Lagrange-d'Alembert Spark Integrators For Nonholonomic Lagrangian Systems

1y ago
12 Views
1 Downloads
1.94 MB
25 Pages
Last View : 6d ago
Last Download : 3m ago
Upload by : Ciara Libby
Transcription

LAGRANGE-D’ALEMBERT SPARK INTEGRATORS FOR NONHOLONOMIC LAGRANGIAN SYSTEMS LAURENT O. JAY Abstract. We consider Lagrangian systems with ideal nonholonomic constraints. These systems can be expressed as implicit index 2 differential-algebraic equations (DAEs) and can be derived from the Lagrange-d’Alembert principle. Methods based on a discrete Lagrange-d’Alembert principle are called Lagrange-d’Alembert integrators and they generalize variational integrators. We define a new nonholonomically constrained discrete Lagrange-d’Alembert principle based on a discrete Lagranged’Alembert principle for forced Lagrangian systems. The principle that we propose does not make explicit use of any Lagrange multiplier in its formulation. Nonholonomic constraints are considered as first integrals of the underlying forced Lagrangian system of ordinary differential equations. We show that a large class of specialized partitioned additive Runge-Kutta (SPARK) methods for index 2 DAEs satisfies the new discrete principle. Symmetric Lagrange-d’Alembert SPARK integrators of any order can be obtained based for example on Gauss and Lobatto coefficients as already proposed for more general index 2 DAEs. Our results are illustrated by several numerical experiments. Key words. Differential-algebraic equations, discrete mechanics, forcing, Gauss coefficients, ideal constraints, index 2, Lagrange-d’Alembert principle, Lagrangian systems, Lobatto coefficients, nonholonomic constraints, SPARK methods. AMS subject classifications. 65L05, 65L06, 65L80, 70F25, 70G45, 70H03, 70H45. 1. Introduction. In this paper we consider the numerical solution of Lagrangian systems with ideal nonholonomic constraints. Nonholonomic systems in mechanics have a long and intriguing history [1, 5, 28]. Nonholonomic constraints involve velocities and are nonintegrable, i.e., they cannot be derived from holonomic constraints. The dynamics of nonholonomic systems has been the subject of a controversy between Lagrange-d’Alembert mechanics and vakonomic (variational nonholonomic) mechanics. It is nowadays accepted that vakonomic mechanics does not lead to the correct equations of motion of physical systems, but that Lagrange-d’Alembert mechanics generally does [5, 22, 28]. In this paper we consider methods which mimic faithfully at the discrete level the integral Lagrange-d’Alembert principle. Such methods are called Lagrange-d’Alembert (LDA) integrators. They fall under the framework of geometric integration methods and they generalize variational integrators [19, 24]. Geometric integration has attracted quite a lot of interest in recent years, see for example the book [13] and the survey paper [27]. Geometric integration methods can be classified as extrinsic or intrinsic. Intrinsic methods are coordinate-free methods, for example Lie group methods are defined intrinsically in terms of the exponential map or some approximation to it on the corresponding Lie algebra to advance the numerical solution in time. In this paper we will exclusively consider extrinsic methods. Extrinsic methods consider an embedding of the manifold in Rn and make use of coordinates. For unconstrained Hamiltonian and Lagrangian systems important classes of geometric integrators are symplectic/Poisson integrators [13, 14, 21, 31] and variational integrators which are based on a discrete version of Hamilton’s principle [19, 24]. For unconstrained Lagrangian systems with forcing an important class of geometric integrators are Lagrange-d’Alembert (LDA) integrators which are based on a discrete version of the Lagrange-d’Alembert principle [19, 24]. Department of Mathematics, 14 MacLean Hall, The University of Iowa, Iowa City, IA 522421419, USA. E-mail: ljay@math.uiowa.edu and na.ljay@na-net.ornl.gov. This material is based upon work supported by the National Science Foundation under Grant No. 0654044. 1

2 L. O. Jay For ideal nonholonomic constraints the equations of motion do not derive from a standard variational principle, but for example from the Lagrange-d’Alembert principle [3, 4, 5, 20, 28] which is not a variational principle, but a differential principle. For Lagrangian systems and ideal scleronomic (i.e., time-independent) linear nonholonomic constraints in the velocities the Lagrange-d’Alembert principle is equivalent to a skew critical problem which is also not truly a variational principle [3, 7, 8, 25]. Methods based on a similar discrete skew critical problem in Q Q where Q is the configuration space have been first introduced by Cortés in [6, 7] and have also been called Lagrange-d’Alembert (LDA) integrators. A few low order LDA integrators of this type have been developed in [11, 12, 25]. We believe that the discrete constrained Lagrange-d’Alembert principle of Cortés [6, 7] is not sufficiently general to include many methods of interest. Tentatives to extend this principle have been made by Cuell and Patrick. For Lagrangian systems with ideal scleronomic linear nonholonomic constraints these authors have shown the Lagrange-d’Alembert principle to be equivalent to a skew critical problem in the kinematic state space T Q [8, 9, 10]. This result has been extended to Lagrangian systems with ideal nonlinear scleronomic nonholonomic constraints [29]. For Lagrangian systems with ideal linear scleronomic nonholonomic constraints, they have proposed discrete skew critical methods defined directly in the kinematic state space T Q [8, 10]. We also mention the energy-preserving methods for Lagrangian systems with ideal linear scleronomic nonholonomic constraints considered by Betsch in [2]. In this paper we take a radically different and simpler approach. We quote from McLachlan and Perlmutter in [25]: much work remains to be done to clarify the nature of discrete nonholonomic mechanics and to pinpoint the “correct” discrete analog of the Lagrange-d’Alembert principle. The nonholonomically constrained discrete Lagrange-d’Alembert principle that we consider in this paper is certainly correct and sufficiently general to include many methods of interest. We define a general constrained discrete Lagrange-d’Alembert principle directly in Q Q based on the discrete Lagrange-d’Alembert principle for forced Lagrangian systems proposed in [19], see [24, Section 3.2]. The principle that we propose does not make explicit use of any Lagrange multiplier in its formulation contrary to [6, 7, 8, 10, 11, 12, 25]. This is consistent with the fact that the Lagrange-d’Alembert principle for nonholonomic systems is not a variational principle. We remark that nonholonomic constraints can be mathematically realized with forcing as a forced Lagrangian system. The nonholonomically constrained discrete Lagrange-d’Alembert principle that we propose is therefore fully consistent with the unconstrained discrete Lagrange-d’Alembert principle for forced Lagrangian systems. It generalizes the one proposed by Cortés in [6, 7] which appears restrictive. It is an extension in a direction awaited by McLachlan and Perlmutter in [25, Section 8]. This extension was in fact partly suggested without details by Marsden and West in [24, Section 5.3.7]. A large class of specialized partitioned additive Runge-Kutta (SPARK) methods for index 2 DAEs is shown to satisfy the new discrete principle. Symmetric Lagrange-d’Alembert SPARK integrators of any order can be obtained based for example on Gauss and Lobatto coefficients as already proposed for more general index 2 DAEs in [15, 16, 17]. An extension of the results of this paper to submanifolds Q Rn and holonomic constraints will be the subject of a forthcoming paper [18]. The paper is organized as follows. In section 2 the system of DAEs of Lagrangian systems with ideal nonholonomic constraints is given. The underlying forced Lagrangian system is also obtained. In section 3 the Lagrange-d’Alembert principle is

Lagrange-d’Alembert integrators for nonholonomic systems 3 discussed. A forced discrete Lagrange-d’Alembert principle for Lagrangian systems with nonholonomic constraints is proposed in section 4. In section 5 the exact discrete forcing terms for Lagrangian systems with nonholonomic constraints are derived. In section 6 the main Theorem 6.1 gives sufficient conditions for SPARK methods to satisfy the forced discrete Lagrange-d’Alembert principle for Lagrangian systems with nonholonomic constraints. Several examples of Lagrange-d’Alembert SPARK integrators are given in section 7. In section 8 some numerical experiments are given to illustrate the favorable energy preservation property of Lagrange-d’Alembert SPARK integrators. Finally, a short conclusion is given in section 9. 2. Lagrangian systems with ideal nonholonomic constraints. For simplicity in this paper we suppose that the configuration space Q is the linear space Q Rn . The constrained Lagrangian system with Lagrangian L : R T Q R (where T Q Rn Rn ) and ideal nonholonomic constraints k : R T Q Rm (m n) is given by the Lagrange equations of the second kind (2.1a) (2.1b) (2.1c) d q v, dt d v L(t, q, v) q L(t, q, v) K(t, q, v)T ψ, dt 0 k(t, q, v), where (2.1d) K(t, q, v) : kv (t, q, v). In most applications the nonholonomic constraints (2.1c) are affine in the generalized velocities v, i.e., (2.2) 0 k(t, q, v) K(t, q)v b(t, q). Moreover, such ideal affine nonholonomic constraints (2.2) are oftentimes just linear in v, i.e., b(t, q) 0. The assumption (2.2) will actually not be needed in this paper. 2.1. The underlying forced Lagrangian system. Expanding the left-hand side of (2.1b) we get (2.3a) 2vv L(t, q, v) d v K(t, q, v)T ψ 2tv L(t, q, v) 2qv L(t, q, v)v q L(t, q, v). dt From a computational point of view, see Section 6, it is in fact advantageous to consider directly the formulation (2.1b) instead of (2.3a) since (2.3a) requires the calculation of the extra terms 2tv L(t, q, v) and 2qv L(q, v)v, the latter corresponding to Coriolis forces. Differentiating (2.1c) once with respect to t and using (2.1a) we obtain (2.3b) K(t, q, v) d v kt (t, q, v) kq (t, q, v)v. dt In this paper we assume that the matrix 2 vv L(t, q, v) K(t, q, v)T (2.4) K(t, q, v) O is nonsingular.

4 L. O. Jay For example, for ideal nonholonomic constraints one can assume that K(t, q, v) is of full row rank m and that the Lagrangian L is regular, i.e., the Hessian matrix (2.5) 2vv L(t, q, v) is nonsingular, 2vv L(t, q, v) is generally assumed to be positive definite. Under the assumption (2.4), d from (2.3) we can express dt v and ψ as explicit functions of (t, q, v). Hence, under the assumption (2.4) the equations (2.1) are implicit differential-algebraic equations (DAEs) of index 2 [16]. For consistent initial values (q0 , v0 ) at t0 , i.e., such that 0 k(t0 , q0 , v0 ), assuming (2.4) and sufficient smoothness of L and k, we have existence and uniqueness of a solution (q(t), v(t), ψ(t)) to (2.1). Expressing ψ as an implicit function of (t, q, v), i.e., ψ Ψ(t, q, v), we obtain from (2.1ab) the underlying forced Lagrangian system (2.6a) (2.6b) d q v, dt d v L(t, q, v) q L(t, q, v) fL (t, q, v), dt where (2.7) fL (t, q, v) : K(t, q, v)T Ψ(t, q, v) can be interpreted as a forcing term. This corresponds to a mathematical realization of the nonholonomic constraints (2.1c). By construction the functions k(t, q, v) of (2.1c) d are first integrals of the forced Lagrangian system (2.6)-(2.7) since dt k(t, q, v) 0 by definition of Ψ(t, q, v). 2.2. Energy. The energy of the system (2.1) is defined as (2.8) E(t, q, v) : Lv (t, q, v)v L(t, q, v). We have d d E(t, q, v) Lv (t, q, v) v Lv (t, q, v)v̇ Lt (t, q, v) Lq (t, q, v)v Lv (t, q, v)v̇ dt dt Lq (t, q, v)v ψ T K(t, q, v)v Lt (t, q, v) Lq (t, q, v)v ψ T K(t, q, v)v Lt (t, q, v). For ideal scleronomic (time-independent) nonholonomic constraints linear in v 0 K(q)v and time-independent Lagrangians L(t, q, v) L(q, v) the energy is conserved since K(q)v 0 and Lt (q, v) 0. 3. The Lagrange-d’Alembert principle. For ideal affine nonholonomic constraints (2.2) the equations (2.1) can be derived from the Lagrange-d’Alembert principle [3, 4, 5, 20, 28] which is a differential principle. The Lagrange-d’Alembert principle states that the virtual work vanishes d (3.1a) δ W : Lv (t, q, q̇) Lq (t, q, q̇) δ q 0 dt

Lagrange-d’Alembert integrators for nonholonomic systems 5 for all reversible (i.e., with q in the interior of the configuration space Q) virtual displacements δ q satisfying (3.1b) K(t, q)δ q 0. Notice nevertheless that energy can still be created or dissipated along a trajectory of (2.1), see subsection 2.2 above. The definition of ideal nonholonomic constraints is equivalent to the Lagrange-d’Alembert principle (3.1ab) with (3.1b) for ideal nonlinear nonholonomic constraints simply replaced by (3.1c) K(t, q, q̇)δ q 0. This is the Maurer-Appell-Chetaev-Johnsen-Hamel rule, see e.g. [28, p. 820], usually simply called Chetaev’s rule. From (3.1c) we obtain the expression (2.1d) in (2.1b). For ideal nonholonomic constraints a different equivalent and finite-dimensional variational principle leading to (2.1) is Gauss’ principle of least constraint [20, 28, 30], but it is based on the generalized accelerations q̈ and is thus generally seen as an inferior principle. 3.1. The Lagrange-d’Alembert principle as a skew critical problem. For ideal linear scleronomic nonholonomic constraints 0 K(q)q̇, the Lagranged’Alembert principle is equivalent to a skew critical problem [3, 7, 8, 9, 10, 25, 29] described as follows. Given a Lagrangian L(t, q, q̇) C 0 ([t0 , tN ], T Q) and ideal linear scleronomic nonholonomic constraints 0 K(q)q̇ we form the action integral between q0 at t0 and qN at tN Z tN L(t, q(t), q̇(t))dt A(q) : t0 1 which is a functional for q C ([t0 , tN ], Q) satisfying q(t0 ) q0 , q(tN ) qN , and 0 K(q(t))q̇(t). The Lagrange-d’Alembert principle is then equivalent to the skew critical problem (3.2) δA(q)(δq) 0 δq C01 ([t0 , tN ], Q) K(q)δq 0 where δA(q) is the first variation (i.e., the Gâteaux derivative) of the action. Hamilton’s variational principle is not valid in the presence of ideal nonholonomic constraints contrary to ideal holonomic constraints. Observe that in (3.2) we do not have the seemingly more natural condition 1 Kq (q)(q̇, δq) K(q)δq 0. lim K(q εδq)(q̇ εδq) ε ε 0 Hamilton’s principle applied to problems with ideal nonholonomic constraints leads to the generally different vakonomic equations [1, 5, 28] which do not agree with physical experiments [22]. It is worth mentioning that the practical realization of nonholonomic constraints is a problem in itself which may be quite difficult [23, 32]. 3.2. The integral Lagrange-d’Alembert principle for forced Lagrangian systems and for Lagrangian systems with nonholonomic constraints. For forced Lagrangian systems (2.6) the (continuous) integral Lagrange-d’Alembert principle is Z tN fL (t, q(t), q̇(t))T δq(t)dt 0 δq C01 ([t0 , tN ], Q). δA(q)(δq) t0

6 L. O. Jay For systems with nonholonomic constraints fL is given by (2.7) and we can simply add the conditions 0 k(t, q, q̇). This principle is simpler and more general than the skew critical problem of subsection 3.1. 3.3. Nonideal constraints. The constraints (2.1c) are called nonideal when the Lagrange-d’Alembert principle (3.1) does not hold. For example, Chetaev’s rule (3.1c) may be unsuitable in certain situations, see, e.g., [23]. For Lagrangian systems with nonideal nonholonomic constraints (2.1b) is replaced by d v L(t, q, v) q L(t, q, v) K(t, q, v)T ψ N (t, q, v, ψ) dt with Nψ 6 0. For example dry sliding friction can lead to such formulations, see, e.g., [28, Example 3.2.6]. Even holonomic constraints can be nonideal, see, e.g., [33, 34]. Notice that the SPARK methods of Section 6 can deal with systems having nonideal constraints without any particular difficulty. 4. A forced discrete Lagrange-d’Alembert principle for Lagrangian systems with nonholonomic constraints. For Lagrangian systems with nonholonomic constraints we define in this section a general constrained discrete Lagranged’Alembert principle directly in Q Q based on the discrete Lagrange-d’Alembert principle for forced Lagrangian systems. 4.1. The forced discrete Lagrange-d’Alembert principle and Euler-Lagrange equations. For forced Lagrangian systems (2.6) the corresponding forced discrete Lagrange-d’Alembert principle proposed in [19], see [24, Section 3.2], is (4.1) δ N 1 X Ld (tk , qk , tk 1 , qk 1 ) k 0 N 1 X k 0 fd (tk , qk , tk 1 , qk 1 )T δqk fd (tk , qk , tk 1 , qk 1 )T δqk 1 0 n for all variations {δqk }N k 0 with δqk R satisfying δq0 0 δqN . The discrete Lagrangian Ld (tk , qk , tk 1 , qk 1 ) (or local discrete action) is an approximation to the exact discrete Lagrangian (the exact local action) between tk and tk 1 Z tk 1 E L(t, q(t), q̇(t))dt Ld (tk , qk , tk 1 , qk 1 ) Ld (tk , qk , tk 1 , qk 1 ) : tk where q(t) : q(t, tk , qk , tk 1 , qk 1 ). The discrete forces fd (tk , qk , tk 1 , qk 1 ) and fd (tk , qk , tk 1 , qk 1 ) above are approximation to the exact discrete forces between tk and tk 1 (4.2a) fd (tk , qk , tk 1 , qk 1 )T fdE (tk , qk , tk 1 , qk 1 )T Z tk 1 fL (t, q(t), q̇(t))T qk q(t)dt, : tk (4.2b) fd (tk , qk , tk 1 , qk 1 )T fdE (tk , qk , tk 1 , qk 1 )T Z tk 1 fL (t, q(t), q̇(t))T qk 1 q(t)dt. : tk

Lagrange-d’Alembert integrators for nonholonomic systems 7 The discrete principle (4.1) is equivalent to the forced discrete Euler-Lagrange equations (4.3) 4 Ld (tk 1 , qk 1 , tk , qk ) 2 Ld (tk , qk , tk 1 , qk 1 ) fd (tk 1 , qk 1 , tk , qk ) fd (tk , qk , tk 1 , qk 1 ) 0 for k 1, . . . , N 1. These equations (4.3) define a mapping ( R Q R Q R Q R Q, Φ: (tk 1 , qk 1 , tk , qk ) 7 (tk , qk , tk 1 , qk 1 ). From q(t, tk 1 , qk 1 , tk , qk ) q(t, tk , qk , tk 1 , qk 1 ) we have the anti-symmetry properties E LE d (tk 1 , qk 1 , tk , qk ) Ld (tk , qk , tk 1 , qk 1 ), E E fd (tk 1 , qk 1 , tk , qk )) fd (tk , qk , tk 1 , qk 1 ), fdE (tk 1 , qk 1 , tk , qk )) fdE (tk , qk , tk 1 , qk 1 ). Hence, from these properties we could require Ld (tk 1 , qk 1 , tk , qk ) Ld (tk , qk , tk 1 , qk 1 ), fd (tk 1 , qk 1 , tk , qk )) fd (tk , qk , tk 1 , qk 1 ), fd (tk 1 , qk 1 , tk , qk )) fd (tk , qk , tk 1 , qk 1 ), as part of the conditions of the forced discrete Lagrange-d’Alembert principle (4.1). This makes sense from a boundary value problem point of view, but this is not fully justified from an initial value problem point of view. For an initial value problem, we are only interested in integrating in a specific time t direction and nonsymmetric methods may also be appropriate when the forced Lagrangian system (2.6) has no symmetry or reversibility properties. 4.2. The nonholonomically constrained discrete Lagrange-d’Alembert principle and Euler-Lagrange equations. For Lagrangian systems with nonholonomic constraints, the integral Lagrange-d’Alembert principle for Lagrangian systems with nonholonomic constraints stated in subsection 3.2 and the forced discrete Lagrange-d’Alembert principle (4.1) motivate the following definition: Definition 4.1. For Lagrangian systems with ideal nonholonomic constraints (2.1) we define the nonholonomically constrained discrete Lagrange-d’Alembert principle as (4.4a) δ N 1 X Ld (tk , qk , tk 1 , qk 1 ) k 0 N 1 X k 0 fd (tk , qk , tk 1 , qk 1 )T δqk fd (tk , qk , tk 1 , qk 1 )T δqk 1 0, (4.4b) 0 c(tk , qk , tk 1 , qk 1 ) for k 0, . . . , N 1, for all variations {δqk }N k 0 satisfying δq0 0 δqN , with fd and fd as in (4.2), fL as in (2.7), and c(tk , qk , tk 1 , qk 1 ) : k(tk 1 , qk 1 , u(tk 1 , tk , qk , tk 1 , qk 1 ))

8 L. O. Jay with u(t, tk , qk , tk 1 , qk 1 ) d q(t, tk , qk , tk 1 , qk 1 ). dt We assume that c(t, q, t, q) 0 t R q Q. The nonholonomically constrained discrete Lagrange-d’Alembert principle of Definition 4.1 is equivalent to the nonholonomically constrained discrete Euler-Lagrange equations (4.5a) 4 Ld (tk 1 , qk 1 , tk , qk ) 2 Ld (tk , qk , tk 1 , qk 1 ) fd (tk 1 , qk 1 , tk , qk ) fd (tk , qk , tk 1 , qk 1 ) 0, (4.5b) c(tk 1 , qk 1 , tk , qk ) c(tk , qk , tk 1 , qk 1 ) for k 1, . . . , N 1 where we assume that c(t0 , q0 , t1 , q1 ) 0 or c(tN 1 , qN 1 , tN , qN ) 0. Notice that (4.5b) implies (4.4b). Assuming c(t0 , q0 , t1 , q1 ) 0 the equations (4.5) define a mapping ( Cd Cd , Φ: (tk 1 , qk 1 , tk , qk ) 7 (tk , qk , tk 1 , qk 1 ). on the constraint submanifold Cd : {(s, q, t, r) R Q R Q c(s, q, t, r) 0} . We remark that ideal nonholonomic constraints can be mathematically realized with forcing (2.7) as a forced Lagrangian system (2.6). The nonholonomically constrained discrete Lagrange-d’Alembert principle that we propose here is fully consistent with the unconstrained forced discrete Lagrange-d’Alembert principle (4.1) for forced Lagrangian systems as briefly suggested by Mardsen and West in [24, Section 5.3.7]. It also generalizes the one proposed by Cortés in [6, 7] which appears restrictive. It is an extension in a direction awaited by McLachlan and Perlmutter in [25, Section 8]. 5. The exact discrete forcing terms for Lagrangian systems with nonholonomic constraints. Consider a solution to (2.1) q(t) q(t, t0 , q0 , t1 , q1 ) passing through q0 at t0 and q1 at t1 and let v(t) v(t, t0 , q0 , t1 , q1 ) : d q(t, t0 , q0 , t1 , q1 ) t q(t, t0 , q0 , t1 , q1 ). dt Consider the exact discrete Lagrangian (the exact local action) as a function of (t0 , q0 , t1 , q1 ) LE d (t0 , q0 , t1 , q1 ) : Z t1 L(t, q(t), v(t))dt. t0 Notice that for unconstrained systems S1 (q0 , q1 ) : LE d (t0 , q0 , t1 , q1 ) can play the role of a generating function of type I. We denote v0 : v(t0 ), v1 : v(t1 ), p0 : v L(t0 , q0 , v0 ), p1 : v L(t1 , q1 , v1 ).

9 Lagrange-d’Alembert integrators for nonholonomic systems We have q0 LE d (t0 , q0 , t1 , q1 ) Z t1 Lq (t, q(t), v(t)) q0 q(t) Lv (t, q(t), v(t)) q0 v(t)dt t0 Z t1 t0 Lq (t, q(t), v(t)) q0 q(t) d Lv (t, q(t), v(t)) q0 q(t)dt dt t Lv (t, q(t), v(t)) q0 q(t) t10 Z t1 d Lq (t, q(t), v(t)) Lv (t, q(t), v(t)) q0 q(t)dt dt t0 Lq (t1 , q1 , v1 ) q0 q1 Lv (t0 , q0 , v0 ) q0 q0 Z t1 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) q0 q(t)dt pT0 . t0 Defining fdE (t0 , q0 , t1 , q1 ) : p0 q0 LE d (t0 , q0 , t1 , q1 ), we have obtained E T T q0 LE d (t0 , q0 , t1 , q1 ) p0 fd (t0 , q0 , t1 , q1 ) (5.1) where fdE (t0 , q0 , t1 , q1 )T Z t1 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) q0 q(t)dt. t0 From the Fundamental Theorem of Calculus we have Z t1 Z t d v(s)ds q(s)ds q1 q(t) q1 t t1 ds leading to q0 q(t) Z t1 q0 v(s)ds. t Hence, fdE (t0 , q0 , t1 , q1 )T Z t1 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) t0 t1 Z t0 Z t1 t0 t1 (5.2) Z t0 Z t t1 T s t0 s Z t1 q0 v(s)ds dt Ψ(t, q(t), v(t)) K(t, q(t), v(t)) q0 v(s)ds dt t Z Z t0 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) q0 v(s)dt ds Ψ(t, q(t), v(t))T K(t, q(t), v(t))dt q0 v(s)ds. Similarly, we obtain q1 LE d (t0 , q0 , t1 , q1 ) pT1 Z t1 t0 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) q1 q(t)dt.

10 L. O. Jay Defining fdE (t0 , q0 , t1 , q1 ) : p1 q1 LE d (t0 , q0 , t1 , q1 ), we have E T T q1 LE d (t0 , q0 , t1 , q1 ) p1 fd (t0 , q0 , t1 , q1 ) . (5.3) We obtain fdE (t0 , q0 , t1 , q1 )T Z t1 Ψ(t, q(t), v(t))T K(t, q(t), v(t)) q1 q(t)dt t0 and q1 q(t) Z t q1 v(s)ds t0 leading to (5.4) fdE (t0 , q0 , t1 , q1 )T Z t1 t0 Z t1 s Ψ(t, q(t), v(t))T K(t, q(t), v(t))dt q1 v(s)ds. We assume that the values tk are independent of the values qj . For example one considers a constant stepsize h and the values tk : t0 kh for k 0, . . . , N. We calculate for k 1, . . . , N 1 qk N 1 X j 0 E E LE d (tj , qj , tj 1 , qj 1 ) qk Ld (tk 1 , qk 1 , tk , qk ) Ld (tk , qk , tk 1 , qk 1 ) E 4 LE d (tk 1 , qk 1 , tk , qk ) 2 Ld (tk , qk , tk 1 , qk 1 ) pk fdE (tk 1 , qk 1 , tk , qk ) pk fdE (tk , qk , tk 1 , qk 1 ) fdE (tk 1 , qk 1 , tk , qk ) fdE (tk , qk , tk 1 , qk 1 ). This leads to E 4 LE d (tk 1 , qk 1 , tk , qk ) 2 Ld (tk , qk , tk 1 , qk 1 ) fdE (tk 1 , qk 1 , tk , qk ) fdE (tk , qk , tk 1 , qk 1 ) 0 where fdE (tk 1 , qk 1 , tk , qk )T (5.5a) fdE (tk , qk , tk 1 , qk 1 )T (5.5b) Z Z tk Ψ(t, q(t), v(t))T K(t, q(t), v(t))T qk q(t)dt tk 1 Z tk Z tk 1 Z tk 1 tk tk 1 tk tk T Ψ(t, q(t), v(t)) K(t, q(t), v(t))dt qk v(s)ds, s Ψ(t, q(t), v(t))T K(t, q(t), v(t))T qk q(t)dt Z s tk T Ψ(t, q(t), v(t)) K(t, q(t), v(t))dt qk v(s)ds.

Lagrange-d’Alembert integrators for nonholonomic systems 11 5.1. Ideal holonomic constraints. When the constraints (2.1c) are holonomic, we have 0 g(t, q) for a certain function g : R Rn Rm . Therefore, we get 0 q0 g(t, q(t)) gq (t, q(t)) q0 q(t). In this situation, since 0 gt (t, q) gq (t, q)v : k(t, q, v), we have kv (t, q, v) gq (t, q), hence we must have 0 kv (t, q(t), v(t)) q1 q(t). 0 kv (t, q(t), v(t)) q0 q(t), For ideal holonomic constraints we have K(t, q, v) kv (t, q, v), and since kv (t, q, v) gq (t, q) we obtain q1 LE d (t0 , q0 , t1 , q1 ) p1 , q0 LE d (t0 , q0 , t1 , q1 ) p0 , and thus fdE 0, fdE 0. 6. Lagrange-d’Alembert SPARK integrators. Following [15, 16, 17] the application of an s-stage SPARK method to Lagrangian systems (2.1) with nonholonomic constraints, stepsize h, and consistent initial values (t0 , q0 , v0 ) at t0 , i.e., 0 k(t0 , q0 , v0 ), can be expressed as (6.1a) (6.1b) (6.1c) (6.1d) (6.1e) Qi q0 h Pi p0 h q1 q0 h p1 p0 h 0 s X s X j 1 s X aij Vj b aij Fj h j 1 s X bj Vj , j 1 bbj Fj h j 1 s X ωij Kj s X j 1 s X j 1 e aij Rj for i 1, . . . , s, ebj Rj , for i 1, . . . , s 1, j 1 (6.1f) for i 1, . . . , s, 0 k(t1 , q1 , v1 ), where t1 : t0 h and Ti : t0 ci h, Pi : v L(Ti , Qi , Vi ), Fi : q L(Ti , Qi , Vi ), Ri : K(Ti , Qi , Vi )T Ψi , Ki : k(Ti , Qi , Vi ) for i 1, . . . , s, p0 : v L(t0 , q0 , v0 ), p1 : v L(t1 , q1 , v1 ).

12 L. O. Jay The coefficients ωij in (6.1e) can be taken for example as ωij : bj cji 1 for i 1, . . . , s 1, j 1, . . . , s. Under certain assumptions on the coefficients of the SPARK method we obtain a mapping (t1 , q1 , v1 ) Φh (t0 , q0 , v0 ) for h sufficiently small [15, 16, 17]. Instead of considering the unknown quantities in equations (6.1) as implicit functions of (t0 , q0 , v0 , h) for h t1 t0 , we consider them as implicit functions of (t0 , q0 , t1 , q1 ). More precisely, we implicitly define by (6.1) as functions of (t0 , q0 , t1 , q1 ) the quantities v0 , v1 , p0 , p1 , Qi , Vi , Ψi , Pi , Fi , Ri , Ki for i 1, . . . , s. The main result of this paper is as follows: Theorem 6.1. For Lagrangian systems with nonholonomic constraints (2.1) and a corresponding s-stage SPARK method (6.1), suppose t0 , q0 and t1 , q1 to be given. If the SPARK coefficients satisfy bbi bi (6.2a) (6.2b) for i 1, . . . , s, bbi aij bj b aji bbi bj 0 for i, j 1, . . . , s, then we have a nonholonomically constrained discrete Lagrange-d’Alembert integrator in the sense of Definition 4.1 with Ld (t0 , q0 , t1 , q1 ) h s X bi L(Ti , Qi , Vi ), i 1 fd (t0 , q0 , t1 , q1 )T h fd (t0 , q0 , t1 , q1 )T h s X i 1 s X i 1 s X (ebj e aij )ΨTj K(Tj , Qj , Vj ) q1 Vi , bi h bi h j 1 s X j 1 e aij ΨTj K(Tj , Qj , Vj ) q0 Vi . Suppose in addition that the SPARK coefficients satisfy the symmetry conditions (6.3a) (6.3b) (6.3c) (6.3d) cs 1 i ci 1 for i 1, . . . , s, as 1 i,s 1 j aij bs 1 j bj b as 1 i,s 1 j b aij bbs 1 j bbj e as 1 i,s 1 j e aij ebs 1 j ebj for i, j 1, . . . , s, for i, j 1, . . . , s, for i, j 1, . . . , s then the SPARK method (6.1) is symmetric and we have Ld (t1 , q1 , t0 , q0 ) Ld (t0 , q0 , t1 , q1 ), fd (t1 , q1 , t0 , q0 ) fd (t0 , q0 , t1 , q1 ), fd (t1 , q1 , t0 , q0 ) fd (t0 , q0 , t1 , q1 ). Proof. We calculate q0 Ld (t0 , q0 , t1 , q1 ) h s X bi Lq (Qi , Vi ) q0 Qi h i 1 bi FiT I h i 1 h s X s X bi Lv (Qi , Vi ) q0 Vi i 1 s X j 1 aij q0 Vj h s X i 1 bi PiT q0 Vi

13 Lagrange-d’Alembert integrators for nonholonomic systems h s X bi FiT I h2 i 1 h s X i 1 h s X s X s X i 1 j 1 bi pT0 h bj FjT I h2 s X j 1 b aij FjT h s X s X i 1 j 1 j 1 s X pT0 h bi aij FiT q0 Vj bi q0 Vi h2 s X From (6.1c) we have 0 I h s X j 1 e aij RjT q0 Vi (bj aji bi b aij )FjT q0 Vi i 1 i 1 s X s X e aij RjT q0 Vi . bi j 1 bi q0 Vi , i 1 hence s X s X q0 Ld (t0 , q0 , t1 , q1 ) h2 i 1 j 1 h2 s X i 1 Under the assumptions (6.2) we obtain bi s X q0 Ld (t0 , q0 , t1 , q1 ) pT0 h (bj aji bi b aij bj bi )FjT q0 Vi pT0 i 1 s X j 1 bi h e aij RjT q0 Vi . s X j 1 e aij ΨTj K(Tj , Qj , Vj ) q0 Vi which is the discrete analogue of (5.1)-(5.2). Similarly, we get q1 Ld (t0 , q0 , t1 , q1 ) h2 s X s X i 1 j 1 h2 s X i 1 (bj aji bi b aij bbj bi )FjT q1 Vi pT1 s X aij ebj )RjT q1 Vi . bi (e j 1 Under the assumptions (6.2) we obtain q1 Ld (t0 , q0 , t1 , q1 ) pT1 h s X i 1 bi h s X j 1 (ebj e aij )ΨTj K(Tj , Qj , Vj ) q1 Vi which is the discrete analogue of (5.3)-(5.4). Under the additional symmetry conditions (6.3) the SPARK method (6.1) is symmetric, and the internal values of the adjoint method satisfy T i Ts 1 i , Qi Qs 1 i , V i Vs 1 i , Ψi Ψs 1 i , see

14 L. O. Jay [16, 17]. We have Ld (t1 , q1 , t0 , q0 ) h h h h s X i 1 s X i 1 s X i 1 s X i 1 bi L(T i , Qi , V i ) bi L(Ts 1 i , Qs 1 i , Vs 1 i ) bs 1 i L(Ti , Qi , Vi ) bi L(Ti , Qi , Vi ) Ld (t0 , q0 , t1 , q1 ). We have fd (t1 , q1 , t0 , q0 )T ( h) s X i 1 h s X i 1 h s X i 1 h s X i 1 bi ( h) bi h s X j 1 bi h s X j 1

of coordinates. For unconstrained Hamiltonian and Lagrangian systems important classes of geometric integrators are symplectic/Poisson integrators [13, 14, 21, 31] and variational integrators which are based on a discrete version of Hamilton's prin-ciple [19, 24]. For unconstrained Lagrangian systems with forcing an important class

Related Documents:

Contents at a Glance Preface xi Introduction 1 I: Spark Foundations 1 Introducing Big Data, Hadoop, and Spark 5 2 Deploying Spark 27 3 Understanding the Spark Cluster Architecture 45 4 Learning Spark Programming Basics 59 II: Beyond the Basics 5 Advanced Programming Using the Spark Core API 111 6 SQL and NoSQL Programming with Spark 161 7 Stream Processing and Messaging Using Spark 209

The overview of Spark and how it is better Hadoop, deploying Spark without Hadoop, Spark history server and Cloudera distribution Spark Basics Spark installation guide, Spark configuration, memory management, executor memory vs. driver memory Working with Spark Shell, the concept of resilient distributed datasets (RDD) Learning to do functional .

Spark vs. MapReduce (2/2) Amir H. Payberah (SICS) Spark and Spark SQL June 29, 2016 23 / 71. Spark vs. MapReduce (2/2) Amir H. Payberah (SICS) Spark and Spark SQL June 29, 2016 23 / 71. Challenge How to design a distributed memory abstraction that is bothfault tolerantande cient? Solution

2. Spark English-Teachers Manual Book II 10 3. Spark English-Teachers Manual Book III 19 4. Spark English-Teachers Manual Book IV 31 5. Spark English-Teachers Manual Book V 45 6. Spark English-Teachers Manual Book VI 59 7. Spark English-Teachers Manual Book VII 73 8. Spark English-Teachers Manual Book VIII 87 Revised Edition, 2017

34 Craig Fraser F Figure 1.(Based on Figure 9, Trait6 (1743)). component v of u parallel to the wall.The velocities u, v and w of the particle correspond to the motions a,; and a of A in the statement of d' Alembert's principle. In Problem IX of Chapter Three of Part Two d'Alembert applies his

(d) Reinstall the spark plug. 3. REMOVE SPARK PLUGS 4. VISUALLY INSPECT SPARK PLUGS Check the spark plug for thread damage and insulator dam-age. If abnormal, replace the spark plug. Recommended spark plug: ND PK20R11 NGK BKR6EP11 5. INSPECT ELECTRODE GAP Maximum electrode gap for used spark plug

Here is a tip sheet guide to get you started using Adobe Spark. Adobe Spark Video is a free App that can be accessed either through your Web Browser at . https://spark.adobe.com, or, you can download the Spark Video app for any mobile device. To download Adobe Spark to your mobile device, visit your App store and search for “Adobe Spark video”

B. Anatomi dan Fisiologi 1. Anatomi Tulang adalah jaringan yang kuat dan tangguh yang memberi bentuk pada tubuh. Skelet atau kerangka adalah rangkaian tulang yang mendukung dan melindungi organ lunak, terutama dalam tengkorak dan panggul. Tulang membentuk rangka penunjang dan pelindung bagi tubuh dan tempat untuk melekatnya otot-otot yang menggerakan kerangka tubuh. Tulang juga merupakan .