Abstract

A Riemannian-geometry approach for modeling and control of dynamics of object manipulation under holonomic or non-holonomic constraints is presented. First, position/force hybrid control of an endeffector of a multijoint redundant (or nonredundant) robot under a holonomic constraint is reinterpreted in terms of “submersion” in Riemannian geometry. A force control signal constructed in the image space of the constraint gradient is regarded as a lifting (or pressing) in the direction orthogonal to the kernel space. By means of the Riemannian distance on the constraint submanifold, stability of position control under holonomic constraints is discussed. Second, modeling and control of two-dimensional object grasping by a pair of multijoint robot fingers are challenged, when the object is of arbitrary shape. It is shown that rolling contact constraints induce the Euler equation of motion, in which constraint forces appear as wrench vectors affecting the object. The Riemannian metric is introduced on a constraint submanifold characterized with arclength parameters. An explicit form of the quotient dynamics is expressed in the kernel space with accompaniment of a pair of first-order differential equations concerning the arclength parameters. An extension of Dirichlet-Lagrange's stability theorem to redundant systems under constraints is suggested by introducing a Morse-Lyapunov function.

1. Introduction

Among roboticsists, it is implicitly known that robot motions can be interpreted in terms of orbits on a high-dimensional torus or trajectories in an 𝑛-dimensional configuration space. Planning of robot motions has been investigated traditionally on the basis of kinematics on a configuration space as an 𝑛-dim numerical space 𝐑𝑛 [1].

This paper first emphasizes a mathematical observation that, given a robot as a multibody mechanism with n degrees of freedom whose endpoint is free, the set of all its postures can be regarded as a Riemannian manifold (𝑀, 𝑔) associated with the Riemannian metric 𝑔 that constitutes the robot inertia matrix. A geodesic connecting any two postures can correspond to an orbit expressed on a local coordinate chart and generated by a solution to the Euler-Lagrange equation of robot motion that originates only from the force of inertia [2, 3]. It should be emphasized that once the Riemannian manifold is given corresponding to the 𝑛 degrees of freedom robot, the collection of all the geodesic paths describes the “law of inertia” for the manifold. It is also important to note that geodesic paths are invariant under any choice of local coordinates. This Riemannian geometry viewpoint is extended in this paper to an important class of multibody dynamics physically interacting with an object or with environment through holonomic or/and nonholonomic (but Pfaffian) constraints. Holonomic constraints are defined as a set of infinitely differentiable functions from a product manifold of multibody Riemannian manifolds onto an open set of a 2- or 3-dimensional Euclidean space called the task space. Such a mapping can be treated as a submersion from the product Riemannian manifold to 𝑚(=2-or3-) dimensional Euclidean space. Hence, holonomic constraints induce a Riemannian submanifold with a naturally induced metric. An Euler-Lagrange equation is formulated in an implicit function form under such constraints. It is also shown that if the gravity term can be explicitly compensated and there arises no viscous friction then the geodesic motion is invariant, that is, it is governed by the “law of inertia,” under any adequate lifting (or pressing) through the joint torque injection in the direction along the constraint gradient. An explicit form of the Euler equation whose solution corresponds to a geodesic on the submanifold is given also as a quotient dynamics corresponding to the kernel space as an orthogonal compliment to the image space spanned from all the constraint gradients. Based upon these observations, the well-known methodology of hybrid (position/force) control for a robot whose end effector is constrained on a surface is re-examined and shown to be effective even if the robot is of redundancy in its degrees of freedom.

In a latter part of the paper, modeling of dynamics of grasping and manipulation of a two-dimensional rigid object with arbitrary shape by using a pair of multijoint robot fingers with spherical finger ends is challenged. It is shown that rolling contact constraints between finger ends and the object surfaces induce not only two holonomic constraints but also two nonholonomic constraints that restrict tangent vectors on the original Riemannian manifold that is a product of three manifolds expressing a set of whole postures of the two fingers and the object. An Euler-Lagrange equation for expressing the dynamics of such physical interaction is derived through applying the variational principle together with deriving a set of the first-order differential equations expressing the contact positions of the object with both the finger ends. The Riemannian distance is introduced on the kernel space as an orthogonal compliment to the image space of all the gradients vectors of both contact and rolling constraints. In other words, rolling constraints are expressed in terms of the first fundamental forms of given contours of the object and restrict only the tangent vector fields at both the contact points. An explicit Euler-Lagrange equation corresponding to a path on the constraint submanifold is derived together with a set of the first-order differential equations expressed in terms of the second fundamental forms of the object contours. Thus, it is shown that rolling constraints can be characterized by means of arc length parameters of the object contours that express locations of the contact points and in the sequel are integrable in the sense of Frobenius. A coordinated control signal called “blind grasping” without referring to the object kinematics or external sensing is proposed and shown to be effective in realizing stable grasping in the sense of stability on a submanifold. A sketch of the convergence proof is given on the basis of an extension of the Dirichlet-Lagrange theorem to a system of degrees of freedom redundancy by finding a Morse-Lyapunov function and using its physical properties and mathematical meanings.

2. Riemannian Manifold: A Set of All Postures

Lagrange's equation of motion of a multijoint system with 2 degrees of freedom (DOF) shown in Figure 1 is described by the formula𝐻(𝑞)̈𝑞+12̇𝐻(𝑞)+𝑆(𝑞,̇𝑞)̇𝑞+𝑔(𝑞)=𝑢,(1)where 𝑞=(𝑞1,𝑞2)𝑇 denotes the vector of joint angles, 𝐻(𝑞) denotesthe inertia matrix, 𝑆(𝑞,̇𝑞)̇𝑞 the gyroscopic force term including centrifugal and Coriolis forces, 𝑔(𝑞) the gradient vector of a potential function 𝑃(𝑞) due to the gravity with respect to 𝑞, and 𝑢 the joint torque generated by joint actuators [4]. It is well known that the inertia matrix 𝐻(𝑞) is symmetric and positive definite, and there exist a positive constant 𝑚 together with a positive definite constant diagonal matrix 𝐻0 such that𝑚𝐻0𝐻(𝑞)𝐻0(2)for any 𝑞. It should be also noted that 𝑆(𝑞,̇𝑞) is skew symmetric and linear and homogeneous in ̇𝑞. More in detail, the 𝑖𝑗th entry of 𝑆(𝑞,̇𝑞) denoted by 𝑠𝑖𝑗 can be described in the form [3]𝑠𝑖𝑗=12𝜕𝜕𝑞𝑗𝑛𝑘=1̇𝑞𝑘𝑖𝑘𝜕𝜕𝑞𝑖𝑛𝑘=1̇𝑞𝑘𝑗𝑘,(3)where 𝐻(𝑞)=(ℎ𝑖𝑗(𝑞)), from which it follows apparently that 𝑠𝑖𝑗=−𝑠𝑗𝑖. Since we assume that the objective system to be controlled is a series of rigid links serially connected through each rotational joint with single DOF, every entry of 𝐻(𝑞) is a constant or a sinusoidal function of components of joint angle vector 𝑞. That is, each element of 𝐻(𝑞) and 𝑔(𝑞) is differentiable of class 𝐶 (infinitely differentiable in 𝑞).

When two joint angles 𝜃1 and 𝜃2 are given in 𝜃𝑖∈(−𝜋,𝜋], 𝑖=1,2, for the 2 DOF robot arm shown in Figure 1, the posture 𝑝(𝜃1,𝜃2) is determined naturally. Denote the set of all such possible postures by 𝑀 and introduce a family of subsets of 𝑀 such that, for any 𝑝∈𝑀 with joint angles 𝑝=(𝜃1,𝜃2) and any number 𝛼>0, a set of all 𝑝′=(𝜃1,𝜃2) is defined as𝑈𝑝,𝛼=𝑝𝑝−𝑝𝐻<𝛼,(4)where𝑝−𝑝𝐻=𝑖,𝑗𝑖𝑗(𝑝)𝜃𝑖−𝜃𝑖𝜃𝑗−𝜃𝑗(5)can be regarded as an open subset of 𝑀. Then, the set 𝑀 with such a family of open subsets can be regarded as a topological manifold. It is possible to show that the manifold 𝑀 becomes Hausdorff and compact. Further, every point 𝑝 of 𝑀 has a neighborhood 𝑈 that is homeomorphic to an open subset Ω of 2-dimensional numerical space 𝐑2. Such a homeomorphism 𝜙∶𝑈→Ω is called a coordinate chart. In fact, a neighborhood 𝑈𝑝,𝛼 of posture 𝑝 with joint angles (𝜃1,𝜃2) in Figure 1 can be mapped to an open set Ω in 𝐑2 with 2 dimensional numerical coordinates (𝑞1,𝑞2) with the origin 𝑂 (Figure 2). In this case, it is possible to see that the original set 𝑀 of robot postures can be visualized as a torus shown in 𝐑3 (see Figure 3) in which angles 𝑞1 and 𝑞2 are defined. It is quite fortunate to see that, in the case of typical robots like the one shown in Figure 1, the local coordinates (𝑞1,…,𝑞𝑛) can be identically chosen as a set of 𝑛 independent joint angles (𝜃1,…,𝜃𝑛) by setting 𝑞𝑖=𝜃𝑖(𝑖=1,…,𝑛). It is also interesting to see that the torus in Figure 3 is made to be topologically coincident with the set of all arm endpoints 𝑃=(𝑥,𝑦,𝑧). As discussed in detail in mathematical text books [5, 6], the topological manifold (𝑀,𝑝) of such a torus can be regarded as a differentiable manifold of class 𝐶.

Now, it is necessary to define a tangent vector to an abstract differentiable manifold 𝑀 at 𝑝∈𝑀. Let 𝐼 be an interval (−𝜖,𝜖) and define a curve 𝑐(𝑡) by a mapping 𝑐∶𝐼→𝑀 such that 𝑐(0)=𝑝. A tangent vector to 𝑀 at 𝑝 is an equivalence class of curves 𝑐∶𝐼→𝑀 for the equivalence relation defined by

𝑐∼𝑐if and only if, in a coordinate chart (𝑈,𝜙) around 𝑝, we have (𝜙∘𝑐)(0)=(𝜙∘𝑐)(0), where symbol ()′ means differentiation of () with respect to 𝑡∈𝐼. This definition of tangent vectors to 𝑀 at 𝑝 does not depend on choice of the coordinate chart at 𝑝, as discussed in text books [5, 6]. Let us denote the set of all tangent vectors to 𝑀 at 𝑝 by 𝑇𝑝𝑀 and call it the tangent space at 𝑝𝑀. It has an 𝑛-dimensional linear space structure like 𝐑𝑛. We also denote the disjoint union of the tangent spaces to 𝑀 at all the points of 𝑀 by 𝑇𝑀 and call it the tangent bundle of 𝑀.

Now, we are in a position to define a Riemannian metric on a differentiable manifold (𝑀,𝑝) as a mapping 𝑔𝑝𝑇𝑝𝑀×𝑇𝑝𝑀𝐑 such that 𝑝𝑔𝑝 is of class 𝐶 and 𝑔𝑝(𝑢,𝑣) for 𝑢𝑇𝑝𝑀 and 𝑣𝑇𝑝𝑀 is a symmetric positive definite quadratic form𝑔𝑝(𝑢,𝑣)=𝑛𝑖,𝑗=1𝑔𝑖𝑗(𝑝)𝑢𝑖𝑣𝑗.(6)

Suppose that 𝑀 is a connected Riemannian manifold. If 𝑐𝐼[𝑎,𝑏]𝑀 is a curve segment of class 𝐶, we define the length of 𝑐 to be𝐿(𝑐)=𝑏𝑎=̇𝑐(𝑡)d𝑡𝑏𝑎𝑔𝑐(𝑡)̇𝑐(𝑡),̇𝑐(𝑡)d𝑡,(7)where we assume ̇𝑐(𝑡)0 for any 𝑡𝐼 and call such a curve segment to be regular. A mapping of class 𝐶𝑐[𝑎,𝑏]𝑀 is called a piecewise regular curve segment (for brevity, we call it an admissible curve) if there exists a finite subdivision 𝑎=𝑎0<𝑎1<<𝑎𝑘=𝑏 such that 𝑐(𝑡) for 𝑡[𝑎𝑖1,𝑎𝑖] is a regular curve for 𝑖=1,,𝑘. Then, it is possible to define for any pair of points 𝑝,𝑝𝑀 the Riemannian distance 𝑑(𝑝,𝑝) to be the infimum of all admissible curves from 𝑝 to 𝑝. It is well known [4, 5] that, with the distance function 𝑑 defined above, any connected Riemannian manifold becomes a metric space whose induced topology is coincident with the given manifold topology. An admissible curve 𝑐 in a Riemannian manifold is said to be minimizing if 𝐿(𝑐)𝐿(̃𝑐) for any other admissible curve ̃𝑐 with the same endpoints. It follows immediately from the definition of distance that 𝑐 is minimizing if and only if 𝐿(𝑐) is equal to the distance between its endpoints. Further, it is known that if the Riemannian manifold {𝑀,𝑔} is complete, then for any pair of points 𝑝 and 𝑝 there exists at least a minimizing curve 𝑐(𝑡), 𝑡[𝑎,𝑏], with 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑝. If such a minimizing curve 𝑐(𝑡) is described with the aid of coordinate chart (𝑈,𝜙) as 𝜙(𝑐(𝑡))=(𝑞1(𝑡),,𝑞𝑛(𝑡)), then 𝑞(𝑡)=𝜙(𝑐(𝑡)) satisfies the 2nd-order differential equationd2d𝑡2𝑞𝑘(𝑡)+𝑛𝑖,𝑗=1Γ𝑘𝑖𝑗(𝑐(𝑡))d𝑞𝑖(𝑡)d𝑡d𝑞𝑗(𝑡)d𝑡=0,(𝑘=1,,𝑛),(8)where Γ𝑘𝑖𝑗 denotes Christoffel's symbol defined byΓ𝑘𝑖𝑗=12𝑛=1𝑔𝑘𝜕𝑔𝑖𝜕𝑞𝑗+𝜕𝑔𝑗𝜕𝑞𝑖𝜕𝑔𝑖𝑗𝜕𝑞,(9)and (𝑔𝑘) denotes the inverse of matrix (𝑔𝑘). A curve 𝑞(𝑡)𝐼𝑈 satisfying (8) together with 𝜙1(𝑞(𝑡)) is called a geodesic, and (8) itself is called the Euler-Lagrange equation or the geodesic equation.

Given a 𝐶-class curve 𝑐(𝑡)=𝐼[𝑎,𝑏]𝑀, the quantity1𝐸(𝑐)=2𝑏𝑎̇𝑐(𝑡)2=1d𝑡2𝑏𝑎𝑔𝑐(𝑡)̇𝑐(𝑡),̇𝑐(𝑡)d𝑡(10)is called the energy of the curve. Then, by applying the Cauchy-Schwartz inequality for (7), we have𝐿(𝑐)22(𝑏𝑎)𝐸(𝑐).(11)Further, the equality of (11) follows if and only if ̇𝑐(𝑡) is constant. It is also possible to see that if 𝑐(𝑡) is a geodesic with 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑝, then for any other curve 𝑐(𝑡) with the same endpoints, it holds𝐸(𝑐)=𝐿(𝑐)2𝐿2(𝑏𝑎)𝑐22(𝑏𝑎)𝐸𝑐.(12)The equalities hold if and only if 𝑐(𝑡) is also a geodesic. Conversely, if 𝑐(𝑡) with 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑝 is a 𝐶 curve that minimizes the energy and makes 𝑔𝑐(𝑡)(̇𝑐(𝑡),̇𝑐(𝑡)) constant, then 𝑐(𝑡) becomes a geodesic connecting 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑝. In mechanics, 𝐸(𝑐) is usually called “action of 𝑐,” and 𝑐(𝑡) is considered as the orbit of motion of a multibody system.

3. Riemannian Geometry of Robot Dynamics

Dynamics of a robotic mechanism with 𝑛 rigid bodies connected in series through rotational joints are described by Lagrange's equation of motion, as shown in (1). It is implicitly assumed that the axis of rotation of the first body is fixed in an inertial frame and denoted by 𝑧-axis that is perpendicular to the 𝑥𝑦-plane as shown in Figure 1. If there is no gravity force affecting motion of the robot, then the equation of motion of the robot can be described by the form1𝐻(𝑞)̈𝑞+2̇𝐻(𝑞)+𝑆𝑞,̇𝑞̇𝑞=𝑢,(13)where 𝑢 stands for a vector of control torques emanating from joint actuators. This formula is valid for motions of a revolute joint robot, shown in Figure 1, if it is installed in weightless environment like an artificial satellite, or the gravity term 𝑔(𝑞) (included in (1)) can be compensated by joint actuators through control input 𝑢. In general, we can represent a posture 𝑝 of the robot as a physical entity by a family of joint angles 𝜃𝑖(𝑖=1,,𝑛), which can be expressed by a point Θ=(𝜃1,,𝜃𝑛) in the 𝑛-dimensional numerical space 𝐑𝑛. In fact, we can naturally imagine an isomorphism 𝜙𝑈Ω, where 𝑈𝑀, and Ω is an open subset of 𝐑𝑛. In other words, a local coordinate chart 𝜙(𝑈)(=Ω) can be treated to be identical to 𝑈 itself, an open subset of 𝑀, by regarding 𝑞=(𝑞1,,𝑞𝑛)𝑇 (“𝑇” denotes transpose and hence 𝑞 a column vector) identical to Θ by setting 𝑞𝑖=𝜃𝑖(𝑖=1,,𝑛). In this way, the abstract manifold 𝑀 as the set of all robot postures can be regarded as an 𝑛-dimensional torus 𝑇𝑛 as an 𝑛-tuple direct product of 𝑆1𝑇𝑛=𝑆1××𝑆1. Hence, a robot posture 𝑝𝑀 can be represented by a point Θ on 𝑇𝑛 and also expressed by a joint vector 𝑞 in 𝐑𝑛.

From the definition of inertia matrices, 𝐻(𝑞) is symmetric and positive definite, and the kinetic energy is expressed as a quadratic form𝐾=1𝑞,̇𝑞2̇𝑞𝑇𝐻(𝑞)̇𝑞.(14)Hence, the equation of motion of the robot is expressed by Lagrange's equationd𝜕d𝑡𝐿𝜕𝜕̇𝑞𝑞,̇𝑞𝐿𝜕𝑞𝑞,̇𝑞=𝑢,(15)where 𝐿(𝑞,̇𝑞)=𝐾(𝑞,̇𝑞), and 𝑢 stands for a generalized external force vector. It is interesting to note that in differential geometry, (15) can be described as𝑖𝑘𝑖̈𝑞𝑖+𝑖,𝑗Γ𝑖𝑘𝑗(𝑞)̇𝑞𝑖̇𝑞𝑗=𝑢𝑘,(16)where Γ𝑖𝑘𝑗 denotes Christoffel's symbol of the first kind defined byΓ𝑖𝑘𝑗=12𝜕𝑗𝑘𝜕𝑞𝑖+𝜕𝑖𝑘𝜕𝑞𝑗𝜕𝑖𝑗𝜕𝑞𝑘.(17)For later use, we introduce another Christoffel's symbol called the second kind as shown in the formulaΓ𝑘𝑖𝑗=12𝑛𝑙=1𝑙𝑘𝜕𝑗𝑙𝜕𝑞𝑖+𝜕𝑖𝑙𝜕𝑞𝑗𝜕𝑖𝑗𝜕𝑞𝑙=12𝑛𝑙=1𝑙𝑘Γ𝑖𝑙𝑗,(18)where (𝑙𝑘) denotes the inverse of (𝑙𝑘), the inertia matrix 𝐻(𝑞)=(𝑙𝑘). Since (𝑘𝑙) and (𝑘𝑙) are symmetric, it follows that Γ𝑖𝑘𝑗=Γ𝑗𝑘𝑖 and Γ𝑘𝑖𝑗=Γ𝑘𝑗𝑖. Now, we show that (13) is equivalent to (16) by bearing in mind that ̇𝐻(𝑞)=𝑖{𝜕𝐻(𝑞)/𝜕𝑞𝑖}̇𝑞𝑖, and the skew symmetric matrix 𝑆(𝑞,̇𝑞) is expressed as in (3). In fact, the second term in the bracket () of (17) corresponds to the first term in {} of (3) and the third term of (17) does to the second term in{} of (3). Hence, it follows from (3) that𝑛𝑗=1𝑠𝑘𝑗̇𝑞𝑗=𝑛𝑗=112𝜕𝜕𝑞𝑗𝑛𝑖=1̇𝑞𝑖𝑘𝑖̇𝑞𝑖𝜕𝜕𝑞𝑘𝑛𝑖=1̇𝑞𝑖𝑖𝑗̇𝑞𝑗=𝑛𝑛𝑗=1𝑖=112𝜕𝑖𝑘𝜕𝑞𝑗𝜕𝑖𝑗𝜕𝑞𝑘̇𝑞𝑖̇𝑞𝑗.(19)Substituting this into (16) by comparing the last two terms of (17) with the last bracket {} of (19) results in the equivalence of (13) to (16). It is easy to see that multiplication of (16) by 𝐻1(𝑞) yields̈𝑞𝑘+𝑖,𝑗Γ𝑘𝑖𝑗̇𝑞𝑖̇𝑞𝑗=𝑗𝑘𝑗𝑢𝑗,𝑘=1,,𝑛.(20)If 𝑢=0, this expression is nothing, but the Euler-Lagrange equation shown in (8). By this reason, from now on, we use symbol 𝑔𝑖𝑗(𝑞) instead of 𝑖𝑗(𝑞) for the inertia matrix 𝐻(𝑞) even when robot dynamics are treated.

Now, on the abstract topological manifold 𝑀 as a set of all possible postures of a robot, suppose that a Riemannian metric is given by a scalar product on each tangent space 𝑇𝑝𝑀:𝑣,𝑤=𝑔𝑖𝑗(𝑝)𝑣𝑖𝑤𝑗,(21)where 𝑣=𝑣𝑖(𝜕/𝜕𝑞𝑖)𝑇𝑝𝑀 and 𝑤=𝑤𝑗(𝜕/𝜕𝑞𝑗)𝑇𝑝𝑀, and the summation symbol in 𝑖 and 𝑗 is omitted, and 𝑞=(𝑞1,,𝑞𝑛) represents local coordinates. Then, the manifold (𝑀,𝑝) can be regarded as a Riemannian manifold and becomes complete as a metric space. Then, according to the Hopf-Rinow theorem [5], any two points 𝑝,𝑞𝑀 can be joined by a geodesic of length 𝑑(𝑝,𝑞), that is, a curve satisfying (8) with shortest length, where𝑑(𝑝,𝑞)=inf𝑏𝑎̇𝑐(𝑡)d𝑡=inf𝑏𝑎̇𝑐(𝑡),̇𝑐(𝑡)d𝑡(22)with 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑞.

As discussed in the previous section, geodesics are the critical points of the energy functional 𝐸(𝑐). Further, a geodesic curve 𝑐(𝑡) satisfies ̇𝑐(𝑡)=const. In fact, by regarding 𝑐(𝑡)=𝑞(𝑡) that is an orbit on Ω, we haveddd𝑡̇𝑞,̇𝑞=𝑔d𝑡𝑖𝑗𝑞(𝑡)̇𝑞𝑖(𝑡)̇𝑞𝑗=d(𝑡)d𝑡̇𝑞𝑇(𝑡)𝐺(𝑞(𝑡))̇𝑞(𝑡)=̈𝑞𝑇(𝑡)𝐺(𝑞)̇𝑞(𝑡)+̇𝑞𝑇̇(𝑡)𝐺(𝑞)̇𝑞(𝑡)+̇𝑞𝑇(𝑡)𝐺(𝑞)̈𝑞(𝑡)=0,(23)where the equivalent expression1𝐺(𝑞)̈𝑞+2̇𝐺(𝑞)+𝑆𝑞,̇𝑞̇𝑞=0(24)to (13) with 𝑢=0 is used, 𝐺(𝑞)=(𝑔𝑖𝑗(𝑞)), and 𝑠𝑖𝑗 of 𝑆 is given in the form of (3) (where 𝑖𝑗=𝑔𝑖𝑗). It is also important to note that, on a local coordinate chart Ω⊂𝐑𝑛 corresponding to a neighborhood 𝑈 of 𝑝∈𝑀, an orbit 𝑞(𝑡) parameterized by time 𝑡∈[𝑎,𝑏] and expressed by a solution to (20) (where 𝑢𝑗 is of 𝐶 in 𝑡) should satisfy𝑛𝑗=1̇𝑞𝑗(𝑡)𝑢𝑗(𝑡)=dd𝑡𝐸(𝑞(𝑡)),(25)or𝑏𝑎𝑛𝑗=1̇𝑞𝑗(𝑡)𝑢𝑗(𝑡)d𝑡=𝐸(𝑞(𝑏))−𝐸(𝑞(𝑎)),(26)as long as 𝑞(𝑡)∈Ω, where 𝐸(𝑞(𝑡))=(1/2)⟨̇𝑞(𝑡),̇𝑞(𝑡)⟩. When 𝑢(𝑡)=0, 𝐸(𝑞(𝑡))=const. and then the curve connecting 𝑝=𝜙−1(𝑞(𝑎)) and 𝑝=𝜙−1(𝑞(𝑏)) must be a geodesic. In other words, an inertia-originated movement without being affected by the gravitational field or any external force field produces a geodesic orbit [2]. The most importantly, geodesics together with their length are invariant under any choice of local coordinates.

Before closing this expository section on robot motion from the Riemannian geometry viewpoint, we must emphasize that all the above invariant properties of geodesics of inertia-originated robot motions result from imaging a set of all robot postures as an abstract Riemannian manifold. Choice of a local coordinates is originally arbitrary. Even an 𝑛-dimensional torus 𝑇𝑛 is one of such choice of local coordinates corresponding to the choice of joint angles 𝑞=(𝑞1,…,𝑞𝑛)𝑇. At the same time, it is important to note that, in differential geometry, choice of coordinates in the tangent space 𝑇𝑝𝑀 is indeterminable or free to choose. However, once a local coordinates system for an 𝑛 degrees of freedom robot is chosen by joint angle vector 𝑞=(𝑞1,…,𝑞𝑛)𝑇, then the coordinates in the tangent space 𝑇𝑝𝑀 should be chosen as the vector of joint angular velocities 𝜕/𝜕𝑞=(𝜕/𝜕𝑞1,…,𝜕/𝜕𝑞𝑛) correspondingly to ̇𝑞=(̇𝑞1,…,̇𝑞𝑛)𝑇, from which the Riemannian metric 𝑔𝑖𝑗(𝑞) is defined through the inertia matrix.

4. Constraint Submanifold and Hybrid Position/Force Control

Consider an 𝑛-DOF robotic arm whose last link is a pencil and suppose that the endpoint of the pencil is in contact with a flat surface 𝜑(𝐱)=𝜉, where 𝐱=(𝑥,𝑦,𝑧)𝑇. It is well known that the Lagrange equation of motion of the system is written as𝐺(𝑞)̈𝑞+12̇𝐺+𝑆̇𝑞+𝑔(𝑞)=−𝜆𝜕𝜑(𝐱(𝑞))𝜕𝑞+𝑢,(27)where 𝜕𝜑/𝜕𝑞 can be decomposed into 𝜕𝜑(𝑞)/𝜕𝑞=𝐽𝑇(𝑞)𝜕𝜑/𝜕𝐱 and 𝐽(𝑞)=𝜕𝐱/𝜕𝑞𝑇. On the constraint manifold 𝐹𝜉={𝑝∶𝑝∈𝑀and𝜑(𝐱(𝑝))=𝜉}, let us consider a smooth curve 𝑐(𝑡)∶𝐼[𝑎,𝑏]→𝐹𝜉 that connects the given two points 𝑐(𝑎)=𝑝 and 𝑐(𝑏)=𝑝, where 𝑝 and 𝑝 belong to 𝐹𝜉. The length of such a curve constrained to 𝐹𝜉 is defined as𝐿(𝑐)=𝑏𝑎𝑔𝑖𝑗(𝑐(𝑡))̇𝑐𝑖(𝑡)̇𝑐𝑗(𝑡)d𝑡,(28)and consider the minimization𝑑𝑝,𝑝=inf𝑐∈𝐹𝜉𝐿(𝑐)(29)that should be called the distance between 𝑝 and 𝑝 on the constraint manifold. Then, the minimizing curve called the geodesic denoted identically by 𝑞(𝑡)(=𝑐(𝑡)) must satisfy the Euler equation̈𝑞𝑘(𝑡)+Γ𝑘𝑖𝑗̇𝑞𝑖(𝑡)̇𝑞𝑗(𝑡)=−𝜆(𝑡)⋅grad𝜑(𝐱(𝑡))𝑘,(30)together with the constraint condition 𝜑(𝐱(𝑡))=𝜉, wheregrad𝜑(𝐱(𝑡))=𝐺−1(𝑞(𝑡))𝐽𝑇(𝑞)𝜕𝜑𝜕𝐱,(31)and 𝐽𝑇(𝑞)=𝜕𝐱𝑇/𝜕𝑞. It should be noted that, from the inner product of (30) and 𝑤=𝐽𝑇𝜕𝜑/𝜕𝐱, it follows that𝑘𝑤𝑘̈𝑞𝑘+𝑤𝑘Γ𝑘𝑖𝑗̇𝑞𝑖̇𝑞𝑗=−𝜆𝑘𝑤𝑘𝑔𝑘𝑗𝑤𝑗.(32)Since the holonomic constraint 𝜑(𝐱(𝑞))=𝜉 implies ⟨𝑤,̇𝑞⟩=0, it follows that4𝑘=1𝑤𝑘̈𝑞𝑘+dd𝑡𝑤𝑘̇𝑞𝑘=0.(33)Substituting this into (32), we obtain𝜆(𝑡)=1𝑤𝑇𝐺−1𝑤𝑘̇𝑤𝑘̇𝑞𝑘𝑖,𝑗𝑤𝑘Γ𝑘𝑖𝑗̇𝑞𝑖̇𝑞𝑗.(34)From the Riemannian geometry, the constraint force 𝜆(𝑡) with the grad{𝜑(𝐱(𝑞))} stands for a component of the image space of 𝑤(=𝐽𝑇(𝑞)𝜕𝜑/𝜕𝐱) that is orthogonal to the kernel 𝑇𝐹𝜉 of 𝑤. In other words, this component is cancelled out by the image space component of the left hand side of (32). From the physical point of view, 𝜆(𝑡) should be regarded as a magnitude of the constraint force that presses the surface 𝜑(𝐱(𝑞))=𝜉 in its normal direction. In order to compromise the mathematical argument with such physical reality, let us suppose that the joint actuators can supply the control torques𝑢=𝜆𝑑⋅𝐽𝑇(𝑞)𝜕𝜑𝜕𝐱+𝑔(𝑞).(35)Then, by substituting this into (27), we obtain the Lagrange equation of motion under the constraint 𝜑(𝐱(𝑞))=𝜉:𝐺(𝑞)̈𝑞+12̇𝐺+𝑆̇𝑞=−𝜆−𝜆𝑑𝐽𝑇(𝑞)𝜕𝜑𝜕𝐱=−Δ𝜆𝐽𝑇(𝑞)𝜕𝜑𝜕𝐱,(36)where Δ𝜆=𝜆−𝜆𝑑. It should be noted that introduction of the first term of control signal of (35) does not affect the solution orbit on the constraint manifold, and further it keeps the constraint condition during motion by rendering 𝜆(𝑡)(=𝜆𝑑+Δ𝜆(𝑡)) positive. In a mathematical sense, exertion of the joint torque 𝜆𝑑𝐽𝑇(𝜕𝜑/𝐱) plays a role of “pressing” or “lifting” of the image space spanned from the gradient of the constraint equation. Further, note that (36) is of an implicit function form with the Lagrange multiplier Δ𝜆. To affirm the argument of treatment of the geodesics through this implicit form, we show an explicit form of the Lagrange equation expressed on the orthogonally projected space (kernel space) by introducing the orthogonal transformatioṅ𝑞=𝑃,𝑤‖𝑤‖−1̇𝜼̇𝑧=𝑄̇𝑞,(37)where 𝑃 is a 4×3 matrix whose column vectors with the unit norm are orthogonal to 𝑤, and ̇𝜼 denotes a 3×1 matrix (3-dim. vector) and ̇𝑧 a scalar. Since 𝑄 is an orthogonal matrix, 𝑄−1=𝑄𝑇. Hence, if ̇𝑞∈ker(𝑤)(=𝑇𝐹𝜉), then ̇𝑧=0. Restriction of (36) to the kernel space of 𝑤 can be attained by multiplying (36) by 𝑃𝑇 from the left such that𝑃𝑇𝐺(𝑞)dd𝑡(𝑃̇𝜼)+𝑃𝑇12̇𝐺+𝑆𝑃̇𝜼=0,(38)which is reduced to the Eular equation in ̇𝜼:𝐺(𝑞)̈𝜼+12̇𝐺+𝑆̇𝜼=0,(39)or equivalentlÿ𝜂𝑘+Γ𝑘𝑖𝑗̇𝜂𝑖̇𝜂𝑗=0,𝑘=1,…,𝑛−1(=3),(40)where 𝐺(𝑞)=𝑃𝑇𝐺(𝑞)𝑃, Γ𝑘𝑖𝑗 the Christoffel symbol for (𝑔𝑖𝑗)=𝐺, and𝑆=𝑃𝑇𝑆𝑃−12̇𝑃𝑇𝐺𝑃+12𝑃𝑇𝐺̇𝑃,(41)which is skew symmetric, too. Note that the transformation 𝑄 is isometric, and (40) stands for the geodesic equation on the constraint Riemannian submanifold.

5. Hybrid Control for Redundant Systems and Stability on a Submanifold

Let us now reconsider a hybrid position/force feedback control scheme, which is of the form𝑢=𝑔(𝑞)+𝜆𝑑𝐽𝑇(𝑞)𝜕𝜑𝜕𝐱+𝐶̇𝑞+𝐽𝑇(𝑞)𝜁𝑘̇𝐱+𝑘Δ𝐱,(42)where 𝜑(𝐱)=𝑧, ̇𝐱=(̇𝑥,̇𝑦,0)𝑇, Δ𝐱=(𝑥−𝑥𝑑,𝑦−𝑦𝑑,0). This type of hybrid control is an extension of the McClamroch and Wang's method [7, 8] to cope with a robotic system that is subject to redundancy in DOFs. From now on, we redefine the Jacobian matrix 𝐽(𝑞) as 2×4 matrix 𝐽(𝑞)=(𝜕𝑥/𝜕𝑞𝑇,𝜕𝑦/𝜕𝑞𝑇) because we consider a special case of 𝜑(𝐱)=𝑧, that is, the constraint 𝑧−𝜉=0, and solution trajectories 𝑞(𝑡) that satisfy 𝑧(𝑞(𝑡))−𝜉=0. In relation to this, denote also the two-dimensional vector (𝑥,𝑦) by 𝑥 and (Δ𝑥,Δ𝑦) by Δ𝐱. It should be noted that the orthogonality relationship among (𝑥,𝑦,𝑧) coordinates in 𝐄3 does not imply the orthogonality among 𝜕𝑥/𝜕𝑞, 𝜕𝑦/𝜕𝑞, and 𝜕𝑧/𝜕𝑞(=𝜕𝜑/𝜕𝑞). Therefore, 𝜕𝑧/𝜕𝑞 is not orthogonal to 𝐽𝑇(𝑞)(=(𝜕𝑥/𝜕𝑞,𝜕𝑦/𝜕𝑞)), and hence the last term of the right hand side of (42) contains both components of image (𝑤) and ker(𝑤).

Now, substituting this into (27) yields𝐺(𝑞)̈𝑞+12̇𝐺+𝑆+𝐶̇𝑞+𝐽𝑇(𝑞)𝜁𝑘̇𝐱+𝑘Δ𝐱=−Δ𝜆𝜕𝑧(𝑞)𝜕𝑞.(43)Evidently the inner product of (43) and ̇𝑞 under the constraint 𝑧(𝑞)=𝜉 leads todd𝑡𝐸𝑞,̇𝑞=−̇𝑞𝐶̇𝑞−𝜁𝑘̇𝐱2,(44)where𝐸𝑞,̇𝑞=12̇𝑞𝑇𝐺(𝑞)̇𝑞+𝑘2‖Δ𝐱‖2.(45)Unfortunately, this quantity is not positive definite in the tangent bundle 𝑇𝐹𝜉. Nevertheless, it is possible to see that magnitudes of ̇𝑞(𝑡) and Δ𝐱(𝑡) remain small if at initial time 𝑡=0 both magnitudes ‖̇𝑞(0)‖ and ‖Δ𝐱(0)‖ are set small. Next, let us introduce the equilibrium manifold defined by the setEM1=𝑞∶𝑧(𝑞)=𝜉,𝑥(𝑞)=𝑥𝑑,𝑦(𝑞)=𝑦𝑑,(46)which is of one dimension. Fortunately as discussed in the paper [9], it is possible to confirm that a modified scalar function𝑊𝛼=𝐸𝑞,̇𝑞+𝛼̇𝑞𝑇𝐺(𝑞)𝑃𝜑𝐽𝑇(𝑞)Δ𝐱(47)becomes positive definite in 𝑇𝐹𝜉 with an appropriate positive parameter 𝛼>0, provided that the physical scale of the robot is ordinary, and feedback gains 𝐶 and 𝑘 are chosen adequately. Here, in (47), 𝑃𝜑 is defined as 𝐼4−𝑤𝑤𝑇/‖𝑤‖2. Furthermore, if in a Riemannian ball in 𝐹𝜉 defined as 𝐵(𝑞(0),𝑟0)={𝑞∶𝑑(𝑞,𝑞(0))<𝑟0}, the Jacobian matrix 𝐽(𝑞) is nondegenerate, then it can be expected that there exist positive numbers 𝜎𝑚 and 𝜎𝑀 such that𝜎𝑀𝐼2≥𝐽𝐽𝑇≥𝐽𝑃𝜑𝐽𝑇≥𝜎𝑚𝐼2.(48)Further, to simplify the argument, we choose 𝐶=𝑐𝐼4 with a constant 𝑐>0. Then, it follows from differentiation of 𝑉(=̇𝑞𝑇𝐺(𝑞)𝑃𝜑𝐽𝑇(𝑞)Δ𝐱) thaṫ𝑉=−𝜁𝑘̇𝐱+𝑘Δ𝐱𝑇𝐽𝑃𝜑𝐽𝑇Δ𝐱−𝑐̇𝑞𝑇𝑃𝜑𝐽𝑇Δ𝐱+̇𝑞𝑇𝐺𝑃𝜑𝐽𝑇̇𝐱−ℎ(̇𝑞,𝐺)Δ𝐱,(49)whereℎ(̇𝑞,𝐺)=̇𝑞𝑇12̇𝐺𝑃𝜑−𝑆𝑃𝜑−𝐺̇𝑃𝜑𝐽𝑇−𝐺𝑃𝜑̇𝐽𝑇.(50)This 1×2-vector is quadratic in components of ̇𝑞, and each coefficient of ̇𝑞𝑖̇𝑞𝑗 (for 𝑖,𝑗=1,…,4) is at most of order of the maximum spectre radius of 𝐺(𝑞) denoted by 𝑔𝑀. Hence, it follows from (49) that||̇𝑞,𝐺Δ𝐱||≤𝑔𝑀𝑙0̇𝑞2‖Δ𝐱‖,(51)where 𝑙0 signifies a constant that is of order of the maximum link length (because the Jacobian matrix 𝜕𝐱/𝜕𝑞𝑇 is homogeneously related to the three link lengths of the robot shown in Figure 4). Further, note that ̇𝑞𝑇𝑃𝜑=0 in this case and remark thaṫ𝑞𝑇𝐺𝑃𝜑𝐽𝑇̇𝐱≤12̇𝑞𝑇𝐺𝐺̇𝑞+12̇𝐱𝑇𝐽𝑃𝜑𝐽𝑇̇𝐱,(52)−𝜁𝑘̇𝐱𝑇𝐽𝑃𝜑𝐽𝑇Δ𝐱≤𝜁𝑘2Δ𝐱𝑇𝐽𝑃𝜑𝐽𝑇Δ𝐱+𝜁2̇𝐱𝑇𝐽𝐽𝑇̇𝐱.(53)Thus, substituting (51) to (52) into (49) yieldṡ𝑉≤−𝑘−12𝜁𝑘2Δ𝐱𝑇𝐽𝑃𝜑𝐽𝑇Δ𝐱+12̇𝑞𝑇𝐺𝐺̇𝑞+1+𝜁2̇𝐱𝑇𝐽𝐽𝑇̇𝐱+𝑔𝑀𝑙0̇𝑞2‖Δ𝐱‖.(54)With the aid of a positive parameter 𝛼>0, it is easy to see that𝑊𝛼≥𝐸−𝛼2̇𝑞𝑇𝐺(𝑞)𝐺(𝑞)̇𝑞−𝛼2Δ𝐱𝑇𝐽𝑃𝜑𝐽𝑇Δ𝐱=12̇𝑞𝑇{𝐺−𝛼𝐺𝐺}̇𝑞+12Δ𝐱𝑇𝑘𝐼2−𝛼𝐽𝑃𝜑𝐽𝑇Δ𝐱.(55)

Now, suppose that the robot has an ordinary physical scale with𝑔𝑀≤0.001[kgm2],𝑙0≤0.2[m].(56)Then, it is possible to see thaṫ𝑊𝛼=̇𝐸+𝛼̇𝑉≤−𝛼𝑘𝜎𝑚2‖Δ𝐱‖2𝑐3̇𝑞2,(57)provided that ‖Δ𝐱(𝑡)‖<0.3[m], (1+𝜁𝑘)≤𝑘, and 𝜁𝑘>𝛼(1+𝜁)/2. From the practical point of view of control design for the handwriting robot, 𝜁 is set around 𝜁=2/2, and 𝑘 is chosen between 5.0∼20.0[kg/s2] (see [9]). On account of (48) and (56),𝑊𝛼1−𝛼𝛾0𝐸,(58)and similarly,𝑊𝛼1+𝛼𝛾0𝐸,(59)where 𝛾0=max(𝑔𝑀,𝜎𝑀/𝑘). Then, if the damping factor can be selected to satisfy both inequalities 𝑐≥(3/2)(𝛼𝜎𝑚𝑔𝑀) and 𝑐>𝛼𝑔𝑀𝑙0, then (57) is reduced tȯ𝑊𝛼−𝛼𝜎𝑚1+𝛼𝛾0𝑊𝛼,(60)where 𝛾0 can be set as 𝛾0=𝜎𝑀/𝑘 which is larger than 𝑔𝑀. Hence, by choosing 𝛼=1/2𝛾0=𝑘/2𝜎𝑀, it follows from (60) and (59) that𝑊𝛼(𝑡)=𝑊𝛼(0)𝑒−𝛾𝑡,(61)where 𝛾=𝑘𝜎𝑚/3𝜎𝑀.

Now, suppose that 𝑞 in EM1∩𝐵(𝑞(0),𝑟0) is the minimizing point that connects with 𝑞(0) among all geodesics from 𝑞(0) to any point of EM1∩𝐵(𝑞(0),𝑟0). We call 𝑞 a reference point corresponding to 𝑞(0).

Definition 1 (stable Riemannian ball on a submanifold). If for any 𝜀>0, there exists the number 𝛿(𝜀)>0 such that any solution trajectory (orbit) of (43) starting from an arbitrary initial position inside 𝐵(𝑞,𝛿(𝜀)) with ̇𝑞(0)=0 remains inside 𝐵(𝑞,𝜀) for any 𝑡>0, then the reference point 𝑞 on EM1 is said to be stable on a submanifold (see Figure 5).
It can be concluded from the exponential convergence of 𝑊𝛼 to zero and the inequality 𝐸≤2𝑊𝛼 when 𝛼=1/2𝛾0 that any point inside 𝐵(𝑞,𝛿(𝜀)) included in 𝐵(𝑞(0),𝑟1) for some 𝑟1(<𝑟0) is stable on a submanifold and further such a solution trajectory converges asymptotically to some 𝑞 on the equilibrium manifold EM1 in an exponential speed of convergence. This can be well understood as a natural extension of the well-known Dirichlet-Lagrange stability under holonomic constraints to a system with DOF-redundancy. The details of the proof are presented in Appendix .

It should be noticed from the proof in Appendix that asymptotic convergence of the solution trajectory to some 𝑞 on the equilibrium manifold implies also the asymptotic convergence of constraint force 𝜆(𝑡) to 𝜆𝑑 as 𝑡→∞ because Δ𝜆=𝜆−𝜆𝑑 is expressed asΔ𝜆(𝑡)=1𝑤𝑇𝐺−1𝑤𝑘̇𝑤̇𝑞𝑘𝑖,𝑗𝑤𝑘Γ𝑘𝑖𝑗̇𝑞𝑖̇𝑞𝑗−𝑤𝑇𝐽𝑇𝜁𝑘̇𝐱+𝑘Δ𝐱,(62)and this right hand side converges to zero as 𝑡→∞.

The stability notion of a Riemannian ball in a neighborhood of a reference equilibrium state 𝑞 on EM1 is extended to cope with the case that the initial velocity vector ̇𝑞(0) is not zero. To do this, we define an extended Riemannian ball in the tangent bundle 𝑀×𝑇𝐹𝜉 around (𝑞,̇𝑞=0) in such a way that𝐵𝑞,0;𝑟0,𝑟𝐾=𝑞,̇𝑞∶𝑑𝑞,𝑞<𝑟0,(1/2)̇𝑞𝑇𝐺(𝑞)̇𝑞<𝑟𝐾,(63)where 𝑑(𝑞,𝑞) denotes the distance between 𝑞 and 𝑞 restricted to the submanifold 𝐹𝜉, and 𝐺 is defined below (40).

Definition 2 (asymptotic stability on a submanifold). If for any 𝜀>0, there exist numbers 𝛿(𝜀) and 𝛿𝐾(𝜀) such that any solution trajectory of (43) starting from an arbitrary initial position and velocity inside 𝐵{(𝑞,0);(𝛿(𝜀),𝛿𝐾(𝜀))} remains in 𝐵{(𝑞,0);(𝜀,𝑟𝐾)} and further converges asymptotically to some equilibrium point 𝑞∈EM1 with still state, then the reference point with its posture on EM1 is said to be asymptotically stable on a constraint submanifold.

It should be remarked that Bloch et al. [10] introduces originally the concept of stabilization for a class of nonholonomic dynamic systems based upon a certain configuration space. The redefinition of stability concepts introduced above is free from any choice of configuration spaces (local coordinates) and assumptions on an invertibility condition (that is almost equivalent to nonlinear control based on compensation for nonlinear inertia-originated terms). Liu and Li [11] also gave a geometric approach to modeling of constrained mechanical systems based upon orthogonal projection maps without deriving a compact explicit form of the Euler equation like (40) with a reduced dimension due to constraints. Therefore, the proposed control scheme was developed on the basis of compensation for the inertia-originated nonlinear terms (that is almost equivalent to the computed torque method). A naive idea of stability on a manifold by using different metrics for the constrained submanifold and its tangent space was first presented in [9] and used in stabilization control of robotic systems with DOF redundancy [12, 13].

6. 2-dimensional Stable Grasp of a Rigid Object with Arbitrary Shape

Consider a control problem for stable grasping of a 2D rigid object by a pair of planar multijoint robot fingers with hemispherical fingertips as shown in Figure 6. In this figure, the two robots are installed on the horizontal 𝑥𝑦-plane 𝐄2. We denote the object mass center by 𝑂𝑚 with the coordinates (𝑥𝑚,𝑦𝑚) expressed in the inertial frame. On the other hand, we express a local coordinate system fixed at the object by 𝑂𝑚-𝑋𝑌 together with unit vectors 𝐫𝑋 and 𝐫𝑌 along the 𝑋-axis and 𝑌-axis, respectively (see Figure 7). The left-hand side surface of the object is expressed by a curve 𝑐(𝑠) with local coordinates (𝑋(𝑠),𝑌(𝑠)) in terms of arc length parameter 𝑠 as shown in Figure 7.

First, suppose that at the left-hand contact point 𝑃1 the fingertip of the left finger is contacting with the object. Denote the unit normal at 𝑃1 by 𝐧1 and the unit tangent vector by 𝐞1. Note that 𝐧1 is normal to both the object surface and finger-end sphere at 𝑃1, and 𝐞1 is tangent to them at 𝑃1, too. If we denote position 𝑃1 by local coordinates (𝑋(𝑠),𝑌(𝑠)) fixed at the object (see Figure 7), then the angle from the 𝑋-axis to the unit normal 𝐧1 is assumed to be determined by a function on the curve:𝜃1(𝑠)=arct𝑋(𝑠)𝑌(𝑠),(64)where 𝑋(𝑠)=d𝑋(𝑠)/d𝑠 and 𝑌′=d𝑌(𝑠)/d𝑠. In this paper, all angles are set positive in counterclockwise direction. Then,𝑃1𝑃1=𝑙𝑛1(𝑠)=−𝑋(𝑠)cos𝜃1(𝑠)+𝑌(𝑠)sin𝜃1(𝑠),(65)which is dependent only on 𝑠 and, therefore, a shape function of the object. On the other hand, this length can be expressed by using the inertial frame coordinates in the following way (see Figure 8):𝑃1𝑃1=𝑥𝑚−𝑥01cos𝜃+𝜃1(𝑠)𝑦𝑚−𝑦01sin𝜃+𝜃1(𝑠)−𝑟1.(66)Hence, the left-hand contact constraint can be expressed by the holonomic constraint𝑄1=−𝑥𝑚−𝑥01cos𝜃+𝜃1+𝑦𝑚−𝑦01sin𝜃+𝜃1+𝑟1+𝑙𝑛1(𝑠)=0,(67)where 𝑙𝑛1(𝑠) denotes the right-hand side of (65) and 𝜃1=𝜃1(𝑠) for abbreviation.

Next, we note that the length 𝑂𝑚𝑃1 can be regarded also as a shape function of the object given by𝑂𝑚𝑃1=𝑙𝑒1(𝑠)=𝑋(𝑠)sin𝜃1+𝑌(𝑠)cos𝜃1.(68)On the other hand, this quantity can be also expressed as𝑂𝑚𝑃1=𝑌1(𝑡)=−𝑥𝑚−𝑥01sin𝜃+𝜃1𝑦𝑚−𝑦01cos𝜃+𝜃1.(69)As discussed in [13], in the light of the book [14], pure rolling at 𝑃1 is defined by the condition that the translational velocity of 𝑃1 on the finger sphere is equal to that of 𝑃1 on the object contour along the 𝐞1-axis. Hence, by denoting 𝑝1=𝑞11+𝑞12+𝑞13, we have−𝑟1̇𝑝1+𝑟1̇𝜃=−̇𝑌1=̇𝑥𝑚−̇𝑥01sin𝜃+𝜃1+̇𝑦𝑚−̇𝑦01cos𝜃+𝜃1+𝑥𝑚−𝑥01cos𝜃+𝜃1𝑦𝑚−𝑦01sin𝜃+𝜃1)̇𝜃,(70)which from (67) can be reduced to−𝑟1̇𝑝1−𝑙𝑛1̇𝜃+̇𝑥01−̇𝑥sin𝜃+𝜃1+̇𝑦01−̇𝑦cos𝜃+𝜃1=0.(71)This constraint form is Pfaffian. As to the contact point 𝑃2 at the right-hand finger end, a similar nonholonomic constraint can be obtained. Thus, by introducing Lagrange's multipliers 𝑓1 and 𝑓2 associated with holonomic constraints 𝑄𝑖=0(𝑖=1,2), it is possible to construct a Lagrangian:𝐿=12̇𝑞𝑇𝑖𝐺𝑖𝑞𝑖̇𝑞𝑖+12𝑀̇𝑥2+𝑀̇𝑦2+𝐼̇𝜃2−𝑓1𝑄1−𝑓2𝑄2,(72)where 𝑞𝑖 denote the joint vector for finger 𝑖, 𝐺𝑖(𝑞𝑖) the inertia matrix for finger 𝑖, 𝑀 and 𝐼 denote the mass and inertia moment of the object. Since both the rolling constraints are Pfaffian, it is possible to associate (71) and its corresponding form at 𝑃2 with another multipliers 𝜆𝑖(𝑖=1,2) and regard them as external forces. Thus, by applying the variational principle to the Lagrangian together with the external forces, we obtain the Lagrange equation of motion of the overall fingers-object system:𝐼̈𝜃−𝑓1𝑌1+𝑓2𝑌2−𝜆1𝑙𝑛1+𝜆2𝑙𝑛2=0,(73)𝑀̈𝑥̈𝑦−𝑓1𝐧1−𝑓2𝐧2−𝜆1𝐞1−𝜆2𝐞2=0,(74)𝐺𝑖𝑞𝑖̈𝑞𝑖+12̇𝐺𝑖+𝑆𝑖̇𝑞𝑖+𝑓𝑖𝐽𝑇𝑖𝑞𝑖𝐧𝑖+𝜆𝑖𝐽𝑇𝑖𝑞𝑖𝐞𝑖+(−1)𝑖𝑟𝑖𝐩𝑖=𝑢𝑖,𝑖=1,2,(75) where 𝐩1=(1,1)𝑇, 𝐩2=(1,1,1)𝑇, and 𝐽𝑇𝑖(𝑞𝑖)=𝜕(𝑥0𝑖/𝑦0𝑖)/𝜕𝑞𝑖, and𝐧𝑖=(−1)𝑖−cos𝜃+𝜃𝑖sin𝜃+𝜃𝑖,𝐞𝑖=sin𝜃+𝜃𝑖cos𝜃+𝜃𝑖(76)for 𝑖=1,2. It is important to note that the object dynamics of (73) and (74) can be recast in the form𝐻0̈𝐳+𝑓1𝐰1+𝑓2𝐰2+𝜆1𝐰3+𝜆2𝐰4=0,(77)where 𝐳=(𝑥,𝑦,𝜃), and𝐰𝑖=−𝐧𝑖(−1)𝑖𝑌𝑖,𝐰𝑖+2=−𝐞𝑖(−1)𝑖𝑙𝑛𝑖,𝑖=1,2,(78)which are of a two-dimensional wrench vector. This implies that if the sum of all wrench vectors converges to zero, then the force/torque balance is established. Further, if we define𝐗=𝑥,𝑦,𝜃,𝑞𝑇1,𝑞𝑇2𝑇,𝐮=0,0,0,𝑢𝑇1,𝑢𝑇2𝑇,𝐺(𝐗)=diag𝑀,𝑀,𝐼,𝐺1𝑞1,𝐺2𝑞2,𝑆(𝐗,̇𝐗)=diag0,0,0,𝑆1,𝑆2,(79)then (73) (75) can be written in the form𝐺(𝐗)̈𝐗+12̇𝐺+𝑆̇𝐗+𝑖=1,2𝑓𝑖𝜕𝑄𝑖𝜕𝐗+𝜆𝑖𝜕𝑅𝑖𝜕𝐗=𝐮,(80)where𝑅𝑖=−(−1)𝑖𝑟𝑖𝜃+𝜃𝑖𝑠𝑖−𝐩𝑇𝑖𝑞𝑖+𝑌𝑖+𝜓𝑖𝑠𝑖,(81)and 𝜓𝑖 signifies a function depending only on the shape parameter 𝑠𝑖. The details of the integral form of (81) will be discussed in Appendix . Comparing (80) with (27), it must be understood that a similar but extended argument of Section 4 can be applied even for a case of physical interaction with complex holonomic constraints and nonholonomic (but Pfaffian) constraints. More precisely, when 𝐮=0 in (80), the image space of the Riemannian manifold (𝐗,𝐺(𝐗)) with constraints 𝑄1=0 and 𝑄2=0 is spanned from the gradients𝐺−1(𝐗)𝜕𝑄1𝜕𝐗,𝜕𝑄2𝜕𝐗,𝜕𝑅1𝜕𝐗,𝜕𝑅2𝜕𝐗,(82)and the kernel space should be defined as the orthogonal compliment to the image space. Note that the rolling constraints of (71) and another for 𝑖=2 induce further restriction of the tangent space to a more subdimensional linear space. Here, 𝜕𝑅𝑖/𝜕𝐗 should be taken at the time instant 𝑡 under the condition that 𝑠𝑖 are fixed. Nevertheless, as shown in Appendix , it is quite interesting to know that 𝑅𝑖 for 𝑖=1,2 can be regarded as a constant function in change of 𝑡 even under any infinitesimally small variation of 𝑠𝑖, that is, d𝑅𝑖/d𝑡=0 for 𝑖=1,2. In other words, 𝑅𝑖=constant(𝑖=1,2), and these expressions imply a holonomic constraint. Such a geometric structure of decomposition of the tangent space to an image space of constraint gradients and its orthogonal compliment is known in general mathematical terminology (see [15, 16]) but has not yet been critically spelled out in the case of existence of rolling contact connections between curved rigid bodies.

The Euler equation of (80) can be expressed in a similar form to (1) by introducing the constraint vector Φ and the vector of Lagrange multipliers 𝝀 asΦ=𝑄1,𝑅1,𝑄2,𝑅2,𝝀=𝑓1,𝜆1,𝑓2,𝜆2𝑇.(83)Then, (80) can be written in the form𝐺(𝐗)̈𝑋+12̇𝐺+𝑆̇𝐗+𝜕Φ𝜕𝐗𝝀=𝐮,(84)which must be valid for the constraint ̇𝐗𝑇(𝜕Φ/𝜕𝐗)=0. We denote the (𝑛−4)-dimensional kernel space of 𝜕Φ/𝜕𝐗 by 𝑉𝐗 and its 4-dimensional orthogonal compliment as the image space of 𝐺−1(𝑋)𝜕Φ/𝜕𝐗 by 𝐻𝐗.

We are now in a position to derive an update law of the length parameters 𝑠1 and 𝑠2 along the object contours. Firstly, it is important to verify that the holonomic constraint 𝑄1=0 is invariant under any infinitesimally small variation of 𝑠1. In fact, we see from (64) to (69) that𝜕𝑄1𝜕𝑠=𝑥𝑚−𝑥01sin𝜃+𝜃1+𝑦𝑚−𝑦01cos𝜃+𝜃1×𝜕𝜃1𝜕𝑠+𝜕𝜕𝑠𝑙𝑛1(𝑠)=−𝑙𝑒1(𝑠)𝜕𝜃1(𝑠)𝜕𝑠+−𝑋(𝑠)cos𝜃1+𝑌(𝑠)sin𝜃1+𝑋(𝑠)sin𝜃1+𝑌(𝑠)cos𝜃1𝜕𝜃1(𝑠)𝜕𝑠=−𝑙𝑒1(𝑠)𝜕𝜃1𝜕𝑠+0+𝑙𝑒1(𝑠)𝜕𝜃1𝜕𝑠=0,(85)where we denote 𝑠1=𝑠 for simplicity. Next, it is possible to show that if 𝜃1(𝑠) is variable, then 𝜃1(𝑠)=d𝜃1(𝑠)/d𝑠=𝜅1(𝑠), where 𝜅1(𝑠) is the curvature of the left-hand contour of the object. Then, it is obvious from Figures 6 to 8 that𝑟1−Δ𝑝1+Δ𝜃+Δ𝜃1Δ𝑡=−Δ𝑠=−𝜅−11(𝑠)Δ𝜃1Δ𝑡,(86)where 𝜅1(𝑠)≠0. Since 𝜃1(𝑠)=𝜅1(𝑠), (86) is reduced to𝑟1−Δ𝑝1+Δ𝜃Δ𝑡=−1+𝑟1𝜅1(𝑠)Δ𝑠Δ𝑡.(87)This equation should be accompanied with the Euler-Lagrange equation (80). This can be also regarded as a nonholonomic constraint to (80). As to the right-hand contour, it follows thatd𝑠2d𝑡=−𝑟21+𝑟2𝜅2𝑠2−̇𝑝2+̇𝜃.(88)

7. Lifting in Horizontal Space and Force/Torque Balance

First, in order to find an adequate lifting that belongs to the image space 𝐻𝐗 and realizes the force/torque balance (see (77)) in the sense that𝑓1𝑑𝐰1+𝑓2𝑑𝐰2+𝜆1𝑑𝐰3+𝜆2𝑑𝐰4=0,(89) we remark that𝑥𝑚−𝑥01𝑦𝑚−𝑦01=𝑅𝜃+𝜃1𝑟1+𝑙𝑛1𝑠1−𝑌1(90)𝑥𝑚−𝑥02𝑦𝑚−𝑦02=𝑅𝜃+𝜃2𝑙𝑛𝑠2+𝑟2𝑌2,(91)where𝑅𝜃+𝜃1=𝐧1(𝜃),𝐞1(𝜃)=cos𝜃+𝜃1)sin𝜃+𝜃1−sin𝜃+𝜃1)cos𝜃+𝜃2.(92)We also define𝑅𝜃+𝜃2=cos𝜃+𝜃2sin𝜃+𝜃2−sin𝜃+𝜃2cos𝜃+𝜃2,𝑅𝜃=cos𝜃sin𝜃−sin𝜃cos𝜃.(93)Then, it follows from (90) and (91) that𝑥01−𝑥02𝑦01−𝑦02=𝑅𝜃𝑅𝜃1𝑟1+𝑙𝑛1𝑠1−𝑌1+𝑅𝜃2𝑟2+𝑙𝑛2𝑠2𝑌2=𝑅𝜃𝑙𝐬,𝑌1,𝑌2−𝑑𝐬,𝑌1,𝑌2,(94)where 𝐬=(𝑠1,𝑠2) and𝑙𝐬,𝑌1,𝑌2=𝑟1+𝑙𝑛1cos𝜃1+𝑟2+𝑙𝑛2cos𝜃2−𝑌1sin𝜃1+𝑌2sin𝜃2,𝑑𝐬,𝑌1,𝑌2=𝑟1+𝑙𝑛1sin𝜃1+𝑟2+𝑙𝑛2sin𝜃2+𝑌1cos𝜃1−𝑌2cos𝜃2.(95)Thus, let us define𝑓𝑖𝑑=𝑓𝑑𝑟1+𝑟2𝑙cos𝜃𝑖+𝑑sin𝜃𝑖,𝜆𝑖𝑑=(−1)𝑖𝑓𝑑𝑟1+𝑟2−𝑙sin𝜃𝑖+𝑑cos𝜃𝑖,𝑖=1,2,(96)and remark that they satisfy𝑓1𝑑𝐧1+𝑓2𝑑𝐧2+𝜆1𝑑𝐞1+𝜆2𝑑𝐞2=0.(97)Further, note that𝑓1𝑑𝐰1+𝑓2𝑑𝐰2+𝜆1𝑑𝐰3+𝜆2𝑑𝐰4=(0,0,𝑆𝑁)𝑇,(98)where𝑆𝑁=𝑓𝑑𝑟1+𝑟2𝑙cos𝜃1+𝑑sin𝜃1𝑌1+𝑙cos𝜃2+𝑑sin𝜃2𝑌2𝑙sin𝜃1−𝑑cos𝜃1𝑙𝑛1𝑙sin𝜃2−𝑑cos𝜃2𝑙𝑛2=𝑓𝑑𝑟1+𝑟2{𝑙𝑟1sin𝜃1+𝑟2sin𝜃2−𝑑𝑟1cos𝜃1+𝑟2cos𝜃2.(99)This shows according to (98) that if 𝑆𝑁 tends to vanish, then the force/torque balance is established, that is, the total sum of wrench vectors exerted to the object becomes zero.

8. Control Signal for Blind Grasping and a Morse-Lyapunov Function

From the practical standpoint of designing a control signal for stable grasping, it is important to see that objects to be grasped are changeable, but the pair of robot fingers is always the same. That is, for designing control signals, we are unable to use physical parameters of the object such as the location (𝑥𝑚,𝑦𝑚) of its mass center and object geometry. On the contrary, it is possible to assume the knowledge of finger kinematics like finger link lengths and locations of the centers of finger-end spheres and to use measurement data of finger joint angles and angular velocities. In view of these points, let us propose a family of control signals defined as𝑢𝑖=−𝑐𝑖̇𝑞𝑖+(−1)𝑖𝑓𝑑𝑟1+𝑟2𝐽𝑇𝑖𝑞𝑖𝑥01−𝑥02𝑦01−𝑦02−𝑟𝑖𝑁𝑖𝐩𝑖,𝑖=1,2,(100)where𝑁𝑖(𝑡)=𝛾−1𝑖𝑟𝑖𝐩𝑇𝑖𝑞𝑖(𝑡)−𝑞𝑖(0),(101)and 𝐩1=(1,1,1)𝑇, 𝐩2=(1,1)𝑇, 𝑐𝑖 denotes a damping factor, and 𝛾𝑖>0 a positive gain specified later.

The closed loop dynamics of motion of the overall fingers-object system can be derived by substituting 𝑢𝑖 of (100) into (75) for 𝑖=1,2. In order to spell out the dynamics in a more physically meaningful way for later discussions, first note the following three equalities:12𝑥01−𝑥022+𝑦01−𝑦022=12𝑙2+𝑑2𝑓𝑖𝑑𝜕𝑄𝑖𝜕𝑞𝑖+𝜆𝑖𝑑𝜕𝑅𝑖𝜕𝑞𝑖=(−1)𝑖𝐽𝑇𝑖−𝑓𝑖𝑑cos𝜃+𝜃𝑖−sin𝜃+𝜃𝑖+𝜆𝑖𝑑sin𝜃+𝜃𝑖cos𝜃+𝜃𝑖+𝜆𝑖𝑑𝑟𝑖𝐩𝑖=(−1)𝑖𝑓𝑑𝑟1+𝑟2−𝐽𝑇𝑖𝑅𝜃𝑙−𝑑+𝜆𝑖𝑑𝑟𝑖𝐩𝑖=(−1)𝑖𝑓𝑑𝑟1+𝑟2𝐽𝑇𝑖𝑥01−𝑥02𝑦01−𝑦02+𝜆𝑖𝑑𝑟𝑖𝐩𝑖,𝑖=1,2,(102)Then, by substituting (98), (102) into (73) to (75), it is possible to express the closed loop dynamics in the following forms:𝐼̈𝜃−Δ𝑓1𝑌1+Δ𝑓2𝑌2−Δ𝜆1𝑙𝑛1+Δ𝜆2𝑙𝑛2−𝑆𝑁=0,𝑀̈𝑥̈𝑦+Δ𝑓1𝐧1(𝜃)+Δ𝑓2𝐧2(𝜃)−Δ𝜆1𝐞1(𝜃)−Δ𝜆2𝐞2(𝜃)=0,𝐺𝑖̈𝑞𝑖+12̇𝐺𝑖+𝑆𝑖̇𝑞𝑖+𝑐𝑖̇𝑞𝑖+Δ𝑓𝑖𝐽𝑇𝑖𝐧𝑖(𝜃)+Δ𝜆𝑖𝐽𝑇𝑖𝐞𝑖(𝜃)+(−1)𝑖𝑟𝑖𝐩𝑖+𝑟𝑖(−1)𝑖𝜆𝑖𝑑+𝑁𝑖𝐩𝑖=0,𝑖=1,2.(103) Similarly, to the form of (84), these equations can be written in the following way:𝐺(𝐗)̈𝐗+12̇𝐺+𝑆+𝐶̇𝐗+𝜕Φ𝜕𝐗Δ𝝀+𝑆𝑁𝐛𝜃+𝑖=1,2𝑁𝑖+(−1)𝑖𝜆𝑖𝑑𝐛𝑖=0,(104)whereΔ𝝀=𝑓1−𝑓1𝑑,𝑓2−𝑓2𝑑,𝜆1−𝜆1𝑑,𝜆2−𝜆2𝑑𝑇,𝐛𝜃=(0,0,1,0,0,0,0,0)𝑇,𝐛1=(0,0,0,1,1,1,0,0)𝑇,𝐛2=(0,0,0,0,0,0,1,1)𝑇.(105)At this stage, it is important to note that in accordance with four constraints Φ=0, the velocity vector ̇𝐗 belongs to the kernel space of (𝜕Φ/𝜕𝐗)𝑇 and, therefore, ̇𝐗𝑇𝜕Φ/𝜕𝐗=0. Further by using (100) and taking inner product of (84) and ̇𝐗, we obtaindd𝑡𝐾+𝑓𝑑2𝑟1+𝑟2𝑥01−𝑥022+𝑦01−𝑦022+𝑖=1,2𝛾𝑖2𝑁2𝑖=−𝑖=1,2𝑐𝑖̇𝑞𝑖2,(106)where 𝐾 denotes the system's kinetic energy defined by𝐾=𝑖=1,212̇𝑞𝑇𝑖𝐺𝑖𝑞𝑖̇𝑞𝑖+𝑀2̇𝑥2+̇𝑦2+𝐼2̇𝜃2.(107)The relation of (106) must be equivalently derived by taking inner product of ̇𝐗 and the closed loop dynamics of equation (104). To verify this, let us define𝑝1=𝑞11+𝑞12+𝑞13=𝐩𝑇1𝑞1,𝑝2=𝑞21+𝑞22=𝐩𝑇2𝑞2,(108)𝑃=𝑓𝑑2𝑟1+𝑟2𝑙2𝐬,𝑌1,𝑌2+𝑑2𝐬,𝑌1,𝑌2+𝑖=1,2𝛾𝑖2𝑁2𝑖,(109)where 𝑙 and 𝑠 are defined in (95). It should be remarked that, with the aid of expressions of integral form of rolling constraints shown in (71), we have𝜕𝑙𝜕𝜃=−𝜕𝑌1𝜕𝜃sin𝜃1+𝜕𝑌2𝜕𝜃sin𝜃2=−𝑟1sin𝜃1−𝑟2sin𝜃2,𝜕𝑑𝜕𝜃=𝑟1cos𝜃1+𝑟2cos𝜃2,(110)and hence,𝜕𝑃𝜕𝜃=𝑓𝑑𝑟1+𝑟2𝑙𝜕𝑙𝜕𝜃+𝑑𝜕𝑑𝜕𝜃=−𝑓𝑑𝑟1+𝑟2𝑙𝑟1sin𝜃1+𝑟2sin𝜃2−𝑑𝑟1cos𝜃1+𝑟2cos𝜃2=𝑆𝑁.(111)Similarly, from (66), (69), and (101), we have𝜕𝑃𝜕𝑝𝑖=𝑓𝑑𝑟1+𝑟2𝑙𝜕𝑙𝜕𝑝𝑖+𝑑𝜕𝑑𝜕𝑝𝑖+𝛾𝑖𝑁𝑖𝜕𝑁𝑖𝜕𝑝𝑖=(−1)𝑖𝑓𝑑𝑟1+𝑟2𝑙𝜕𝑌𝑖𝜕𝑝𝑖sin𝜃𝑖−𝑑𝜕𝑌𝑖𝜕𝑝𝑖cos𝜃𝑖+𝛾𝑖𝑁𝑖𝜕𝑁𝑖𝜕𝑝𝑖=−𝑟𝑖𝑓𝑑𝑟1+𝑟2𝑙sin𝜃𝑖−𝑑cos𝜃𝑖+𝑟𝑖𝑁𝑖=(−1)𝑖𝑟𝑖𝜆𝑖𝑑+𝑟𝑖𝑁𝑖,𝑖=1,2.(112)Thus, the inner product of ̇𝐗 and (104) or the relation of (106) is reduced todd𝑡{𝐾+𝑃}=−𝑖=1,2𝑐𝑖̇𝑞𝑖2.(113)In other words, the closed loop dynamics of (104) can be expressed in the form𝐺(𝐗)̈𝐗+12̇𝐺+𝑆+𝐶̇𝐗+𝜕𝑃𝜕𝐗+𝜕Φ𝜕𝐗Δ𝜆=0.(114)This is interpreted as a Lagrange equation of the Lagrangian𝐿=𝐾−𝑃−ΦΔ𝝀,(115)in accompany with the external damping torques 𝑐𝑖̇𝑞𝑖 for 𝑖=1,2 through finger joints. The scalar function 𝑃 defined by (109) is a quadratic function of 𝑌1, 𝑌2, 𝑝1, and 𝑝2, and hence, it is regarded as a quadratic function of 𝜃, 𝑝1, and 𝑝2 since 𝑌𝑖 can be regarded as a linear function of 𝜃 and 𝑝𝑖 for 𝑖=1,2 because of (71). Hence, 𝑃 can be regarded as a Morse function defined on the Riemannian submanifold induced by two constraints described by 𝑄𝑖=0 for 𝑖=1,2 and four constraints in the tangent space described by ̇𝑄𝑖 and ̇𝑅𝑖=0 for 𝑖=1,2.

9. Physical Insights into Gradient and Hessian of the Morse Function

The physical meaning of control signals for blind grasping defined by (100) is quite simple. The first term of the right-hand side of (100) plays a role of damping for rotational motion of finger joints. Damping for motion of the object is exerted from velocity constraints ̇𝑄𝑖=0 and ̇𝑅𝑖=0 for 𝑖=1,2 as discussed in detail in the previous paper [1]. The second term plays a role of fingers-thumb opposition that induces minimization of the distance between 𝑂01 and 𝑂02, centers of finger-end spheres. The distance is equivalent to 𝑙2+𝑑2 as discussed in Section 5. The third term plays an important role in suppressing excess movements of rotation of finger joints. These characteristics of the control signal condense into the Lagrange equation of motion with (1) the potential 𝑃(𝐬,𝜃,𝑝1,𝑝2) of (109), (2) the lifting (𝜕Φ/𝜕𝐗)𝝀𝑑, where 𝝀𝑑=(𝑓1𝑑,𝑓2𝑑,𝜆1𝑑,𝜆2𝑑)𝑇, and (3) the gradient 𝜕𝑃/𝜕𝐗 of the potential. In other words, (113) implies that if the artificial potential 𝑃 attains its minimum and, at the same time, makes the gradient 𝜕𝑃/𝜕𝐗 vanish at some 𝐬=𝐬(𝐬=(𝑠1,𝑠2)𝑇), and, moreover, the artificial potential is positive definite in 𝐗 under the two holonomic constraints, then the equilibrium position that minimizes 𝑃 would be asymptotically stable. Unfortunately, 𝑃 is a quadratic function with respect to only 𝜃, 𝑝1, and 𝑝2, and, therefore, 𝑃 is only nonnegative definite in 𝐗. Nevertheless, it is possible to show that the Hessian matrix of 𝑃 with respect to 𝜃, 𝑝1, and 𝑝2 becomes positive definite in these three variables as shown in Table 1 that is calculated by partially differentiating the gradients 𝜕𝑃/𝜕𝜃, 𝜕𝑃/𝜕𝑝1, and 𝜕𝑃/𝜕𝑝2 of (111) to (112) with respect to 𝜃, 𝑝1, and 𝑝2 again. Apparently, the Hessian becomes positive definite provided that 𝛾𝑖 is chosen as being of similar order of (𝑟𝑖/𝑓𝑑), and 𝜃𝑖(𝑠𝑖) remains in a region such that |𝜃𝑖(𝑠𝑖)|<𝜋/6 for 𝑖=1,2. Then, by using a similar argument used in proving the stability on a submanifold for DOF redundant systems [9, 13], a solution trajectory to the Lagrange equation converges exponentially to the constraint equilibrium manifold. In this paper, we omit the details of the argument, but show a physical meaning of the gradient 𝜕𝑃/𝜃 together with 𝜕𝑃/𝜕𝑝𝑖(𝑖=1,2).

As shown in Figure 9, all rotational motions of the object emerge around axes that are perpendicular to the 𝑥𝑦-plane. Here, we consider the rotational moment that may emerge at the contact point 𝑃2 exerted by the pressing force 𝐟1 to the object from the other contact point 𝑃1. Another force 𝐟2 at 𝑃2 that presses the object generates the torque around the 𝑧-axis at 𝑃1. The sum of these torques around the 𝑧-axis can be expressed as𝐫1×𝐟1+𝐫2×𝐟2=𝑃1𝑃2×𝑟1𝑓𝑑𝑟1+𝑟2𝐧1+𝑃2𝑃1×𝑟2𝑓𝑑𝑟1+𝑟2𝐧2=𝑋𝑠1−𝑋𝑠2𝑌𝑠1−𝑌𝑠20×𝑟1𝑓𝑑𝑟1+𝑟2cos𝜃1−sin𝜃10+𝑋𝑠2−𝑋𝑠1𝑌𝑠2−𝑌𝑠10×𝑟2𝑓𝑑𝑟1+𝑟2−cos𝜃2sin𝜃20=𝑓𝑑𝑟1+𝑟200𝑋1−𝑋2𝑟1sin𝜃1+𝑟2sin𝜃2𝑌1−𝑌2𝑟1cos𝜃1+𝑟2cos𝜃2,(116) where we denote 𝑋𝑖=𝑋(𝑠𝑖) and 𝑌𝑖=𝑌(𝑠𝑖) for abbreviation. We see also that from geometric relations of the vectors 𝑂𝑚𝑃𝑖, and quantities 𝑌𝑖 and 𝑙𝑛𝑖 expressed in local coordinates 𝑂𝑚-𝑋𝑌 (see Figure 9), it follows that𝑙−𝑑=𝑟1cos𝜃1−sin𝜃1+𝑟2cos𝜃2−sin𝜃2+𝑅𝜃1𝑙𝑛1−𝑌1+𝑅𝜃2𝑙𝑛2−𝑌2=𝑟1cos𝜃1−sin𝜃1+𝑟2cos𝜃2−sin𝜃2𝑋𝑠1−𝑋𝑠2𝑌𝑠1−𝑌𝑠2.(117)Applying this relation to (116), we see that𝐫1×𝐟1+𝐫2×𝐟2=𝑓𝑑𝑟1+𝑟20,0,𝑙𝑖=1,2𝑟𝑖sin𝜃𝑖−𝑑𝑖=1,2𝑟𝑖cos𝜃𝑖𝑇=0,0,𝑆𝑁𝑇.(118)

As a summary of the argument, it is possible to conclude that a solution trajectory to (114) converges asymptotically to some equilibrium state 𝐗=𝐗 with some 𝐬=𝐬 with 𝜕𝑃/𝜕𝐗=0 at 𝐗=𝐗 and 𝐬=𝐬.

10. Conclusions

A natural extension of hybrid position/force control for robots with redundant degrees of freedom is presented from the standpoint of a Riemannian geometric approach. It is shown that any supply of the constant pressing force lying in the image space of the constraint gradient to the environment through joint actuations is not relevant to motions in the kernel space orthogonally compliment to the image space. An extension of problems of grasping and manipulation of rigid objects to the case of 2-dimensional objects with arbitrary shape is also treated from the viewpoint of Dirichlet-Lagrange's stability by introducing a nonnegative definite Morse-Lyapunov function on a Riemannian manifold together with damping shaping.

Appendices

A. A Proof of Stability

From exponential decay of 𝑊𝛼(𝑡) as 𝑡→∞ shown in (61) and (58) and (59) (where 𝛼 is set as 𝛼=1/2𝛾0), it follows that𝐸(𝑡)≤𝑒−𝛾𝑡1−𝛼𝛾0𝑊𝛼(0)1+𝛼𝛾01−𝛼𝛾0𝑒−𝛾𝑡𝐸(0)=3𝑒−𝛾𝑡𝐸(0).(A.1)Hence, 𝐸(𝑡) is also exponentially convergent to zero. Since ̇𝑞(𝑡) is also exponentially convergent in 𝑡, 𝑞(𝑡) converges to some 𝑞 that belongs to the equilibrium manifold EM1. Along the trajectory, the Riemannian distance 𝐿(𝑞(0),𝑞(𝑇)) must satisfy𝐿𝑞(0),𝑞(𝑇)𝑇0̇𝑞𝑇(𝑡)𝐺𝑞(𝑡)̇𝑞(𝑡)d𝑡,(A.2)and it is possible to suppose that 𝐿(𝑞(0),𝑞(𝑇)) converges to 𝐿(𝑞(0),𝑞) as 𝑇→∞. On the other hand, for any 𝑇>0, it follows from (A.1) that𝑇0̇𝑞𝑇(𝑡)𝐺𝑞(𝑡)̇𝑞(𝑡)d𝑡≤𝑇02𝐸(𝑡)d𝑡6𝐸(0)2𝛾.(A.3) Hence, substituting this into (A.2) yields 𝑅𝑞(0),𝑞(𝑇)6𝐸(0)2𝛾.(A.4)In particular, we have𝐿𝑞(0),𝑞=6𝐸(0)2𝛾.(A.5)Thus, it follows that𝐿𝑞,𝑞(𝑇)≤𝐿𝑞(0),𝑞(𝑇)+𝐿𝑞(0),𝑞6𝐸(0)2𝛾+𝐿𝑞(0),𝑞.(A.6)First, we consider the case that at initial time 𝑡=0 the angular velocity vector is zero, that is, ̇𝑞(0)=0. Then, since 𝑞inEM1 and, therefore, 𝐱(𝑞)=𝐱𝑑 and 𝐽(𝑞) is nondegenerate in a local coordinate chart containing 𝑞, there exists a number 𝛿0(𝜀)>0 such that any 𝑞(0) satisfying𝐿𝑞(0),𝑞<𝛿0(𝜀)(A.7)implies𝐱(𝑞(0))−𝐱𝑑23𝑘𝛾𝜀2.(A.8)Now, let us define𝛿(𝜀)=min𝜀2,𝛿0(𝜀).(A.9)Then, for any 𝑞(0) satisfying 𝐿(𝑞(0),𝑞)<𝜀/2, it follows from (A.6) and (A.8) that𝐿𝑞,𝑞(𝑇)𝜀2+𝐿𝑞(0),𝑞<𝜀,(A.10)which completes the proof of stability of the reference point 𝑞 on EM1.

B. Integrability of Rolling Constraints

Suppose that the contact point 𝑃1 on the left-hand finger end with the object contour moves along the finger end circle starting from the initial time 𝑡=0 to the present time 𝑡=𝑡 (Figure 6). The length of its movement can be expressed by𝑠1(𝑡)−𝑠1(0)=𝑟1𝜑1(𝑡)−𝜑1(0)}=𝑟1𝜃(𝑡)+𝜃1𝑠1(𝑡)−𝑝1(𝑡)𝜃(0)+𝜃1𝑠1(0)−𝑝1(0),(B.1)that is equivalent to the formula of (86). We show that 𝜓𝑖(𝑠𝑖) for 𝑖=1,2 can be chosen as𝜓𝑖𝑠𝑖=(−1)𝑖𝑙𝑒𝑖𝑠𝑖+𝑠𝑖,𝑖=1,2.(B.2)In fact, let us define in accordance with (81) and (B.2)𝑅𝑖=𝑟𝑖𝜃+𝜃𝑖𝑠𝑖−𝑝𝑖−(−1)𝑖𝑌𝑖−𝑙𝑒𝑖𝑠𝑖+𝑠𝑖(B.3)for 𝑖=1,2 and take the derivative of them in 𝑡. Then, it follows thatd𝑅𝑖d𝑡=𝑟𝑖̇𝜃−̇𝑝𝑖−(−1)𝑖𝜕𝑌𝑖𝜕𝑞𝑇𝑖̇𝑞𝑖+𝜕𝑌𝑖𝜕𝐱𝑇̇𝐱+𝜕𝑌𝑖𝜕𝜃̇𝜃+𝜕𝑟𝑖𝜃𝑖−(−1)𝑖𝑌𝑖−𝑙𝑒𝑖𝜕𝑠𝑖d𝑠𝑖d𝑡+d𝑠𝑖d𝑡.(B.4)Note that from the meanings of (64) to (69), it follows that𝜕𝑙𝑒𝑖𝜕𝑠𝑖=−(−1)𝑖𝑋𝑠𝑖sin𝜃𝑖+𝑌𝑠𝑖)cos𝜃𝑖+𝑋𝑠𝑖cos𝜃𝑖−𝑌𝑠𝑖sin𝜃𝑖𝜕𝜃𝑖𝜕𝑠𝑖=(−1)𝑖𝑙𝑛𝑖𝑠𝑖𝜕𝜃𝑖𝜕𝑠𝑖−(−1)𝑖,𝜕𝑌𝑖𝜕𝑠𝑖=(−1)𝑖𝑥𝑚−𝑥0𝑖cos𝜃+𝜃𝑖+𝑦𝑚−𝑦0𝑖sin𝜃+𝜃𝑖𝜕𝜃𝑖𝜕𝑠𝑖=(−1)𝑖𝑟𝑖+𝑙𝑛𝑖𝑠𝑖𝜕𝜃𝑖𝜕𝑠𝑖.(B.5)Substituting these two equations into (B.4) and noticing equality (70), we have now shown thatd𝑅𝑖d𝑡=0,𝑖=1,2.(B.6)

C. Frenet-Serret Form

It should be remarked that, instead of specifying 𝜃1(𝑠1) and 𝜃2(𝑠2) through (64), it is possible to use the moving coordinates frame called the Frenet frame defined by the pair of unit vectors (𝐞1(𝑠),𝐧1(𝑠)) expressed by local coordinates 𝑂𝑚-𝑋𝑌. Then, evolution of the frame (𝐞1(𝑠),𝐧1(𝑠)) along the curve length must be determined by the Frenet-Serret formuladd𝑠𝐞1(𝑠),𝐧1(𝑠)=𝐞1(𝑠),𝐧1(𝑠)0−𝜅1(𝑠)𝜅1(𝑠)0,(C.1)where 𝜅1(𝑠) denotes the curvature of the left-hand object contour at arc length 𝑠. On the other hand, these two unit vectors are expressed in terms of the inertial frame 𝑂-𝑥𝑦 by the formula𝐧1(𝜃)=𝑅𝜃𝐧1(𝑠)=𝐫𝑋,𝐫𝑌𝐧1(𝑠)=cos𝜃sin𝜃−sin𝜃cos𝜃cos𝜃1(𝑠)−sin𝜃1(𝑠)=cos𝜃+𝜃1−sin𝜃+𝜃1,,𝐞1(𝜃)=𝑅𝜃𝐞1(𝑠)=sin𝜃+𝜃1cos𝜃+𝜃1,(C.2) where 𝑅𝜃∈𝑆𝑂(2) and it is subject todd𝑡𝑅𝜃=𝑅𝜃0𝜔𝑧−𝜔𝑧0,(C.3)where 𝜔𝑧=̇𝜃. Then, it is evident that𝑙𝑛1(𝑠)𝐧1(𝜃)−𝑙𝑒1(𝑠)𝐞1(𝜃)=𝑥𝑚−𝑥1𝑦𝑚−𝑦1,𝑥1𝑦1=𝑥01𝑦01+𝑟1𝐧1(𝑠).(C.4) From these equalities, it follows that𝑙𝑛1(𝑠)=𝐧𝑇1(𝜃)𝑥𝑚−𝑥01𝑦𝑚−𝑦01−𝑟1,𝑙𝑒1(𝑠)=−𝐞𝑇1(𝜃)𝑥𝑚−𝑥01𝑦𝑚−𝑦01=𝑌1(𝑡),(C.5)which show equalities (66) and (68). Hence, (74) that expresses translational motion of the object can be written in the Frenet form 𝑀̈𝑥̈𝑦−𝑓1𝐧1(𝜃)−𝜆1𝐞1(𝜃)−𝑓2𝐧2(𝜃)−𝜆2𝐞2(𝜃).(C.6)Note that at the right-hand contact point the Frenet form is defined by (−𝐞2(𝑠),𝐧2(𝑠)), and hence the Frenet-Serret formula is written asdd𝑠𝐞2,−𝐧2=𝐞2,−𝐧20−𝜅2(𝑠)𝜅2(𝑠)0.(C.7)