullo & Lewis (2005) and the references therein. As recognized in Koditschek (1989); Bullo et al. (1995); Maithripala et al. (2006), a key point in control design is how to define the tracking error. The tracking error defined on a Riemannian manifold relying on an error function and a transport map in Bullo & Murray (1999) may be simplified if the manifold is endowed with a Lie group structure (Maithripala et al., 2006), where the error notion can be globally defined explicitly and is easier to be manipulated for stability analysis of the closed-loop system (Maithripala & Berg, 2015; Saccon et al., 2013; De Marco et al., 2018; Lee, 2012; Sarlette et al., 2010). A similar situation is encountered in observer designs using an estimation error defined on a Riemannian manifold (Aghannan & Rouchon, 2003) versus an estimation error defined by the group operation on Lie groups bonnabel2009non. The ability to define a global error on Lie groups provides a powerful tool for treating the error as an object in the state space globally and controlling it as a physical system so that the tracking problem can be reduced to stabilizing the error dynamics to the group identity (Bullo et al., 1995; Maithripala et al., 2006; Spong & Bullo, 2005). Moreover, a separation principle can be proved in the geometric approach to control designs (Maithripala et al., 2006; Maithripala & Berg, 2015) when part of the state in the control law is estimated by an exponentially convergent observer designed on the Lie group (Bonnabel et al., 2009), similar to an LTI system. This opens a wide field of applications for systems on Lie groups, such as rigid body motion control and trajectory tracking in 22D and 33D spaces, given the significant advances in both geometric control designs (Bullo & Lewis, 2005; Spong & Bullo, 2005; Lee, 2011; Akhtar & Waslander, 2020; Rodríguez-Cortés & Velasco-Villa, 2022) and observer designs (Aghannan & Rouchon, 2003; Bonnabel et al., 2009; Mahony et al., 2008; Lageman et al., 2009; Zlotnik & Forbes, 2018).

GSMC on Lie groups has been considered using two main approaches: developing the SMC in the underlying Lie algebra or developing it on the Lie group itself. The main idea in the former approach is first expressing the tracking error defined on the Lie group in its Lie algebra through the locally diffeomorphic logarithmic map (Bullo et al., 1995). Since the Lie algebra is a vector space, a sliding surface can be designed as in the Euclidean case (Culbertson et al., 2021; Liang et al., 2021; Espíndola & Tang, 2022). In the latter approach, the sliding subgroup is designed directly on the Lie group. Since the topological structures of the configuration space (a Lie group) and the tangent space (a vector space) are very different when the underlying Lie group is not diffeomorphic to an Euclidean space, an important question arises as to how to ensure the sliding surface to be indeed a subgroup of the state space formed by the tangent bundle to guarantee the existence of the sliding mode and thus to inherit the salient features of SMC in the Euclidean space.

SMC designs using the second approach have been reported for Lie groups such as SO(3)SO(3), 𝒮3\mathcal{S}^{3} for attitude control, and SE(3)SE(3) for motion controls (Ghasemi et al., 2020; Lopez & Slotine, 2021). However, the issue of whether the tangent bundle is a Lie group and whether the sliding subgroup is properly immersed on the tangent bundle was not addressed in these works. Therefore, the potential problem of lack of robustness due to the nonexistence of the sliding mode might appear. Recently, Gómez et al. (2019) brought this issue to the attention of the control community, and proposed an SMC on the rotation group SO(3)SO(3) with a sliding surface which was ensured to be a Lie subgroup immersed in the tangent bundle SO(3)×3SO(3)\times\mathbb{R}^{3}, and a finite-time convergent controller was devised for attitude control. This design method was applied in Meng et al. (2023) to design a second-order SMC for fault-tolerant control designs.

1.2 Contributions

We generalize the conventional sliding mode control designs for systems in Euclidean spaces to fully-actuated simple mechanical systems whose configuration space is a Lie group for the trajectory-tracking problem. The main contributions can be summarized as follows: (1) we endow the state space formed by the tangent bundle of the error dynamics with a Lie group structure by defining a group operation that is based on a generic kinematic control designed in the Lie algebra of the configuration Lie group; (2) we design a smooth sliding subgroup and show it to be a Lie subgroup of the tangent bundle, therefore, inheriting the Lie group structure of the state space; and (3) we design a coordinate-free geometric sliding mode controller for a fully-actuated mechanical system on a Lie group which drives the error dynamics to the sliding subgroup globally exponential at the reaching stage, the error dynamics then converges to the identity of the tangent bundle almost globally asymptotically at the sliding stage. In addition, rigid body tracking in 33D space is addressed on the special orthogonal groups SO(3)SO(3) and on the unit sphere 𝒮3\mathcal{S}^{3}, respectively, by applying the proposed geometric sliding mode control.

1.3 Organization

The rest of the paper is organized as follows. Section 2 presents the notation and background materials for simple mechanical systems with Lie groups as the configuration space. Section 3 first endows the state space formed by the tangent bundle with a Lie group structure under a group operation, which is defined based on a generic kinematic control law in the Lie algebra of the configuration space. Then, a smooth sliding subgroup is defined, which is a Lie subgroup immersed in the tangent bundle. The convergence to the identity of the tangent bundle of the reduced-order dynamics constrained on the sliding surface is analyzed based on Lyapunov stability. Section 4 gives the design of the GSMC, composed of a reaching law to the sliding subgroup and the convergence property of the sliding subgroup. Attitude tracking of a rigid body in 33D space is addressed in Section 5 respectively on the rotational group SO(3)SO(3) and the unit sphere 𝒮3\mathcal{S}^{3}, and simulation results under the GSMC developed on SO(3)SO(3) are presented in Section 6 for illustration and comparison. Conclusions are drawn in Section 7.

2 Mechanical systems on Lie groups

This section provides the notation and introduces the motion equations for a fully-actuated simple mechanical system on Lie groups. More details can be found in Bullo & Lewis (2005) and Abraham et al. (2012).

Given a finite-dimension Lie group GG, the identity of the group is denoted by eGe\in G. TeGT_{e}G denotes the tangent space in the identity, which also defines its Lie algebra 𝔤TeG\mathfrak{g}\triangleq T_{e}G in the Lie bracket [,]𝔤[\cdot,\cdot]\in\mathfrak{g}. Let Lg(h)=ghGL_{g}(h)=gh\in G and Rg(h)=hgGR_{g}(h)=hg\in G be the left and right translation maps, respectively, g,hG\forall g,h\in G, and denote its corresponding tangent maps TeLg(ν)=gνTgGT_{e}L_{g}(\nu)=g\cdot\nu\in T_{g}G and TeRg(ν)=νgTgGT_{e}R_{g}(\nu)=\nu\cdot g\in T_{g}G, νTeG\forall\nu\in T_{e}G, it describes the natural isomorphism TeGTgGT_{e}G\simeq T_{g}G, which induces the equivalence TGG×TeGTG\simeq G\times T_{e}G for the tangent bundle TG=G×TgGTG=G\times T_{g}G. The inverse tangent map from TgGT_{g}G to TeGT_{e}G is denoted by ν=g1vgL\nu=g^{-1}\cdot v^{L}_{g}, where vgL=νL(g)TgGv^{L}_{g}=\nu_{L}(g)\in T_{g}G, being νLΓ(TG)\nu_{L}\in\Gamma^{\infty}(TG) a left-invariant vector field, with Γ(TG)\Gamma^{\infty}(TG) denoting the set of CC^{\infty}-sections of TGTG, and respectively for a right-invariant vector field νRΓ(TG)\nu_{R}\in\Gamma^{\infty}(TG), it follows that vgR=νR(g)TgGv^{R}_{g}=\nu_{R}(g)\in T_{g}G, and accordingly ν=vgRg1\nu=v^{R}_{g}\cdot g^{-1}.

The cotangent space at gGg\in G is denoted by TgGT^{*}_{g}G, while 𝔤\mathfrak{g}^{*} describes the dual space of the Lie algebra 𝔤\mathfrak{g}. Likewise, the cotangent bundle is denoted by TGG×𝔤T^{*}G\simeq G\times\mathfrak{g}^{*}. Given a \mathbb{R}-vector space VV, its dual space VV^{*}, and a bilinear map B:V×VB\vcentcolon V\times V\to\mathbb{R}, the flat map B:VVB^{\flat}\vcentcolon V\to V^{*} is defined as B(v);u=B(u,v)\langle B^{\flat}(v);u\rangle=B(u,v), u,vV\forall u,v\in V, B(v)VB^{\flat}(v)\in V^{*}, where α;v=α(u)\langle\alpha;v\rangle=\alpha(u) denotes the image in \mathbb{R} of vVv\in V under the covector αV\alpha\in V^{*}. If the flat map is invertible, then the inverse, known as the sharp map, is denoted by B:VVB^{\sharp}\vcentcolon V^{*}\to V

The inner product on a smooth manifold \mathcal{M} is denoted by ,\langle\langle\cdot,\cdot\rangle\rangle\in\mathbb{R}. A Riemannian metric 𝔾\mathbb{G} on a Lie group GG assigns the inner product 𝔾(g)(Xg,Yg)\mathbb{G}(g)\cdot\left(X_{g},Y_{g}\right) on each TgGT_{g}G, Xg,YgTgG\forall X_{g},Y_{g}\in T_{g}G. Moreover, when 𝔾\mathbb{G} is left-invariant (resp. right-invariant), it induces an inner product in the Lie algebra 𝔤\mathfrak{g} by 𝕀(ξ,ζ)=𝔾(g)(ξL(g),ζL(g))\mathbb{I}\left(\xi,\zeta\right)=\mathbb{G}(g)\cdot\left(\xi_{L}(g),\zeta_{L}(g)\right), ξ,ζ𝔤\forall\xi,\zeta\in\mathfrak{g}. The kinetic energy is given by KE(vg)=(1/2)𝔾(g)(vg,vg)=(1/2)𝕀(ν,ν)\mathrm{KE}(v_{g})=(1/2)\mathbb{G}(g)\cdot\left(v_{g},v_{g}\right)=(1/2)\mathbb{I}(\nu,\nu), where 𝕀\mathbb{I} is the kinetic energy tensor, which induces a kinetic energy metric 𝔾\mathbb{G} on GG. In the rotational motion of a rigid body, 𝕀\mathbb{I} also represents the inertia tensor.

In the sequel, only the left invariance will be used. The proposed control methodology can be developed similarly for the right invariance. Also, subscripts and superscripts LL will be dropped when the meaning is clear. A left-invariant covariant derivative (affine connection) on a Lie group is denoted by ξLζLΓ(TG)\nabla_{\xi_{L}}\zeta_{L}\in\Gamma^{\infty}(TG) for any vector fields ξL,ζLΓ(TG)\xi_{L},\zeta_{L}\in\Gamma^{\infty}(TG). In addition, the Levi-Civita connection associated with the Riemannian metric 𝔾\mathbb{G} is denoted by 𝔾\overset{\mathbb{G}}{\nabla}, which is unique and torsion-free. A left-invariant affine connection on a Lie group is uniquely determined by a bilinear map B:𝔤×𝔤𝔤B:\mathfrak{g}\times\mathfrak{g}\to\mathfrak{g} called the restriction of the left-invariant connection. In particular, the restriction for the left-invariant Levi-Civita connection 𝔾\overset{\mathbb{G}}{\nabla} is defined as

𝔤ξζ12[ξ,ζ]12𝕀(adξ𝕀(ζ)+adζ𝕀(ξ)),\overset{\mathfrak{g}}{\nabla}_{\xi}\zeta\triangleq\frac{1}{2}\left[\xi,\zeta\right]-\frac{1}{2}\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\xi}\mathbb{I}^{\flat}(\zeta)+\mathrm{ad}^{*}_{\zeta}\mathbb{I}^{\flat}(\xi)\right), (1)

where the adjoint map ad:𝔤×𝔤𝔤\mathrm{ad}\vcentcolon\mathfrak{g}\times\mathfrak{g}\to\mathfrak{g} is defined as adξζ=[ξ,ζ]\mathrm{ad}_{\xi}\zeta=\left[\xi,\zeta\right], and adξ:𝔤𝔤\mathrm{ad}^{*}_{\xi}\vcentcolon\mathfrak{g}^{*}\to\mathfrak{g}^{*} is the dual map defined as adξα;ζ=α;[ξ,ζ]\langle\mathrm{ad}^{*}_{\xi}\alpha;\zeta\rangle=\langle\alpha;\left[\xi,\zeta\right]\rangle. Furthermore, the adjoint action Ad:G×𝔤𝔤\mathrm{Ad}\vcentcolon G\times\mathfrak{g}\to\mathfrak{g} is Adgζ=gζg1\mathrm{Ad}_{g}\zeta=g\cdot\zeta\cdot g^{-1}, gG\forall g\in G. So, the left-invariant Levi-Civita connection is explicitly expressed as

𝔾ξLζL(dζ(ξ)+𝔤ξζ)L,\overset{\mathbb{G}}{\nabla}_{\xi_{L}}\zeta_{L}\triangleq\left(\mathrm{d}\zeta(\xi)+\overset{\mathfrak{g}}{\nabla}_{\xi}\zeta\right)_{L}, (2)

where dζ(ξ)ddt|t=0ζ(gexp(ξt))\mathrm{d}\zeta(\xi)\triangleq\frac{d}{dt}|_{t=0}\;\zeta\left(g\;\mathrm{exp}(\xi t)\right), being exp:𝔤G\mathrm{exp}\vcentcolon\mathfrak{g}\to G the exponential map on GG, which is a local CC^{\infty}-diffeomorphism, and whose inverse is called the logarithmic map denoted by log:G𝔤\mathrm{log}\vcentcolon G\to\mathfrak{g}. By the left-invariance of vector fields ξL,ζLΓ(TG)\xi_{L},\zeta_{L}\in\Gamma^{\infty}(TG), the covariant derivative (2) is expressed in terms of ξ,ζ𝔤\xi,\zeta\in\mathfrak{g} as follows

ξζdζ(ξ)+𝔤ξζ.\nabla_{\xi}\zeta\triangleq\mathrm{d}\zeta(\xi)+\overset{\mathfrak{g}}{\nabla}_{\xi}\zeta.

Consider a differentiable curve g:IGg:I\to G, where II is the set of all intervals. Then a body velocity ν:I𝔤\nu:I\to\mathfrak{g} is defined as tTg(t)Lg1(t)(g˙(t))t\mapsto T_{g(t)}L_{g^{-1}(t)}\left(\dot{g}(t)\right), for all tIt\in I, and therefore

g˙(t)=g(t)ν(t).\dot{g}(t)=g(t)\cdot\nu(t). (3)

A forced mechanical system is governed by the intrinsic Euler-Lagrange equations

𝔾g˙(t)g˙(t)=Fu+Δd,\overset{\mathbb{G}}{\nabla}_{\dot{g}(t)}\dot{g}(t)=F_{u}+\Delta_{d}, (4)

where Fu=a=1mua(t)𝔾(Tg(t)Lg1(t)(fa))F_{u}=\sum^{m}_{a=1}u^{a}(t)\mathbb{G}^{\sharp}\left(T^{*}_{g(t)}L_{g^{-1}(t)}\left(f^{a}\right)\right) is the control force applied to the system on TgGT_{g}G, being ua:Iu^{a}:I\to\mathbb{R} the control inputs, and fa(g)𝔤f^{a}(g)\in\mathfrak{g}^{*} the control forces. Furthermore, ΔdTgG\Delta_{d}\in T_{g}G represents the vector field version of constraint forces, such as potential external forces, uncontrolled conservative plus dissipative forces, and unmodeled disturbances.

In view of (3) and the left-invariance of g˙\dot{g}, the Levi-Civita connection in (4) can be explicitly expressed using (1)-(2) as

𝔾g˙(t)g˙(t)=g(t)(ν˙(t)+𝔤ν(t)ν(t)),\overset{\mathbb{G}}{\nabla}_{\dot{g}(t)}\dot{g}(t)=g(t)\cdot\left(\dot{\nu}(t)+\overset{\mathfrak{g}}{\nabla}_{\nu(t)}\nu(t)\right),

resulting in the controlled Euler-Poincaré equation

ν˙(t)+𝔤ν(t)ν(t)=fu+δd,\dot{\nu}(t)+\overset{\mathfrak{g}}{\nabla}_{\nu(t)}\nu(t)=f_{u}+\delta_{d}, (5)

with fu=a=1mua(t)𝕀(fa)𝔤f_{u}=\sum^{m}_{a=1}u^{a}(t)\mathbb{I}^{\sharp}\left(f^{a}\right)\in\mathfrak{g}, and δd=g1Δd𝔤\delta_{d}=g^{-1}\cdot\Delta_{d}\in\mathfrak{g}.

The underlying mechanical system on the Lie group GG is then defined by the configuration Lie group GG, the inertia tensor 𝕀\mathbb{I}, and the external forces fu+δdf_{u}+\delta_{d}.

3 Lie Group Structure of the State Space and the Sliding Subgroup

In this section, we will endow the tangent bundle TGG×𝔤TG\simeq G\times\mathfrak{g} with a Lie group structure by a properly designed group operation. For this purpose, an intrinsic control for kinematics is first proposed (3). Then we design a smooth sliding subgroup that is immersed in the tangent bundle so that it inherits the Lie group structure of the state space.

3.1 Intrinsic kinematic control

The purpose of this subsection is to design a control law ν(t)𝔤\nu(t)\in\mathfrak{g} for the kinematics (3) to render g(t)eg(t)\to e, the group identity. Let V:G0V:G\to\mathbb{R}_{\geq 0} be an infinitely differentiable proper Morse function, which satisfies V(g)>0V(g)>0, gG\{e}\forall g\in G\backslash\{e\}, dV(g)=0\mathrm{d}V(g)=0 and V(g)=0g=eV(g)=0\iff g=e. Morse functions, a class of error functions (Koditschek, 1989; Bullo et al., 1995), are guaranteed to exist on many Lie groups of practical interest considered in this paper (Maithripala & Berg, 2015; Bullo & Lewis, 2005). They represent potential energy that can be used to measure the distance between the configuration gg and the identity ee on GG. The following definition specifies the class of kinematic controls considered in the paper.

Definition 3.1 (Kinematic control law).

Let g:IGg:I\to G be a differentiable curve governed by (3), for all tIt\in I. A kinematic control law is a map νu:G𝔤\nu_{u}:G\to\mathfrak{g} that satisfies the following properties.

  1. i.

    νu(e)=0\nu_{u}(e)=0,

  2. ii.

    νu(g1)=νu(g)\nu_{u}\left(g^{-1}\right)=-\nu_{u}\left(g\right),

  3. iii.

    dV(g(t));g(t)νu(g(t))<0\langle\mathrm{d}V(g(t));-g(t)\cdot\nu_{u}(g(t))\rangle<0, g(t)G\𝒪u\forall g(t)\in G\backslash\mathcal{O}_{u}, where 𝒪u{gG\{e}|gνu(g)=0}\mathcal{O}_{u}\triangleq\left\{g\in G\backslash\{e\}\;|\;g\cdot\nu_{u}(g)=0\right\},

  4. iv.

    dV(g(t));g(t)νu(g(t))=y(g(t))V(g(t))\langle\mathrm{d}V(g(t));-g(t)\cdot\nu_{u}(g(t))\rangle=-y\left(g(t)\right)V\left(g(t)\right), g(t)𝒰\forall g(t)\in\mathcal{U}, where y:𝒰>0y:\mathcal{U}\to\mathbb{R}_{>0}, and 𝒰G\𝒪u\mathcal{U}\subset G\backslash\mathcal{O}_{u} is a neighborhood of ee.

Some comments on the class of kinematic controls are in order. Properties (i)-(ii) are instrumental to building a particular Lie subgroup on the tangent bundle. Properties (iii)-(iv) represent the sliding (convergence) property of the reduced-order dynamics on the sliding subgroup (Lemma 5 below). In particular, Property (iii) states the almost-global asymptotic stability for system (3) in closed loop with the kinematic control law ν(t)=νu(g(t))\nu(t)=-\nu_{u}(g(t)). Note that since 𝒪u\mathcal{O}_{u} is the set of closed-loop equilibria other than g(t)=eg(t)=e, they are critical points of V(g)V(g). Since V(g)V(g) is a Morse function, the set 𝒪u\mathcal{O}_{u} consists of a finite number of isolated points. In addition, this set is nowhere dense, which means that it cannot separate the configuration space. Therefore, the complement G\𝒪uG\backslash\mathcal{O}_{u} is open and dense, i.e., G\𝒪uG\backslash\mathcal{O}_{u} is a submanifold of GG (Maithripala et al., 2006). Finally, Property (iv) establishes the local exponential stability of the closed-loop system, where the existence of the neighborhood 𝒰\mathcal{U} is immediate because V(g)V(g) is a Morse function, which has a unique minimum at eGe\in G by definition.

Note that both V(g)V(g) and νu(g)\nu_{u}(g) are of free design, provided that the properties in Definition 3.1 hold. However, it is worth considering the kinematic control law in the logarithmic coordinate, that is, νu(g)=log(g)\nu_{u}(g)=\mathrm{log}(g), or some parallel vectors to log(g)\mathrm{log}(g) (Akhtar & Waslander, 2020), as this map has been found to provide the strongest stability results, for example, almost global and local exponential convergence to the identity through a geodesic path (Bullo et al., 1995).

3.2 Lie Group structure for the state space

For systems described in (3) and (5), the state space is the tangent bundle TGG×𝔤TG\simeq G\times\mathfrak{g}. To endow it with a Lie group structure, we consider the binary operation :TG×TGTG\star:\ TG\times TG\mapsto TG defined in the following

h1h2\displaystyle h_{1}\star h_{2}\triangleq
(g1g2,ν1+ν2+λνu(g1)+λνu(g2)λνu(g1g2)),\displaystyle\big{(}g_{1}g_{2},\;\nu_{1}+\nu_{2}+\lambda\nu_{u}(g_{1})+\lambda\nu_{u}(g_{2})-\lambda\nu_{u}(g_{1}g_{2})\big{)}, (6)

h1=(g1,ν1),h2=(g2,ν2)TG\forall h_{1}=(g_{1},\nu_{1}),\;h_{2}=(g_{2},\nu_{2})\in TG, and λ>0\lambda\in\mathbb{R}_{>0}.

Lemma 1 (The state space TGTG as a Lie group).

The tangent bundle TGG×𝔤TG\equiv G\times\mathfrak{g} endowed with the binary operation (3.2) is a Lie group, with

  1. i.

    Identity element: f(e,0)TGf\triangleq(e,0)\in TG,

  2. ii.

    Inverse element: h1(g1,ν)TGh^{-1}\triangleq\left(g^{-1},-\nu\right)\in TG, h=(g,ν)TG\forall h=(g,\nu)\in TG.

{pf}

Being TGTG a smooth manifold with (3.2) a smooth operation, it only remains to verify the group axioms as follows.

  1. 1.

    h=(g,ν)TG\forall h=(g,\nu)\in TG, it satisfies

    hf\displaystyle h\star f =(ge,ν+0+λνu(g)+λνu(e)λνu(ge))\displaystyle=\left(ge,\;\nu+0+\lambda\nu_{u}(g)+\lambda\nu_{u}(e)-\lambda\nu_{u}(ge)\right)
    =fh=h,\displaystyle=f\star h=h,

    where Property of Definition 3.1(i) is used.

  2. 2.

    The group operation between h=(g,ν)TGh=(g,\nu)\in TG and its inverse h1=(g1,ν)TGh^{-1}=\left(g^{-1},-\nu\right)\in TG verifies

    h1h\displaystyle h^{-1}\star h
    =(g1g,ν+ν+λνu(g1)+λνu(g)λνu(g1g))\displaystyle=\big{(}g^{-1}g,-\nu+\nu+\lambda\nu_{u}(g^{-1})+\lambda\nu_{u}(g)-\lambda\nu_{u}(g^{-1}g)\big{)}
    =(gg1,νν+λνu(g)+λνu(g1)λνu(gg1))\displaystyle=\big{(}gg^{-1},\nu-\nu+\lambda\nu_{u}(g)+\lambda\nu_{u}(g^{-1})-\lambda\nu_{u}(gg^{-1})\big{)}
    =hh1=f,\displaystyle=h\star h^{-1}=f,

    where Properties (i)-(ii) of Definition 3.1 are used.

  3. 3.

    The associativity h1(h2h3)=(h1h2)h3h_{1}\star\left(h_{2}\star h_{3}\right)=\left(h_{1}\star h_{2}\right)\star h_{3} is proved straightforwardly by substitution, using the properties of Definition 3.1.

\blacksquare

Remark 2 (Tangent bundle TGTG).

The definition of the group operation (3.2) relying on the kinematic control νu(g)\nu_{u}(g) in Definition 3.1 is crucial to define a sliding Lie subgroup immersed in TGTG in the next subsection. In fact, a group operation to endow TGTG to be a Lie group may simply be h1h2=(g1g2,ν1+ν2)h_{1}\star h_{2}=\left(g_{1}g_{2},\;\nu_{1}+\nu_{2}\right). However, this operation does not allow to design of a useful sliding subgroup, in particular, it fails to prove closure under the group operation, as will be seen below.

Remark 3 (Associativity).

The associativity proved in Lemma 1 ensures the proposed Lie group TGTG to be globalizable (Olver et al., 1996), that is, the local Lie group TGTG can be extended to be a global topological group. This fact allows us to develop a sliding mode control defined globally on the state space in contrast to the Lie groups defined locally in Gómez et al. (2019) and Meng et al. (2023).

3.3 Sliding Subgroup on TGTG

In this subsection, we define a smooth sliding subgroup on the tangent bundle. The following lemma shows that HTGH\subset TG is an immersed submanifold of TGTG that inherits the topology and smooth structure of the tangent bundle TGTG (Lee, 2013).

Lemma 4 (Sliding Lie subgroup).

Define

H{h=(g,ν)TG|s(h)=0}TG,H\triangleq\left\{h=(g,\nu)\in TG\;|\;s(h)=0\right\}\subset TG, (7)

where h=(g,ν)TG\forall h=(g,\nu)\in TG, the map s:TG𝔤s:TG\mapsto\mathfrak{g} is defined as

s(h)=ν+λνu(g).s(h)=\nu+\lambda\nu_{u}(g). (8)

Then HTGH\subset TG is a Lie subgroup under the group operation (3.2).

{pf}

The smoothness of HH is immediate, because the map defined in (8) is smooth. The proof consists thus in showing that subset HH inherits the group structure of the Lie group TGTG, by verifying the following:

  1. i.

    Identity: The identity of the tangent bundle f=(e,0)Hf=(e,0)\in H. This is immediate by Definition 3.1(i) since s(f)=0+λνu(e)=0s(f)=0+\lambda\nu_{u}(e)=0.

  2. ii.

    Inverse. h=(g,ν)H\forall h=(g,\nu)\in H, s(h)=0ν=λνu(g)s(h)=0\implies\nu=-\lambda\nu_{u}(g). By Definition 3.1(ii) it follows that

    s(h1)\displaystyle s\left(h^{-1}\right) =ν+λνu(g1)\displaystyle=-\nu+\lambda\nu_{u}\left(g^{-1}\right)
    =(λνu(g))+λνu(g1)=0.\displaystyle=-\left(-\lambda\nu_{u}(g)\right)+\lambda\nu_{u}\left(g^{-1}\right)=0.

    This proves that h1Hh^{-1}\in H for all hHh\in H.

  3. iii.

    Closure. Given h1=(g1,ν1),h2=(g2,ν2)Hh_{1}=(g_{1},\nu_{1}),\ h_{2}=(g_{2},\nu_{2})\in H, then s(h1)=0ν1=λνu(g1)s(h_{1})=0\implies\nu_{1}=-\lambda\nu_{u}(g_{1}) and s(h2)=0ν2=λνu(g2)s(h_{2})=0\implies\nu_{2}=-\lambda\nu_{u}(g_{2}). By (3.2) h1h2=(g1g2,λνu(g1g2))h_{1}\star h_{2}=\big{(}g_{1}g_{2},-\lambda\nu_{u}(g_{1}g_{2})\big{)}. Thus, s(h1h2)=λνu(g1g2)+λνu(g1g2)=0s\left(h_{1}\star h_{2}\right)=-\lambda\nu_{u}(g_{1}g_{2})+\lambda\nu_{u}(g_{1}g_{2})=0. That is, HH is closed under the group operation.

\blacksquare

The following lemma shows that once a trajectory reaches the sliding subgroup it will stay on it and converges to the group identity.

Lemma 5 (Properties of the sliding subgroup HH).

Consider the sliding Lie subgroup HTGH\subset TG in (7). Then HH is forward invariant, i.e., h(tr)Hh(t_{r})\in H for some trIt_{r}\in I h(t)H,ttr\implies h(t)\in H,\forall t\geq t_{r}. Moreover, h(t)(e, 0)h(t)\to(e,\ 0) almost globally asymptotically.

{pf}

Consider a differentiable curve g:IGg:I\to G of the dynamics (3). Let V:GV:G\to\mathbb{R} be a proper Morse function with the unique minimum at eGe\in G. Then, along the trajectory g(t)g(t) and tI\forall t\in I, it yields

ddtV(g(t))=dV(g(t));g˙(t)=dV(g(t));g(t)ν(t).\frac{\mathrm{d}}{\mathrm{d}t}V\left(g(t)\right)=\langle\mathrm{d}V\left(g(t)\right);\dot{g}(t)\rangle=\langle\mathrm{d}V\left(g(t)\right);g(t)\cdot\nu(t)\rangle.

Assume that h(tr)Hh(t_{r})\in H, for some trIt_{r}\in I. Then s(h)=0s(h)=0 gives ν(t)=λνu(g(t))\nu(t)=-\lambda\nu_{u}(g(t)). Therefore,

ddtV(g(t))\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}V\left(g(t)\right) =dV(g(t));g(t)ν(t)\displaystyle=\langle\mathrm{d}V\left(g(t)\right);g(t)\cdot\nu(t)\rangle
=dV(g(t));λg(t)νu(g(t)),\displaystyle=\langle\mathrm{d}V\left(g(t)\right);-\lambda g(t)\cdot\nu_{u}\left(g(t)\right)\rangle,

In light of Definition 3.1(iii), it follows that ddtV(g(t))<0\frac{\mathrm{d}}{\mathrm{d}t}V\left(g(t)\right)<0, for all g(tr)G\𝒪ug(t_{r})\in G\backslash\mathcal{O}_{u}, and ddtV(g(t))=0g=e\frac{\mathrm{d}}{\mathrm{d}t}V\left(g(t)\right)=0\iff g=e, where 𝒪u\mathcal{O}_{u} is a nowhere-dense set with a finite number of points given in Definition 3.1(iii). Therefore, h(t)h(t) will remain on HH for all ttrt\geq t_{r}, and the equilibrium g(t)=eg(t)=e of (3) is almost globally asymptotically stable for all g(tr)G\𝒪ug(t_{r})\in G\backslash\mathcal{O}_{u} and locally exponentially stable g(0)𝒰\forall g(0)\in\mathcal{U}, according to Definition 3.1(iv).

\blacksquare

4 Geometric Sliding Mode Control (GSMC)

In this section, we design a control law, called the reaching law, for fuf_{u} in the Euler-Poincaré equation (5) to drive the trajectory h(t)=(g(t),ν(t))TGh(t)=(g(t),\nu(t))\in TG to the sliding subgroup HH. Then the tracking control objective will be achieved as a consequence of Lemma 5.

4.1 Reaching Law

The Euler-Lagrange dynamics (4), ignoring disturbance forces Δd\Delta_{d}, is expressed as

ν(t)ν(t)=fu.\nabla_{\nu(t)}\nu(t)=f_{u}. (9)

which is defined on TGTG, being h(t)=(g(t),ν(t))h(t)=\left(g(t),\nu(t)\right) the state variable.

The intrinsic acceleration for the sliding variable (8) is calculated, by using (9), through the covariant derivative of s(h)𝔤s(h)\in\mathfrak{g} with respect to itself as

s(h)s(h)\displaystyle\nabla_{s(h)}s(h) =ddts(h)+𝔤s(h)s(h)\displaystyle=\frac{d}{dt}s(h)+\overset{\mathfrak{g}}{\nabla}_{s(h)}s(h)
=ν˙+λν˙u(g)+𝔤s(h)s(h).\displaystyle=\dot{\nu}+\lambda\dot{\nu}_{u}(g)+\overset{\mathfrak{g}}{\nabla}_{s(h)}s(h).

Substituting the Euler-Poincaré equation (5) yields

s(h)s(h)\displaystyle\nabla_{s(h)}s(h)
=\displaystyle= 𝔤ν(t)ν(t)+λν˙u(g)+𝔤s(h)s(h)+fu\displaystyle-\overset{\mathfrak{g}}{\nabla}_{\nu(t)}\nu(t)+\lambda\dot{\nu}_{u}(g)+\overset{\mathfrak{g}}{\nabla}_{s(h)}s(h)+f_{u}
=\displaystyle= 𝕀(adν(t)𝕀(ν(t)))𝕀(ads(h)𝕀(s(h)))+λν˙u(g)+fu,\displaystyle\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\nu(t)}\mathbb{I}^{\flat}(\nu(t))\right)-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(s(h))\right)+\lambda\dot{\nu}_{u}(g)+f_{u}, (10)

where the skew-symmetry of the Lie bracket [,]𝔤[\cdot,\cdot]\in\mathfrak{g} in (1) is used. The reaching law is then proposed as follows

fu=𝕀(adλνu(g)𝕀(ν(t)))λν˙u(g)kss(h),f_{u}=\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\lambda\nu_{u}(g)}\mathbb{I}^{\flat}(\nu(t))\right)-\lambda\dot{\nu}_{u}(g)-k_{s}s(h), (11)

with ks>0k_{s}>0 a design parameter.

Theorem 6 (Reaching Controller).

The reaching law (11) drives exponentially the trajectories of the closed-loop system (4.1) to the sliding subgroup HH h(0)TG\forall h(0)\in TG, i.e., s(h(t))0s(h(t))\to 0 globally exponentially.

{pf}

Consider the function W:TGW:TG\to\mathbb{R} defined below

W(h)=12𝕀(s(h),s(h)).W(h)=\frac{1}{2}\mathbb{I}(s(h),s(h)). (12)

Its time evolution along trajectories of (4.1) is given by

W˙(h)\displaystyle\dot{W}(h) =𝕀(s(h)s(h),s(h))\displaystyle=\mathbb{I}\left(\nabla_{s(h)}s(h),\;s(h)\right)
=𝕀(𝕀(adν(t)𝕀(ν(t)))𝕀(ads(h)𝕀(s(h)))\displaystyle=\mathbb{I}\left(\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\nu(t)}\mathbb{I}^{\flat}(\nu(t))\right)-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(s(h))\right)\right.
+λν˙u(g)+fu,s(h)),\displaystyle\quad\quad+\lambda\dot{\nu}_{u}(g)+f_{u},\;s(h)\Bigl{)},

which in closed loop with the controller (11) yields

W˙(h)\displaystyle\dot{W}(h) =𝕀(𝕀(adν(t)𝕀(ν(t)))𝕀(ads(h)𝕀(s(h)))\displaystyle=\mathbb{I}\left(\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\nu(t)}\mathbb{I}^{\flat}(\nu(t))\right)-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(s(h))\right)\right.
+𝕀(adλνu(g)𝕀(ν(t)))kss(h),s(h)),\displaystyle\quad\quad+\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\lambda\nu_{u}(g)}\mathbb{I}^{\flat}(\nu(t))\right)-k_{s}s(h),\;s(h)\Bigl{)},
=𝕀(𝕀(ads(h)𝕀(ν(t)))𝕀(ads(h)𝕀(s(h)))\displaystyle=\mathbb{I}\left(\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(\nu(t))\right)-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(s(h))\right)\right.
kss(h),s(h)).\displaystyle\quad\quad-k_{s}s(h),\;s(h)\Bigl{)}.

By Lemma 12 in Appendix A the term 𝕀(𝕀(adζ𝕀(η)),ζ)=0\mathbb{I}\left(\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\zeta}\mathbb{I}^{\flat}(\eta)\right),\;\zeta\right)=0, for any ζ,η𝔤\zeta,\eta\in\mathfrak{g}. Therefore,

W˙(h)=ks𝕀(s(h),s(h))=2ksW(h).\dot{W}(h)=-k_{s}\mathbb{I}\left(s(h),\;s(h)\right)=-2k_{s}W(h).

It follows from Proposition 6.26 of Bullo & Lewis (2005) that W(h(t))0W(h(t))\to 0 exponentially. \blacksquare

Remark 7 (Passivity of the Lagrangian dynamics).

Note that the first two right-hand terms of the control law (11) complete the terms 𝕀ads(h)𝕀(ν(t))𝕀ads(h)𝕀(s(h))\mathbb{I}^{\sharp}\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(\nu(t))-\mathbb{I}^{\sharp}\mathrm{ad}^{*}_{s(h)}\mathbb{I}^{\flat}(s(h)). By exploring the intrinsic passivity properties in Lemma 12 in Appendix A, these terms were not canceled in the above stability analysis. This result was first given for the Lie group SO(3)SO(3) in Koditschek (1989). The lemma 12 extends this result to coordinate-free Lagrangian dynamics on a general Lie group, which has not been explored, to the authors’ knowledge, in the literature for stability analysis.

Remark 8 (The reaching controller ).

The reaching law (11) achieves the convergence of s(h(t))0s(h(t))\to 0 for the Euler-Lagrange dynamics (9), which implies that h(t)TGh(t)\in TG reaches the sliding subgroup HH exponentially. Note that the result of Theorem 6 holds when the external constraint forces δd\delta_{d} can be compensated for by the controller fuf_{u}, which was omitted from the control design. Otherwise, in the presence of bounded δd\delta_{d}, h(t)TGh(t)\in TG will remain bounded and close to HH.

4.2 Tracking Control

Let gr:IGg_{r}:I\to G be a twice differentiable configuration reference, with the corresponding reference body velocity νr:I𝔤\nu_{r}:I\to\mathfrak{g} given by νr(t)gr1(t)g˙r(t)\nu_{r}(t)\triangleq g^{-1}_{r}(t)\cdot\dot{g}_{r}(t). The problem is to design a control law fuf_{u} to track the reference. The Lie group structure of the configuration space GG enables to define the following intrinsic configuration error

ge(t)gr1(t)g(t).g_{e}(t)\triangleq g^{-1}_{r}(t)g(t).

By left invariance the body velocity error is defined as

νe(t)\displaystyle\nu_{e}(t) ge1(t)g˙e(t)=ν(t)ηr(t),\displaystyle\triangleq g^{-1}_{e}(t)\cdot\dot{g}_{e}(t)=\nu(t)-\eta_{r}(t), (13)

with ηr(t)=Adge1νr(t)\eta_{r}(t)=\mathrm{Ad}_{g^{-1}_{e}}\nu_{r}(t). Then, the error dynamics evolving on TGTG is described by

νe(t)νe(t)=fu,\nabla_{\nu_{e}(t)}\nu_{e}(t)=f_{u}, (14)

being the state variable he(t)=(ge(t),νe(t))TGh_{e}(t)=\left(g_{e}(t),\nu_{e}(t)\right)\in TG.

The tracking problem, therefore, boils down to stabilizing the identity f=(e,0)f=(e,0) on TGTG. By using the sliding-model control strategy, the error state is first driven to the sliding subgroup in the reaching stage, and then on the sliding subgroup, the reduced-order dynamics converges to the identity ff ensured by Lemma 5.

In terms of the error state heh_{e} the sliding variable (8) is given by

s(he)=νe(t)+λνu(ge),s(h_{e})=\nu_{e}(t)+\lambda\nu_{u}(g_{e}), (15)

and, its covariant derivative, by using (1)-(2), is

s(he)s(he)\displaystyle\nabla_{s(h_{e})}s(h_{e})
=ddts(he)+𝔤s(he)s(he)\displaystyle=\frac{d}{dt}s(h_{e})+\overset{\mathfrak{g}}{\nabla}_{s(h_{e})}s(h_{e})
=ν˙e(t)+λν˙u(ge)𝕀(ads(he)𝕀(s(he)))\displaystyle=\dot{\nu}_{e}(t)+\lambda\dot{\nu}_{u}(g_{e})-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h_{e})}\mathbb{I}^{\flat}\left(s(h_{e})\right)\right)
=ν˙(t)η˙r(t)+λν˙u(ge)𝕀(ads(he)𝕀(s(he))).\displaystyle=\dot{\nu}(t)-\dot{\eta}_{r}(t)+\lambda\dot{\nu}_{u}(g_{e})-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h_{e})}\mathbb{I}^{\flat}\left(s(h_{e})\right)\right).

Ignoring disturbance δd\delta_{d} it follows from the Euler-Poincaré equation (5) that

s(he)s(he)=\displaystyle\nabla_{s(h_{e})}s(h_{e})= 𝕀(adν(t)𝕀(ν(t)))+fuη˙r(t)\displaystyle\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\nu(t)}\mathbb{I}^{\flat}\left(\nu(t)\right)\right)+f_{u}-\dot{\eta}_{r}(t) (16)
+λν˙u(ge)𝕀(ads(he)𝕀(s(he))).\displaystyle+\lambda\dot{\nu}_{u}(g_{e})-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h_{e})}\mathbb{I}^{\flat}\left(s(h_{e})\right)\right).

We proposed the following tracking controller

fu\displaystyle f_{u} =𝕀(adλνu(ge)ηr(t)𝕀(ν(t)))λν˙u(ge)+η˙r(t)\displaystyle=\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\lambda\nu_{u}(g_{e})-\eta_{r}(t)}\mathbb{I}^{\flat}(\nu(t))\right)-\lambda\dot{\nu}_{u}(g_{e})+\dot{\eta}_{r}(t)
kss(he),\displaystyle\quad-k_{s}s(h_{e}), (17)

where ks>0k_{s}>0 is a design parameter. The following theorem establishes the stability of the equilibrium he=fh_{e}=f in the closed-loop system (16)-(4.2).

Theorem 9 (Tracking Controller).

Consider the error dynamics (16) in closed loop with the controller (4.2). Then, the equilibrium he(t)=fh_{e}(t)=f is

  1. i.

    almost-globally asymptotically stable, for all he(0)TG¯G\𝒪u×𝔤h_{e}(0)\in\overline{TG}\triangleq G\backslash\mathcal{O}_{u}\times\mathfrak{g},

  2. ii.

    locally exponentially stable for all he(0)TU¯𝒰×𝔤h_{e}(0)\in\overline{TU}\triangleq\mathcal{U}\times\mathfrak{g}, where 𝒪u\mathcal{O}_{u} and 𝒰\mathcal{U} are given in Definition 3.1(iii)-(iv).

{pf}

Substituting the controller (4.2) in the error dynamics (16) yields the closed-loop dynamics

s(he)s(he)\displaystyle\nabla_{s(h_{e})}s(h_{e}) =𝕀(ads(he)𝕀(ν(t)))kss(he)\displaystyle=\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h_{e})}\mathbb{I}^{\flat}\left(\nu(t)\right)\right)-k_{s}s(h_{e})
𝕀(ads(he)𝕀(s(he))),\displaystyle\quad-\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{s(h_{e})}\mathbb{I}^{\flat}\left(s(h_{e})\right)\right),

which has an equilibrium point at s(he)=0s(h_{e})=0. The results follow as a consequence of Theorem 6 (the reaching stage) and Lemma 5 (the sliding mode). \blacksquare

Remark 10 (The tracking controller).

Theorem 9 gives a coordinate-free sliding mode control for a mechanical system whose configuration space is a general Lie group. The group structure allows defining globally a tracking error, whose dynamics evolves on the tangent bundle. The Lie subgroup of the sliding subgroup immersed on the tangent bundle ensures the existence of the sliding mode and thus inherits the salient features of the SMC in Euclidean spaces.

Similarly to the Euclidean case, the design of the sliding subgroup and the reaching law may incorporate other control objectives, such as finite-time convergence and controller saturation, which are, however, beyond the scope of the main purposes of this paper.

5 Attitude Tracking of a Rigid Body

In this section, we present the attitude tracking of a rigid body in the 33D space using the proposed GSMC. To illustrate the theoretic development, the problem is addressed using attitude representation by first the rotation matrix on SO(3)SO(3) and then by the unit quaternion 𝒮3\mathcal{S}^{3}.

5.1 GSMC for Attitude Tracking on SO(3)SO(3)

The group of rotations on 3\mathbb{R}^{3} is the Lie group SO(3)={R3×3|RRT=RTR=I3,det(R)=+1}SO(3)=\left\{R\in\mathbb{R}^{3\times 3}\;|\;RR^{T}=R^{T}R=I_{3},\;\mathrm{det}(R)=+1\right\}, with the usual multiplication of matrices as the group operation. The identity of the group is the identity matrix I3I_{3} of 3×33\times 3, and the inverse is the transpose RTSO(3)R^{T}\in SO(3) for any RSO(3)R\in SO(3). The Lie algebra is given by the set of skew-symmetric matrices 𝔰𝔬(3)={S3×3|ST=S}\mathfrak{so}(3)=\left\{S\in\mathbb{R}^{3\times 3}\;|\;S^{T}=-S\right\}, which is isomorphic to 3\mathbb{R}^{3}, i.e., 𝔰𝔬(3)3\mathfrak{so}(3)\simeq\mathbb{R}^{3}. The Lie bracket in 3\mathbb{R}^{3} is defined by the cross product [ζ,η]=adζηζ×η[\zeta,\eta]=\mathrm{ad}_{\zeta}\eta\triangleq\zeta\times\eta, ζ,η3\forall\zeta,\eta\in\mathbb{R}^{3}. Denote the isomorphism :3𝔰𝔬(3)\cdot^{\wedge}\vcentcolon\mathbb{R}^{3}\to\mathfrak{so}(3), and respectively the inverse map :𝔰𝔬(3)3\cdot^{\vee}\vcentcolon\mathfrak{so}(3)\to\mathbb{R}^{3}. Then for a differentiable curve R:ISO(3)R\vcentcolon I\to SO(3) with left-invariant dynamics R˙(t)TRSO(3)\dot{R}(t)\in T_{R}SO(3), the body angular velocity is given by

Ω(t)=RT(t)R˙(t)=[0Ω3(t)Ω2(t)Ω3(t)0Ω1(t)Ω2(t)Ω1(t)0],\Omega^{\wedge}(t)=R^{T}(t)\dot{R}(t)=\left[\begin{array}[]{ccc}0&-\Omega_{3}(t)&\Omega_{2}(t)\\ \Omega_{3}(t)&0&-\Omega_{1}(t)\\ -\Omega_{2}(t)&\Omega_{1}(t)&0\end{array}\right],

for all tIt\in I. The kinetic energy of the rotational motion of a rigid body is calculated as KE(Ω)=12𝕁(Ω,Ω)12𝕁Ω,Ω\mathrm{KE}(\Omega)=\frac{1}{2}\mathbb{J}(\Omega,\Omega)\triangleq\frac{1}{2}\langle\langle\mathbb{J}\Omega,\Omega\rangle\rangle, where 𝕁=𝕁T3×3\mathbb{J}=\mathbb{J}^{T}\in\mathbb{R}^{3\times 3} is the positive-definite inertia tensor. Therefore, adζ𝕁(η)=(𝕁η)ζ\mathrm{ad}^{*}_{\zeta}\mathbb{J}^{\flat}(\eta)=\left(\mathbb{J}\eta\right)^{\wedge}\zeta, and 𝕁(ζ)=𝕁1ζ\mathbb{J}^{\sharp}(\zeta)=\mathbb{J}^{-1}\zeta. Hence, the rotational motion described by the Euler-Lagrange equation (4) is

Ω(t)Ω(t)=τu.\nabla_{\Omega(t)}\Omega(t)=\tau_{u}. (18)

The state (R,ω)(R,\omega) evolves on the tangent bundle TSO(3)SO(3)×3TSO(3)\simeq SO(3)\times\mathbb{R}^{3}, and the control torque τu=𝕁1τ3\tau_{u}=\mathbb{J}^{-1}\tau\in\mathbb{R}^{3} is expressed in the body frame. Furthermore, (18) is explicitly expressed, by using the Euler-Poincaré equation (5) and restriction (1), as

Ω˙(t)𝕁1(𝕁Ω(t))Ω(t)=τu.\dot{\Omega}(t)-\mathbb{J}^{-1}\left(\mathbb{J}\Omega(t)\right)^{\wedge}\Omega(t)=\tau_{u}. (19)

Let Rr:ISO(3)R_{r}\vcentcolon I\to SO(3) be a twice differentiable attitude reference, and Ωr:I3\Omega_{r}:I\to\mathbb{R}^{3}, the reference angular velocity expressed in the body frame, which holds Ωr(t)=(RrT(t)R˙r(t))\Omega_{r}(t)=\left(R^{T}_{r}(t)\dot{R}_{r}(t)\right)^{\vee}. Then, the intrinsic attitude error is

Re(t)RrT(t)R(t).R_{e}(t)\triangleq R^{T}_{r}(t)R(t).

In view of (13) the (left-invariant) velocity error is

Ωe(t)\displaystyle\Omega_{e}(t) (ReT(t)R˙e(t))=Ω(t)σ(t),\displaystyle\triangleq\left(R^{T}_{e}(t)\dot{R}_{e}(t)\right)^{\vee}=\Omega(t)-\sigma(t), (20)
σ(t)\displaystyle\sigma(t) AdRe1Ωr(t)=ReT(t)Ωr(t).\displaystyle\triangleq\mathrm{Ad}_{R^{-1}_{e}}\Omega_{r}(t)=R^{T}_{e}(t)\Omega_{r}(t).

Therefore, the distance between Re(t)R_{e}(t) and I3I_{3} is properly measured with the Morse function V1(Re)21+tr(Re(t))V_{1}(R_{e})\triangleq 2-\sqrt{1+\mathrm{tr}(R_{e}(t))}, proposed by Lee (2012). In fact, V1(Re)=0Re=I3V_{1}(R_{e})=0\iff R_{e}=I_{3} and is positive for all ReSO(3)\{I3}R_{e}\in SO(3)\backslash\{I_{3}\}. Moreover, along the trajectories of (20), it satisfies

ddtV1(Re)\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}V_{1}(R_{e}) =ψ(Re)(Re(t)ReT(t)),Ωe(t),\displaystyle=\left\langle\left\langle\psi(R_{e})\left(R_{e}(t)-R^{T}_{e}(t)\right)^{\vee},\;\Omega_{e}(t)\right\rangle\right\rangle,
ψ(Re)\displaystyle\psi(R_{e}) 121+tr(Re),\displaystyle\triangleq\frac{1}{2\sqrt{1+\mathrm{tr}(R_{e})}},

for all ReSO(3)\𝒪RR_{e}\in SO(3)\backslash\mathcal{O}_{R}, where 𝒪R{RSO(3)|tr(R)\mathcal{O}_{R}\triangleq\{R\in SO(3)|\mathrm{tr}(R) =1}=-1\}. Furthermore, given 𝒰R\mathcal{U}_{R} \triangleq {ReSO(3)\\{R_{e}\in SO(3)\backslash 𝒪R|V1(Re)<2ϵ}\mathcal{O}_{R}\;|\;V_{1}(R_{e})<2-\epsilon\}, for some ϵ>0\epsilon>0 arbitrarily small, V1(Re)V_{1}(R_{e}) verifies (Lee, 2012)

ψ(Re)(ReReT)2\displaystyle\left\|\psi(R_{e})\left(R_{e}-R^{T}_{e}\right)^{\vee}\right\|^{2} V1(Re)2ψ(Re)(ReReT)2,\displaystyle\leq V_{1}(R_{e})\leq 2\left\|\psi(R_{e})\left(R_{e}-R^{T}_{e}\right)^{\vee}\right\|^{2},

for all Re𝒰RR_{e}\in\mathcal{U}_{R}.

Consider the kinematic control law

Ωu(Re)\displaystyle\Omega_{u}(R_{e}) log(Re),\displaystyle\equiv\mathrm{log}(R_{e})^{\vee}, (21)
log(Re)\displaystyle\mathrm{log}(R_{e}) {03×3,Re=I3,ϕ(Re)2sin(ϕ(Re))(ReReT),ReI3,\displaystyle\triangleq\left\{\begin{array}[]{ll}0_{3\times 3},&R_{e}=I_{3},\\ \frac{\phi(R_{e})}{2\sin\left(\phi(R_{e})\right)}\left(R_{e}-R^{T}_{e}\right),&R_{e}\neq I_{3},\end{array}\right. (24)

where ϕ(Re)arccos(12(tr(Re)1))(π,π)\phi(R_{e})\triangleq\arccos{\left(\frac{1}{2}\left(\mathrm{tr}(R_{e})-1\right)\right)}\in(-\pi,\pi), and 0n×m0_{n\times m} is a matrix of size n×mn\times m with zero-entries. It can verify readily Definition 3.1(i)-(ii) by (21). To verify Definition 3.1(iii)-(iv) under the kinematic control Ωe(t)=Ωu(Re)\Omega_{e}(t)=-\Omega_{u}(R_{e}), consider the derivative of the Morse function V1(Re)V_{1}(R_{e}) along the error kinematics R˙e=ReΩe\dot{R}_{e}=R_{e}\Omega_{e}^{\wedge}:

V˙1(Re)=ψ(Re)(ReReT),Ωu(Re)<0,\dot{V}_{1}(R_{e})=\left\langle\left\langle\psi(R_{e})\left(R_{e}-R^{T}_{e}\right)^{\vee},\;-\Omega_{u}(R_{e})\right\rangle\right\rangle<0,

for all ReSO(3)\𝒪RR_{e}\in SO(3)\backslash\mathcal{O}_{R} and V˙1(Re)y1(Re)V1(Re)\dot{V}_{1}(R_{e})\leq-y_{1}(R_{e})V_{1}(R_{e}) for all Re𝒰RR_{e}\in\mathcal{U}_{R}, where

y1(Re)ϕ(Re)4ψ(Re)sinϕ(Re)>0,Re𝒰R.y_{1}(R_{e})\triangleq\frac{\phi(R_{e})}{4\psi(R_{e})\sin\phi(R_{e})}>0,\;\forall R_{e}\in\mathcal{U}_{R}.

This proves that the kinematic control (21) also holds Definition 3.1(iii)-(iv).

Therefore, based on the kinematic control law (21) the following group operation is defined

r1r2\displaystyle r_{1}\star r_{2} (25)
=\displaystyle= (R1R2,Ω1+Ω2+λΩu(R1)+λΩu(R2)λΩu(R1R2)),\displaystyle\bigg{(}R_{1}R_{2},\;\Omega_{1}+\Omega_{2}+\lambda\Omega_{u}(R_{1})+\lambda\Omega_{u}(R_{2})-\lambda\Omega_{u}(R_{1}R_{2})\Big{)},

for any r1=(R1,Ω1)r_{1}=(R_{1},\Omega_{1}), r2=(R2,Ω2)TSO(3)r_{2}=(R_{2},\Omega_{2})\in TSO(3). Thus, the tangent bundle TSO(3)SO(3)×3TSO(3)\simeq SO(3)\times\mathbb{R}^{3} is endowed with a Lie group structure with identity (I3,03×1)TSO(3)(I_{3},0_{3\times 1})\in TSO(3) and inverse r1=(RT,Ω)TSO(3)r^{-1}=(R^{T},-\Omega)\in TSO(3), r=(R,Ω)TSO(3)\forall r=(R,\Omega)\in TSO(3). Likewise, given re(t)=(Re(t),Ωe(t))TSO(3)r_{e}(t)=\left(R_{e}(t),\Omega_{e}(t)\right)\in TSO(3) and in view of (15), the map s:TSO(3)3s:TSO(3)\to\mathbb{R}^{3}

s(re)=Ωe(t)+λΩu(Re),s(r_{e})=\Omega_{e}(t)+\lambda\Omega_{u}(R_{e}), (26)

for some scalar λ>0\lambda>0, defines a Lie subgroup

HR={re(t)=(Re(t),Ωe(t))TSO(3)|s(re)=03×1},H_{R}=\{r_{e}(t)=\left(R_{e}(t),\Omega_{e}(t)\right)\in TSO(3)\;|\;s(r_{e})=0_{3\times 1}\}, (27)

under the group operation (25).

Thus, the tracking controller on SO(3)SO(3) is obtained from (4.2) and (26) as

τu\displaystyle\tau_{u} =𝕁1((𝕁Ω(t))(λΩu(Re)σ(t)))λΩ˙u(Re)+σ˙(t)\displaystyle=\mathbb{J}^{-1}\left(\left(\mathbb{J}\Omega(t)\right)^{\wedge}\left(\lambda\Omega_{u}(R_{e})-\sigma(t)\right)\right)-\lambda\dot{\Omega}_{u}(R_{e})+\dot{\sigma}(t)
kss(re),\displaystyle\quad-k_{s}s(r_{e}), (28)

where ks>0k_{s}>0 is a controller gain. Theorem 9 proves that controller (28) in closed loop with the system (19) renders the equilibrium point re(t)=(I3,03×1)r_{e}(t)=\left(I_{3},0_{3\times 1}\right) almost globally asymptotically stable for all re(0)SO(3)\𝒪r×3r_{e}(0)\in SO(3)\backslash\mathcal{O}_{r}\times\mathbb{R}^{3}, and exponentially stable for all re(0)𝒰r×3r_{e}(0)\in\mathcal{U}_{r}\times\mathbb{R}^{3}.

Remark 11.

Note that in applying Theorem 9 it should define first a tracking error using the group operation on the configuration manifold, and then treat the error dynamics as a physical system. Otherwise, the sliding surface may not be a Lie subgroup. To see this more clear, consider r=(R,Ω)r=\left(R,\Omega\right), rd1=(RrT,Ωr)TSO(3)r^{-1}_{d}=\left(R^{T}_{r},-\Omega_{r}\right)\in TSO(3), then in the following tracking error may be defined by the group operation (25)

re\displaystyle r^{\prime}_{e} =rd1r=(RrTR,Ωr+Ω+λΩu(RrT)+λΩu(R)\displaystyle=r^{-1}_{d}\star r=\left(R^{T}_{r}R,\;-\Omega_{r}+\Omega+\lambda\Omega_{u}(R^{T}_{r})+\lambda\Omega_{u}(R)\right.
λΩu(RrTR))\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\left.-\lambda\Omega_{u}(R^{T}_{r}R)\right)
=(Re,Ωr+Ω+λΩu(RrT)+λΩu(R)λΩu(Re))\displaystyle=\left(R_{e},\;-\Omega_{r}+\Omega+\lambda\Omega_{u}(R^{T}_{r})+\lambda\Omega_{u}(R)-\lambda\Omega_{u}(R_{e})\right)
=(Re,Ω¯e).\displaystyle=\left(R_{e},\bar{\Omega}_{e}\right).

However, HR={re(t)TSO(3)|s(re)=03×1}TSO(3)H^{\prime}_{R}=\{r^{\prime}_{e}(t)\in TSO(3)\;|\;s(r^{\prime}_{e})=0_{3\times 1}\}\subset TSO(3) is not a sliding subgroup for the proposed Morse function V1(Re)V_{1}(R_{e}).

5.2 Attitude Tracking on 𝒮3\mathcal{S}^{3}

The set of 4\mathbb{R}^{4}-vectors evolving on the unit sphere 𝒮3={q4|qTq=1}\mathcal{S}^{3}=\left\{q\in\mathbb{R}^{4}\;|\;q^{T}q=1\right\}, with q=[q0,qT]T𝒮3q=\left[q_{0},\vec{q}^{T}\right]^{T}\in\mathcal{S}^{3}, q0[1,1]q_{0}\in[-1,1], and q3\vec{q}\in\mathbb{R}^{3}, is a Lie group with identity ı=[1,01×3]T𝒮3\imath=\left[1,0_{1\times 3}\right]^{T}\in\mathcal{S}^{3}, and inverse q1=[q0,qT]T𝒮3q^{-1}=\left[q_{0},-\vec{q}^{T}\right]^{T}\in\mathcal{S}^{3}, under the group operation (q1,q2)q1q2𝒮3(q_{1},q_{2})\mapsto q_{1}\otimes q_{2}\in\mathcal{S}^{3} defined as

q1q2Q(q1)q2=[q0,1q1Tq1q0,1I3+q1][q0,2q2],q_{1}\otimes q_{2}\triangleq Q(q_{1})q_{2}=\left[\begin{array}[]{cc}q_{0,1}&-\vec{q}^{T}_{1}\\ \vec{q}_{1}&q_{0,1}I_{3}+\vec{q}_{1}^{\wedge}\end{array}\right]\left[\begin{array}[]{c}q_{0,2}\\ \vec{q}_{2}\end{array}\right],

for any q1=[q0,1,q1T]Tq_{1}=\left[q_{0,1},\vec{q}^{T}_{1}\right]^{T}, q2=[q0,2,q2T]T𝒮3q_{2}=\left[q_{0,2},\vec{q}^{T}_{2}\right]^{T}\in\mathcal{S}^{3}. The Lie algebra is 𝔰3={ω4|ω=[0,ΩT]T,Ω3}\mathfrak{s}^{3}=\left\{\omega\in\mathbb{R}^{4}\;|\;\omega=\left[0,\Omega^{T}\right]^{T},\Omega\in\mathbb{R}^{3}\right\}, which holds 𝔰33\mathfrak{s}^{3}\simeq\mathbb{R}^{3}. Its Lie bracket operation corresponds to the cross product in 3\mathbb{R}^{3}. Thus, denote the isomorphism ¯:3𝔰3\overline{\cdot}\vcentcolon\mathbb{R}^{3}\to\mathfrak{s}^{3} with the inverse map ¯:𝔰33\underline{\cdot}\vcentcolon\mathfrak{s}^{3}\to\mathbb{R}^{3}.

Rodriguez formula qR(q)=I3+2q0q+2q2SO(3)q\mapsto R(q)=I_{3}+2q_{0}\vec{q}^{\wedge}+2\vec{q}^{\wedge 2}\in SO(3) relates each antipodal point ±q\pm q with a physical rotation of a rigid body, i.e., 𝒮3\mathcal{S}^{3} double covers the group SO(3)SO(3). The adjoint action in 𝒮3\mathcal{S}^{3} is defined as Adqζqζ¯q1=R(q)ζ¯\mathrm{Ad}_{q}\zeta\triangleq q\otimes\bar{\zeta}\otimes q^{-1}=\overline{R(q)\zeta}, for any ζ3\zeta\in\mathbb{R}^{3}.

Given a differentiable curve q:I𝒮3q\vcentcolon I\to\mathcal{S}^{3} with a left-invariant vector field q˙(t)Tq𝒮3\dot{q}(t)\in T_{q}\mathcal{S}^{3}, and a twice-differentiable reference configuration qr:I𝒮3q_{r}\vcentcolon I\to\mathcal{S}^{3}, tI\forall t\in I, the body angular velocity Ω¯(t)2q1(t)q˙(t)=2QT(q(t))q˙(t)𝔰3\overline{\Omega}(t)\triangleq 2q^{-1}(t)\otimes\dot{q}(t)=2Q^{T}(q(t))\dot{q}(t)\in\mathfrak{s}^{3} and the reference angular velocity Ω¯r(t)2qr1(t)q˙r(t)𝔰3\overline{\Omega}_{r}(t)\triangleq 2q^{-1}_{r}(t)\otimes\dot{q}_{r}(t)\in\mathfrak{s}^{3} can be defined. We consider the following intrinsic tracking error qe(t)qr1(t)q(t),q_{e}(t)\triangleq q^{-1}_{r}(t)\otimes q(t), and its left-invariant velocity error

Ω¯e(t)\displaystyle\overline{\Omega}_{e}(t) 2qe1(t)q˙e(t)=Ω¯(t)ζ¯(t),\displaystyle\triangleq 2q^{-1}_{e}(t)\otimes\dot{q}_{e}(t)=\overline{\Omega}(t)-\overline{\zeta}(t), (29)
ζ¯(t)\displaystyle\overline{\zeta}(t) =Adqe1Ωr(t),\displaystyle=\mathrm{Ad}_{q^{-1}_{e}}\Omega_{r}(t),

where q˙e(t)Tqe𝒮3\dot{q}_{e}(t)\in T_{q_{e}}\mathcal{S}^{3} is left invariant. Propose the Morse function 𝒮3qV2(q)=12ıq=1q0\mathcal{S}^{3}\ni q\mapsto V_{2}(q)=\frac{1}{\sqrt{2}}\left\|\imath-q\right\|=\sqrt{1-q_{0}}, which satisfies V2(q)=0q=ıV_{2}(q)=0\iff q=\imath, and V2(q)>0V_{2}(q)>0 q𝒮3\{ı}\forall q\in\mathcal{S}^{3}\backslash\{\imath\}. That is, function V2(q)V_{2}(q) has a unique minimum critical zero at identity ı𝒮3\imath\in\mathcal{S}^{3} and is strictly positive for any other q𝒮3q\in\mathcal{S}^{3}. Moreover, it verifies that

ddtV2(qe)=q˙0,e(t)21q0,e(t)=141q0,e(t)qeT(t)Ωe(t),\frac{\mathrm{d}}{\mathrm{d}t}V_{2}(q_{e})=\frac{-\dot{q}_{0,e}(t)}{2\sqrt{1-q_{0,e}(t)}}=\frac{1}{4\sqrt{1-q_{0,e}(t)}}\vec{q}^{T}_{e}(t)\Omega_{e}(t),

which suggests the following kinematic control law

Ωu(qe)log(qe){03×1,qe=ı,arccos(q0,e)qeqe,qeı,\Omega_{u}(q_{e})\equiv\mathrm{log}(q_{e})\triangleq\left\{\begin{array}[]{ll}0_{3\times 1},&q_{e}=\imath,\\ \frac{\arccos(q_{0,e})}{\|\vec{q}_{e}\|}\vec{q}_{e},&q_{e}\neq\imath,\end{array}\right. (30)

for all qe(t)𝒮3\{ı}q_{e}(t)\in\mathcal{S}^{3}\backslash\{-\imath\}. Indeed, when Ωe(t)=Ωu(qe)\Omega_{e}(t)=-\Omega_{u}(q_{e}), it leads to

ddtV2(qe)\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}V_{2}(q_{e}) =arccos(q0,e)41q0,eqe\displaystyle=-\frac{\arccos(q_{0,e})}{4\sqrt{1-q_{0,e}}}\|\vec{q}_{e}\|
=arccos(q0,e)41q0,e1q0,e2\displaystyle=-\frac{\arccos(q_{0,e})}{4\sqrt{1-q_{0,e}}}\sqrt{1-q^{2}_{0,e}}
=arccos(q0,e)41q0,e(1+q0,e)(1q0,e)\displaystyle=-\frac{\arccos(q_{0,e})}{4\sqrt{1-q_{0,e}}}\sqrt{\left(1+q_{0,e}\right)\left(1-q_{0,e}\right)}
=arccos(q0,e)41q0,e1+q0,eV2(qe),\displaystyle=-\frac{\arccos(q_{0,e})}{4\sqrt{1-q_{0,e}}}\sqrt{1+q_{0,e}}V_{2}(q_{e}),
=y2(qe)V2(qe),\displaystyle=-y_{2}(q_{e})V_{2}(q_{e}),

where y2(qe)>0y_{2}(q_{e})>0 for all qe𝒰q{qe𝒮3\{ı}|V2(qe)<2ϵ}q_{e}\in\mathcal{U}_{q}\triangleq\{q_{e}\in\mathcal{S}^{3}\backslash\{\imath\}\;|\;V_{2}(q_{e})<2-\epsilon\}, for some ϵ>0\epsilon>0 arbitrarily small. Consequently, the control law (30) satisfies all properties of Definition 3.1 for the Morse function V2(qe)V_{2}(q_{e}).

The kinematic control law (30) enables the definition of the tangent bundle T𝒮3𝒮3×3T\mathcal{S}^{3}\simeq\mathcal{S}^{3}\times\mathbb{R}^{3} as a Lie group under the group operation

p1p2\displaystyle p_{1}\star p_{2} (31)
=\displaystyle= (q1q2,Ω1+Ω2+λΩu(q1)+λΩu(q2)λΩu(q1q2)),\displaystyle\left(q_{1}\otimes q_{2},\;\Omega_{1}+\Omega_{2}+\lambda\Omega_{u}(q_{1})+\lambda\Omega_{u}(q_{2})-\lambda\Omega_{u}(q_{1}\otimes q_{2})\right),

p1=(q1,Ω1)\forall p_{1}=(q_{1},\Omega_{1}), p2=(q2,Ω2)T𝒮3p_{2}=(q_{2},\Omega_{2})\in T\mathcal{S}^{3}. Note that the identity is (ı,03×1)T𝒮3(\imath,0_{3\times 1})\in T\mathcal{S}^{3}, and inverse, p1=(q1,Ω)T𝒮3p^{-1}=(q^{-1},-\Omega)\in T\mathcal{S}^{3}, p=(q,Ω)T𝒮3\forall p=(q,\Omega)\in T\mathcal{S}^{3}. Therefore, the map

s(pe)=Ωe(t)+λΩu(qe),s(p_{e})=\Omega_{e}(t)+\lambda\Omega_{u}(q_{e}), (32)

where pe(t)=(qe(t),Ωe(t))T𝒮3p_{e}(t)=(q_{e}(t),\Omega_{e}(t))\in T\mathcal{S}^{3}, defines the sliding Lie subgroup

Hq={peT𝒮3|s(pe)=03×1}.H_{q}=\left\{p_{e}\in T\mathcal{S}^{3}\;|\;s(p_{e})=0_{3\times 1}\right\}. (33)

The attitude tracking controller on 𝒮3\mathcal{S}^{3} is thus defined as (4.2) using (29)-(32), which yields

τu\displaystyle\tau_{u} =𝕁1((𝕁Ω(t))(λΩu(qe)ζ(t)))λΩ˙u(qe)+ζ˙(t)\displaystyle=\mathbb{J}^{-1}\left(\left(\mathbb{J}\Omega(t)\right)^{\wedge}\left(\lambda\Omega_{u}(q_{e})-\zeta(t)\right)\right)-\lambda\dot{\Omega}_{u}(q_{e})+\dot{\zeta}(t)
kss(pe).\displaystyle\quad-k_{s}s(p_{e}). (34)

By Theorem 9 controller (34) in closed loop with system (18) achieves the asymptotic convergence of pe(t)(ı,03×1)p_{e}(t)\to(\imath,0_{3\times 1}) for all pe(0)𝒮3\{ı}×3p_{e}(0)\in\mathcal{S}^{3}\backslash\{\imath\}\times\mathbb{R}^{3}, and exponential convergence when pe(0)𝒰q×3p_{e}(0)\in\mathcal{U}_{q}\times\mathbb{R}^{3}.

6 Simulations

To illustrate the theoretical results and for comparison, the proposed GSMC (28) was contrasted with two reported controllers: the ”linearization”-by-state-feedback-like (LSF) controller Eq. (26) of Maithripala et al. (2006), and the PD+ controller Eq. (23) of Lee (2012). For easy comparison, the applied torque control for each controller is rewritten in terms of

si=ω~i+γiφ~i,i=1,2,3,s_{i}=\tilde{\omega}_{i}+\gamma_{i}\tilde{\varphi}_{i},\quad\forall i=1,2,3, (35)

where ω~i3\tilde{\omega}_{i}\in\mathbb{R}^{3} is angular velocity error, φ~i3\tilde{\varphi}_{i}\in\mathbb{R}^{3} is attitude error, and γi>0\gamma_{i}>0 is the control gain.

The proposed GSMC law (28) is expressed as

τ1\displaystyle\tau_{1} =ks𝕁s1+F1,\displaystyle=-k_{s}\mathbb{J}s_{1}+F_{1}, (36)
s1\displaystyle s_{1} =Ωe(t)+λΩu(Re),\displaystyle=\Omega_{e}(t)+\lambda\Omega_{u}(R_{e}), (37)
F1\displaystyle F_{1} =𝕁(λΩ˙u(Re)+σ˙)+(𝕁Ω)(λΩu(Re)σ).\displaystyle=\mathbb{J}\left(-\lambda\dot{\Omega}_{u}(R_{e})+\dot{\sigma}\right)+\left(\mathbb{J}\Omega\right)^{\wedge}\left(\lambda\Omega_{u}(R_{e})-\sigma\right).

Likewise, the LSF controller (26) of Maithripala et al. (2006) is given by

τ2\displaystyle\tau_{2} =k𝕁s2+F2,\displaystyle=-k\mathbb{J}s_{2}+F_{2}, (38)
s2\displaystyle s_{2} =ΩΩr+κkRT(RRrTRrRT),\displaystyle=\Omega-\Omega_{r}+\frac{\kappa}{k}R^{T}\left(RR^{T}_{r}-R_{r}R^{T}\right)^{\vee}, (39)
F2\displaystyle F_{2} =𝕁Ω˙r(𝕁Ω)(Ω)ΩΩr,\displaystyle=\mathbb{J}\dot{\Omega}_{r}-\left(\mathbb{J}\Omega\right)^{\wedge}\left(\Omega\right)-\Omega^{\wedge}\Omega_{r},

where I~=I3\tilde{I}=I_{3} and K=κI3K=\kappa I_{3}, for some κ>0\kappa>0. Finally, the PD+ controller (23) of Lee (2012) is rewritten as

τ3\displaystyle\tau_{3} =kΩs3+F3,\displaystyle=-k_{\Omega}s_{3}+F_{3}, (40)
s3\displaystyle s_{3} =Ωe+kRkΩψ(Re)(ReReT),\displaystyle=\Omega_{e}+\frac{k_{R}}{k_{\Omega}}\psi(R_{e})\left(R_{e}-R^{T}_{e}\right)^{\vee}, (41)
F3\displaystyle F_{3} =𝕁RTRrΩ˙r+(RTRrΩr)𝕁RTRrΩr.\displaystyle=\mathbb{J}R^{T}R_{r}\dot{\Omega}_{r}+\left(R^{T}R_{r}\Omega_{r}\right)^{\wedge}\mathbb{J}R^{T}R_{r}\Omega_{r}.

The inertia tensor was given by

𝕁=[3.60460.07060.14910.07068.68680.04490.14910.04499.3484],\mathbb{J}=\left[\begin{array}[]{ccc}3.6046&-0.0706&0.1491\\ -0.0706&8.6868&0.0449\\ 0.1491&0.0449&9.3484\end{array}\right],

while the reference trajectory was calculated as Ωr(t)=(RrT(t)R˙r(t))=[0,0.1,0]T\Omega_{r}(t)=\left(R^{T}_{r}(t)\dot{R}_{r}(t)\right)^{\vee}=\left[0,0.1,0\right]^{T} (rad/s). Furthermore, the initial conditions were chosen as Ω(0)=(1/(214))[1,2,3]T\Omega(0)=\left(1/\left(2\sqrt{14}\right)\right)\left[1,2,3\right]^{T} (rad/s), Rr(0)=R312(π/4,π,π/4)R_{r}(0)=R_{312}(\pi/4,-\pi,\pi/4), where the expression R312(φ,ϑ,ψ)R_{312}(\varphi,\vartheta,\psi) is a rotation matrix described by the sequence 3-1-2 of Euler angles (Shuster et al., 1993), and the initial attitude was calculated as R(0)=Rr(0)Re(0)R(0)=R_{r}(0)R_{e}(0).

The simulations were carried out under three scenarios according to the distance between Re(0)R_{e}(0) and the desired equilibrium I3I_{3}, and to the undesired equilibrium diag(1,1,1)\mathrm{diag}(1,-1,-1) measured by the Morse function Ψ(Re)12tr(I3Re)\Psi(R_{e})\triangleq\frac{1}{2}\mathrm{tr}(I_{3}-R_{e}) used in Maithripala et al. (2006). Therefore, the initial attitudes Re(0)=R312(0,0.428π,0)R_{e}(0)=R_{312}(0,-0.428\pi,0), Re(0)=R312(0,0.01π,0)I3R_{e}(0)=R_{312}(0,-0.01\pi,0)\approx I_{3}, and Re(0)=R312(0,0.99π,0)diag(1,1,1)R_{e}(0)=R_{312}(0,-0.99\pi,0)\approx\mathrm{diag}(1,-1,-1) were assigned.

Finally, the design parameters for each controller were tuned in such a way that the energy-consumption level measured by 0tτiT(t)τi(t)dt\sqrt{\int^{t}_{0}\tau^{T}_{i}(t)\tau_{i}(t)\mathrm{d}t} in the first scenario is the same. The resulting controller gains were ks=1k_{s}=1, λ=0.5\lambda=0.5 for (36), k=1k=1, κ=0.5\kappa=0.5 for (38), and kΩ=18.5k_{\Omega}=18.5, kR=9.25k_{R}=9.25 for (40). With these design parameters the control gain of (35) was γi=0.5\gamma_{i}=0.5 for all i=1,2,3i=1,2,3.

6.1 Scenario 1. Intermediate case.

Figure 1 shows the performance of the controllers (36), (38), and (40) under the initial condition Re(0)=R312(0,0.428π,0)R_{e}(0)=R_{312}(0,-0.428\pi,0). Fig. 1(a) shows the attitude error Ψ(Re)12tr(I3Re)\Psi(R_{e})\triangleq\frac{1}{2}\mathrm{tr}(I_{3}-R_{e}), it is observed that the proposed controller (36) and controller (38) achieve the convergence ReI3R_{e}\to I_{3} in 1717 (s), while controller (40) achieves it in 3030 (s). Fig. 1(b) illustrates the norm Ωe(t)\|\Omega_{e}(t)\| for each controller, where the angular velocity error Ωe(t)\Omega_{e}(t) is calculated as (20), it can be seen that controller (40) takes 1010 (s) longer than the other controllers to reach Ωe(t)03×1\Omega_{e}(t)\to 0_{3\times 1}. Furthermore, Figs. 1(c) and (d) draw the control effort and the energy consumption respectively, it is observed that, with the selected controller gains, all controllers consume the same amount of energy. Finally, Fig. 2 shows the behavior of the sliding variables (37), (39), and (41) compared to si=0s_{i}=0 according to (35). It is observed that the proposed controller (36) allows convergence s103×1s_{1}\to 0_{3\times 1} to complete the reach phase, while the LSF control scheme (38) presents an oscillatory behavior around the equilibrium point, in addition to the PD + controller (40) that follows closely si=0s_{i}=0 until it reaches equilibrium.

Refer to caption
Refer to caption
Figure 1: Scenario 1: Behavior of controllers (36), (38), and (40) when the initial attitude error is Re(0)=R312(0,0.428π,0)R_{e}(0)=R_{312}(0,-0.428\pi,0).
Refer to caption
Figure 2: Scenario 1: Behavior of the sliding variable (37), (39), and (41) when Re(0)=R312(0,0.428π,0)R_{e}(0)=R_{312}(0,-0.428\pi,0).

6.2 Scenario 2. Starting close to the desired equilibrium point I3I_{3}.

For this scenario, the initial condition was set to Re(0)=R312(0,0.01π,0)R_{e}(0)=R_{312}(0,-0.01\pi,0), which corresponds to an initial condition close to the desired equilibrium I3I_{3}. Figs. 3(a) and (b) show that the controllers (36) and (38) reach the desired equilibrium (Re,Ωe)=(I3,03×1)(R_{e},\Omega_{e})=(I_{3},0_{3\times 1}) at the same time 2020 (s), while the controller (40) takes 55 (s) longer, which coincides with the previous scenario. However, as illustrated in Fig. 3(d), the proposed controller uses less energy than others to reach the desired equilibrium when the system starts close to the desired equilibrium.

Refer to caption
Refer to caption
Figure 3: Scenario 2: Behavior of controllers (36), (38), and (40) when the initial attitude error is close to I3I_{3}, i.e., Re(0)=R312(0,0.01π,0)R_{e}(0)=R_{312}(0,-0.01\pi,0).

6.3 Scenario 3. Starting close to the undesired equilibrium point diag(1,1,1)\mathrm{diag}(1,-1,-1).

Figure 4 displays the performance of the controllers starting close to the undesired equilibrium point diag(1,1,1)\mathrm{diag}(1,-1,-1), i.e., Re(0)=R312(0,0.99π,0)R_{e}(0)=R_{312}(0,-0.99\pi,0). It is observed in Fig. 4(a) that the proposed controller (36) and the PD+ controller (36) present a delay of 11 (s) before beginning the convergence of ReI3R_{e}\to I_{3}, however, the LSF controller (38) has the longest delay of 2.52.5 (s). Notice that the proposed control scheme allows a faster convergence to the desired equilibrium point (Figs. 4(a) and (b)) at a cost of more energy consumption (Figs. 4(c) and (d)).

Refer to caption
Refer to caption
Figure 4: Scenario 3: Behavior of controllers (36), (38), and (40) when the initial attitude error is close to diag(1,1,1)\mathrm{diag}(1,-1,-1)), i.e., Re(0)=R312(0,0.99π,0)R_{e}(0)=R_{312}(0,-0.99\pi,0).

7 Conclusions

This paper presented a geometric sliding mode control for fully actuated mechanical systems evolving on Lie groups, generalizing the conventional sliding mode control in Euclidean spaces. It was shown that the sliding surface (a Lie subgroup) is immersed in the state space (a Lie group) of the system dynamics, and the tracking is achieved by first driving the trajectories of the system to the sliding subgroup and then converging to the group identity of the reduced dynamics restricted on the sliding subgroup, like sliding mode control designs for systems evolving on Euclidean spaces. An application of the result to attitude control was presented for the rotation group SO(3)SO(3) and the unit sphere 𝒮3\mathcal{S}^{3}. The simulation results illustrated the scheme and compared it with similar control designs in the literature.

References

  • (1)
  • Abraham et al. (2012) Abraham, R., Marsden, J. E. & Ratiu, T. (2012), Manifolds, tensor analysis, and applications, Vol. 75, Springer Science & Business Media.
  • Aghannan & Rouchon (2003) Aghannan, N. & Rouchon, P. (2003), ‘An intrinsic observer for a class of lagrangian systems’, IEEE Transactions on Automatic Control 48(6), 936–945.
  • Akhtar & Waslander (2020) Akhtar, A. & Waslander, S. L. (2020), ‘Controller class for rigid body tracking on so(3)’, IEEE Transactions on Automatic Control 66(5), 2234–2241.
  • Bhat & Bernstein (2000) Bhat, S. P. & Bernstein, D. S. (2000), ‘A topological obstruction to continuous global stabilization of rotational motion and the unwinding phenomenon’, Systems & control letters 39(1), 63–70.
  • Bonnabel et al. (2009) Bonnabel, S., Martin, P. & Rouchon, P. (2009), ‘Non-linear symmetry-preserving observers on lie groups’, IEEE Transactions on Automatic Control 54(7), 1709–1713.
  • Bullo & Lewis (2005) Bullo, F. & Lewis, A. D. (2005), Geometric control of mechanical systems: modeling, analysis, and design for simple mechanical control systems, Vol. 49, Springer.
  • Bullo & Murray (1999) Bullo, F. & Murray, R. M. (1999), ‘Tracking for fully actuated mechanical systems: a geometric framework’, Automatica 35(1), 17–34.
  • Bullo et al. (1995) Bullo, F., Murray, R. M. & Sarti, A. (1995), ‘Control on the sphere and reduced attitude stabilization’, IFAC Proceedings Volumes 28(14), 495–501.
  • Culbertson et al. (2021) Culbertson, P., Slotine, J.-J. & Schwager, M. (2021), ‘Decentralized adaptive control for collaborative manipulation of rigid bodies’, IEEE Transactions on Robotics 37(6), 1906–1920.
  • De Marco et al. (2018) De Marco, S., Marconi, L., Mahony, R. & Hamel, T. (2018), ‘Output regulation for systems on matrix lie-groups’, Automatica 87, 8–16.
  • Espíndola & Tang (2022) Espíndola, E. & Tang, Y. (2022), ‘A novel four-dof lagrangian approach to attitude tracking for rigid spacecraft’, (submitted to Automatica), available at arXiv preprint arXiv:2202.05227 .
  • Ghasemi et al. (2020) Ghasemi, K., Ghaisari, J. & Abdollahi, F. (2020), ‘Robust formation control of multiagent systems on the lie group se (3)’, International Journal of Robust and Nonlinear Control 30(3), 966–998.
  • Gómez et al. (2019) Gómez, C. G. C., Castanos, F. & Dávila, J. (2019), Sliding motions on so (3), sliding subgroups, in ‘2019 IEEE 58th Conference on Decision and Control (CDC)’, IEEE, pp. 6953–6958.
  • Koditschek (1989) Koditschek, D. (1989), Autonomous mobile robots controlled by navigation functions, in ‘Proceedings. IEEE/RSJ International Workshop on Intelligent Robots and Systems’.(IROS’89)’The Autonomous Mobile Robots and Its Applications’, IEEE, pp. 639–645.
  • Lageman et al. (2009) Lageman, C., Trumpf, J. & Mahony, R. (2009), ‘Gradient-like observers for invariant dynamics on a lie group’, IEEE Transactions on Automatic Control 55(2), 367–377.
  • Lee (2013) Lee, J. M. (2013), Smooth manifolds, in ‘Introduction to smooth manifolds’, Springer, pp. 1–31.
  • Lee (2011) Lee, T. (2011), Geometric tracking control of the attitude dynamics of a rigid body on so (3), in ‘Proceedings of the 2011 American Control Conference’, IEEE, pp. 1200–1205.
  • Lee (2012) Lee, T. (2012), ‘Exponential stability of an attitude tracking control system on so (3) for large-angle rotational maneuvers’, Systems & Control Letters 61(1), 231–237.
  • Liang et al. (2021) Liang, W., Chen, Z. & Yao, B. (2021), ‘Geometric adaptive robust hierarchical control for quadrotors with aerodynamic damping and complete inertia compensation’, IEEE Transactions on Industrial Electronics 69(12), 13213–13224.
  • Lopez & Slotine (2021) Lopez, B. T. & Slotine, J.-J. E. (2021), Sliding on manifolds: Geometric attitude control with quaternions, in ‘2021 IEEE International Conference on Robotics and Automation (ICRA)’, IEEE, pp. 11140–11146.
  • Mahony et al. (2008) Mahony, R., Hamel, T. & Pflimlin, J.-M. (2008), ‘Nonlinear complementary filters on the special orthogonal group’, IEEE Transactions on automatic control 53(5), 1203–1218.
  • Maithripala & Berg (2015) Maithripala, D. S. & Berg, J. M. (2015), ‘An intrinsic pid controller for mechanical systems on lie groups’, Automatica 54, 189–200.
  • Maithripala et al. (2006) Maithripala, D. S., Berg, J. M. & Dayawansa, W. P. (2006), ‘Almost-global tracking of simple mechanical systems on a general class of lie groups’, IEEE Transactions on Automatic Control 51(2), 216–225.
  • Meng et al. (2023) Meng, Q., Yang, H. & Jiang, B. (2023), ‘Second-order sliding-mode on so (3) and fault-tolerant spacecraft attitude control’, Automatica 149, 110814.
  • Olver et al. (1996) Olver, P. et al. (1996), ‘Non-associative local lie groups’, Journal of Lie theory 6, 23–52.
  • Rodríguez-Cortés & Velasco-Villa (2022) Rodríguez-Cortés, H. & Velasco-Villa, M. (2022), ‘A new geometric trajectory tracking controller for the unicycle mobile robot’, Systems & Control Letters 168, 105360.
  • Saccon et al. (2013) Saccon, A., Hauser, J. & Aguiar, A. P. (2013), ‘Optimal control on lie groups: The projection operator approach’, IEEE Transactions on Automatic Control 58(9), 2230–2245.
  • Sarlette et al. (2010) Sarlette, A., Bonnabel, S. & Sepulchre, R. (2010), ‘Coordinated motion design on lie groups’, IEEE Transactions on Automatic Control 55(5), 1047–1058.
  • Shuster et al. (1993) Shuster, M. D. et al. (1993), ‘A survey of attitude representations’, Navigation 8(9), 439–517.
  • Spong & Bullo (2005) Spong, M. W. & Bullo, F. (2005), ‘Controlled symmetries and passive walking’, IEEE Transactions on Automatic Control 50(7), 1025–1031.
  • Utkin (1977) Utkin, V. (1977), ‘Variable structure systems with sliding modes’, IEEE Transactions on Automatic control 22(2), 212–222.
  • Zlotnik & Forbes (2018) Zlotnik, D. E. & Forbes, J. R. (2018), ‘Gradient-based observer for simultaneous localization and mapping’, IEEE Transactions on Automatic Control 63(12), 4338–4344.

Appendix A A passivity-like lemma

Lemma 12.

Given that the inner product on 𝔤\mathfrak{g} is a symmetric bilinear map, for any ζ,η𝔤\zeta,\eta\in\mathfrak{g}, it holds

𝕀(ζ,𝕀(adζ𝕀(η)))\displaystyle\mathbb{I}\left(\zeta,\;\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\zeta}\mathbb{I}^{\flat}(\eta)\right)\right) =𝕀(𝕀(adζ𝕀(η)));ζ\displaystyle=\langle\mathbb{I}^{\flat}\left(\mathbb{I}^{\sharp}\left(\mathrm{ad}^{*}_{\zeta}\mathbb{I}^{\flat}(\eta)\right)\right);\zeta\rangle
=adζ𝕀(η);ζ\displaystyle=\langle\mathrm{ad}^{*}_{\zeta}\mathbb{I}^{\flat}(\eta);\zeta\rangle
=𝕀(η);[ζ,ζ]\displaystyle=\langle\mathbb{I}^{\flat}(\eta);\left[\zeta,\zeta\right]\rangle
=0,\displaystyle=0,

because of the skew symmetry of the Lie bracket operation [,]𝔤[\cdot,\cdot]\in\mathfrak{g}.