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

Implicit Nonholonomic Mechanics with Collisions

Álvaro Rodríguez Abella    Leonardo Colombo Department of Mathematics and Computer Science, Saint Louis University (Madrid Campus), Avenida del Valle, 34, Madrid, 28003, Madrid, Spain. email: alvrod06@ucm.es Centre for Automation and Robotics (CSIC-UPM), Ctra. M300 Campo Real, Km 0,200, Arganda del Rey - 28500 Madrid, Spain. email: leonardo.colombo@csic.es
Abstract

In this paper, variational techniques are used to analyze the dynamics of nonholonomic mechanical systems with impacts. Implicit nonholonomic smooth Lagrangian and Hamiltonian systems are extended to a nonsmooth context appropriate for collisions. In particular, we provide a variational formulation for implicit nonholonomic mechanical systems with collisions, for those collisions that preserve energy and momentum at the impact. Lastly, the theoretical results are illustrated by examining the example of a rolling disk hitting a wall.

keywords:
Implicit Lagrangian systems, Implicit Hamiltonian systems, Nonholonomic systems, Lagrange–d’Alembert–Pontryagin principle, Collision dynamics.
thanks: ARA has been partially supported by Ministerio de Ciencia e Innovación (Spain) under grant PID2021-126124NB-I00. LC acknowledges financial support from Grant PID2022-137909NB-C21 funded by MCIN/AEI/ 10.13039/501100011033.

1 Introduction

A nonholonomic system is a mechanical system subject to constraint functions which are, roughly speaking, functions on the velocities that are not derivable from position constraints. They arise, for instance, in mechanical systems that have rolling or certain types of sliding contact. There are multiple applications in the context of wheeled motion, mobile robotics and robotic manipulation.

A geometrical formulation for mechanical systems with one-sided constraints was developed by Lacomba and Tulczyjew (see Lacomba and Tulczyjew (1990)). Ibort et al. studied the geometric aspects of Lagrangian systems subject to impulsive and one-sided constraints (Ibort et al. (1998)). This was extended to the Hamiltonian formalism by Cortés and Vinogradov (2006).

Mechanical systems subject to collisions are confined within a region of space with a boundary. Collision with the boundary for elastic impacts activates a constraint on the momentum and on the energy after and before the collision. The problem of collisions has been extensively treated in the literature since the early days of mechanics (see Brogliato (1999) for a comprenshive review and references therein). More recently, much work has been done on the rigorous mathematical foundation of impact problems (see Haddad et al. (2006) and Westervelt et al. (2018)). Nonholonomic systems subject to impacts or impulse effects have been previously studied in Clark and Bloch (2019) and Colombo et al. (2022).

In mechanics, implicit Lagrangian and Hamiltonian systems appear in controlled mechanical systems. An important class of implicit mechanical systems studied in Yoshimura and Marsden (2006b) are those with nonholonomic constraints. The aim of this paper is to take one step further and consider implicit mechanical systems subject to nonholonomic constraints and elastic collisions, which occurs when the nonholonomic system impacts the boundary of the configuration space under some suitable conditions. The goal of this paper is to provide a variational formulation for nonholonomic implicit mechanical systems with collisions. In particular, those collisions that preserve energy and momentum at the impact.

The remainder of the paper is structured as follows. Section 2 introduces nonholonomic systems from an explicit point of view via the Lagrange–d’Alembert principle and from an implicit point of view via the Lagrange–d’Alembert–Pontryagin principle. In Section 3 we define the configuration space and the phase space with the objective of introducing the action functional in Section 4, where we derive variationally, via the Hamilton–-d’Alembert–-Pontryagin principle, the equations of motion for implicit nonholonomic Lagrangian systems subject to elastic collisions. We extend the framework to the Hamiltonian side in Section 5, where we derive nonholonomic implicit Hamiltonian systems subject to collisions from a variational perspective. Finally, we study the vertical rolling disk hitting a wall in Section 6.

2 Nonholonomic systems

Let QQ be a differentiable manifold with dim(Q)=n\hbox{dim}(Q)=n. Throughout the text, qiq^{i} will denote a particular choice of local coordinates on this manifold and TQTQ denotes its tangent bundle, with TqQT_{q}Q denoting the tangent space at a specific point qQq\in Q. Usually vqv_{q} denotes a vector at TqQT_{q}Q and, in addition, the coordinate chart qiq^{i} induces a natural coordinate chart on TQTQ denoted by (qi,q˙i)(q^{i},\dot{q}^{i}) with dim(TQ)=2n\hbox{dim}(TQ)=2n. Let TQT^{*}Q be its cotangent bundle, locally described by the positions and the momentum for the system, i.e., (q,p)TQ(q,p)\in T^{*}Q with dim(TQ)=2n\hbox{dim}(T^{*}Q)=2n. The cotangent bundle at a point qQq\in Q is denoted as TqQT_{q}^{*}Q.

A kk-dimensional distribution ΔQ\Delta_{Q} on a manifold Q,Q, is a kk-dimensional subspace ΔQ(q)\Delta_{Q}(q) of TqQT_{q}Q for each qQq\in Q. ΔQ\Delta_{Q} is smooth if for each qQq\in Q there exist a neighborhood UU of qq and kk CC^{\infty} vector fields X1,,XkX_{1},\ldots,X_{k} on UU that span ΔQ\Delta_{Q} at each point of UU, that is, ΔQ(q)=span{X1(q),,Xk(q)}\Delta_{Q}(q)=\hbox{span}\{X_{1}(q),\ldots,X_{k}(q)\}. The rank of ΔQ\Delta_{Q} at qQq\in Q is the dimension of the subspace ΔQ(q)\Delta_{Q}(q), i.e. ϱ:Q,\varrho:Q\to\mathbb{R}, ϱ(q)=dimΔQ(q).\varrho(q)=\dim\Delta_{Q}(q). For any q0Qq_{0}\in Q it is clear that ϱ(q)ϱ(q0)\varrho(q)\geq\varrho(q_{0}) in a neighborhood of q0.q_{0}. If ϱ\varrho is a constant function, then ΔQ\Delta_{Q} is called a regular distribution. In an analogous fashion as for distributions, it is possible to define codistributions.

A smooth regular codistribution ΔQ~\widetilde{\Delta_{Q}} on QQ is a subbundle of TQT^{*}Q with kk-dimensional fiber. Given the concept of codistribution, it is possible to define the annihilator of a distribution. Let ΔQTQ\Delta_{Q}\subset TQ be a distribution, the annihilator of ΔQ\Delta_{Q} is the codistribution defined as

ΔQ(q)={αTqQ|α(v)=α,v=0,vΔQ(q)}\Delta_{Q}^{\circ}(q)=\{\alpha\in T_{q}^{*}Q\,\,|\,\,\alpha(v)=\langle\alpha,v\rangle=0,\,\,\forall v\in\Delta_{Q}(q)\}

for every qQq\in Q.

Linear constraints on the velocities are locally given by equations of the form

ϕa(qi,q˙i)=μia(q)q˙i=0,1am,\phi^{a}(q^{i},\dot{q}^{i})=\mu^{a}_{i}(q)\dot{q}^{i}=0,\quad 1\leq a\leq m,

depending, in general, on their configuration coordinates and their velocities. From an intrinsic point of view, the linear constraints are defined by a regular distribution ΔQ\Delta_{Q} on QQ of constant rank nmn-m such that the annihilator of ΔQ\Delta_{Q} is locally given at each point of QQ by

ΔQ(q)=span{μa(q)=μiadqi;1am},\Delta_{Q}^{\circ}(q)=\operatorname{span}\left\{\mu^{a}(q)=\mu_{i}^{a}dq^{i}\;;1\leq a\leq m\right\},

where the 11-forms μa\mu^{a} are independent at each point of QQ.

We will restrict ourselves to the case of linear nonholonomic constraints. In this case, the constraints are given by a nonintegrable distribution ΔQ\Delta_{Q}. In addition to these constraints, we need to specify the dynamical evolution of the system, usually by fixing a Lagrangian function L:TQL\colon TQ\to\mathbb{R}. The central concepts permitting the extension of mechanics from the Newtonian point of view to the Lagrangian one are the notions of virtual displacements and virtual work. These concepts were formulated in the developments of mechanics and in their application to statics. In nonholonomic dynamics, the procedure is given by the Lagrange–d’Alembert principle. This principle allows us to determine the set of possible values of the constraint forces from the set ΔQ\Delta_{Q} of admissible kinematic states alone. The resulting equations of motion are

[ddt(Lq˙i)Lqi]δqi=0,\left[\frac{d}{dt}\left(\frac{\partial L}{\partial\dot{q}^{i}}\right)-\frac{\partial L}{\partial q^{i}}\right]\delta q^{i}=0,

where δqi\delta q^{i} denotes the virtual displacements verifying

μiaδqi=0.\mu^{a}_{i}\delta q^{i}=0.

By using Lagrange multipliers, we obtain

ddt(Lq˙i)Lqi=λaμia.\frac{d}{dt}\left(\frac{\partial L}{\partial\dot{q}^{i}}\right)-\frac{\partial L}{\partial q^{i}}={\lambda}_{a}\mu^{a}_{i}. (1)

The term on the right-hand side represents the constraint force or reaction force induced by the constraints and the functions λa\lambda_{a} are the Lagrange multipliers which, after being computed using the constraint equations, allow us to obtain a set of second-order differential equations.

Alternatively to the use of Lagrange multipliers, the phase space may be enlarged to the Pontryagin bundle TQTQTQ\oplus T^{*}Q and the Lagrange–d’Alembert–Pontryagin principle may be considered. This variational principle is given by

δt0t1(L(q(t),v(t))+p(t),q˙(t)v(t))𝑑t=0,\delta\int_{t_{0}}^{t_{1}}\left(L(q(t),v(t))+\langle p(t),\dot{q}(t)-v(t)\rangle\right)dt=0,

where v(t)ΔQ(q(t))v(t)\in\Delta_{Q}(q(t)) and the variations (δq(t),δv(t),δp(t))(\delta q(t),\delta v(t),\delta p(t)) are such that δq(t)ΔQ(q(t))\delta q(t)\in\Delta_{Q}(q(t)) and vanishes at the endpoints. Then stationary condition for a curve (q(t),v(t),p(t))(q(t),v(t),p(t)) yields the implicit Lagrange–d’Alembert equations on TQTQTQ\oplus T^{*}Q (see Yoshimura and Marsden (2006b)):

p=Lv,q˙=vΔQ(q),p˙LqΔQ(q).p=\frac{\partial L}{\partial v},\quad\dot{q}=v\in\Delta_{Q}(q),\quad\dot{p}-\frac{\partial L}{\partial q}\in\Delta_{Q}^{\circ}(q).

3 Configuration space and phase space

Let QQ be a smooth manifold with boundary, denoted by Q\partial Q, L:TQL:TQ\to\mathbb{R} be a (possibly degenerate) Lagrangian, and ΔQTQ\Delta_{Q}\subset TQ be a (possibly nonholonomic) constraint distribution. According to Section 2, the annihilator of ΔQ\Delta_{Q} is denoted by ΔQTQ\Delta_{Q}^{\circ}\subset T^{*}Q.

Given [τ0,τ1][\tau_{0},\tau_{1}]\subset\mathbb{R} and τ~[τ0,τ1]\tilde{\tau}\in[\tau_{0},\tau_{1}], the path space with a unique collision (at τ=τ~\tau=\tilde{\tau}) is defined as Ω(Q,τ~)=𝒯×𝒬(τ~)\Omega(Q,\tilde{\tau})=\mathcal{T}\times\mathcal{Q}(\tilde{\tau}), where

𝒯={αTC([τ0,τ1])αT(τ)>0,τ[τ0,τ1]}\mathcal{T}=\left\{\alpha_{T}\in C^{\infty}([\tau_{0},\tau_{1}])\mid\alpha_{T}^{\prime}(\tau)>0,~{}\tau\in[\tau_{0},\tau_{1}]\right\}

and

𝒬(τ~)={αQC0([τ0,τ1],Q)αQ(τ~)Q,\displaystyle\mathcal{Q}(\tilde{\tau})=\big{\{}\alpha_{Q}\in C^{0}([\tau_{0},\tau_{1}],Q)\mid\alpha_{Q}(\tilde{\tau})\in\partial Q, (2)
αQ is piecewise C2 and has only one singularity at τ~}.\displaystyle\alpha_{Q}\text{ is piecewise }C^{2}\text{ and has only one singularity at }\tilde{\tau}\big{\}}.

We only consider one singularity at τ=τ~\tau=\tilde{\tau} for brevity, but similar results hold for a finite amount of singularities, {τ~i1iN}[τ0,τ1]\{\tilde{\tau}_{i}\mid 1\leq i\leq N\}\subset[\tau_{0},\tau_{1}].

Remark 1

Systems with collisions are a particular instance of hybrid systems. For systems with elastic impacts, the guard is given by S={vqTqQqQ,g(vq,nq)>0}S=\{v_{q}\in T_{q}Q\mid q\in\partial Q,~{}g(v_{q},n_{q})>0\}, where gg is a Riemannian metric on QQ and nn is the outward-pointing, unit, normal vector field on the boundary. Similarly, the reset map is given by R(vq)=vqvqR(v_{q})=v_{q}^{{}_{\parallel}}-v_{q}^{\perp}, where vq=g(vq,nq)nqv_{q}^{\perp}=g(v_{q},n_{q})\,n_{q} and vq=vqvqTqQv_{q}^{{}_{\parallel}}=v_{q}-v_{q}^{\perp}\in T_{q}\partial Q. Recall that hybrid systems may experience Zeno behaviour if a trajectory undergoes infinitely many impacts in finite time. In order to avoid this situation, we ask the system to satisfy two conditions (cf. (Goodman and Colombo, 2020, Remark 2.1)):

(i)(i) SR¯(S)=S\cap\overline{R}(S)=\emptyset, where R¯(S)\overline{R}(S) is the closure of R(S)TQR(S)\subset TQ. This condition is clearly satisfied in our case. Indeed, for each vqSv_{q}\in S we have vq0v_{q}^{\perp}\neq 0 and, thus, R(vq)vqg=2vqg>0\parallel R(v_{q})-v_{q}\parallel_{g}=2\parallel v_{q}^{\perp}\parallel_{g}>0, being g\parallel\cdot\parallel_{g} the norm induced by the metric gg.

(ii)(ii) The set of collision times is closed and discrete. This condition, which depends on the topology of the configuration manifold, prevents the existence of an accumulation point and will be assumed in the following.

Under these assumptions, our development is valid in a neighborhood of each collision.

Lemma 1

(Fetecau et al., 2003, Corollary 2.3) Ω(Q,τ~)=𝒯×𝒬(τ~)\Omega(Q,\tilde{\tau})=\mathcal{T}\times\mathcal{Q}(\tilde{\tau}) is a smooth manifold.

Remark 2

Given αT𝒯\alpha_{T}\in\mathcal{T}, we denote [t0,t1]=αT([τ0,τ1])[t_{0},t_{1}]=\alpha_{T}([\tau_{0},\tau_{1}]) and, in order to distinguish between τ\tau-derivatives and tt-derivatives, we use different symbols; namely, αT=dαT/dτ\alpha_{T}^{\prime}=d\alpha_{T}/d\tau and α˙T1=dαT1/dt\dot{\alpha}_{T}^{-1}=d\alpha_{T}^{-1}/dt, where αT1:[t0,t1][τ0,τ1]\alpha_{T}^{-1}:[t_{0},t_{1}]\to[\tau_{0},\tau_{1}] is the inverse of αT\alpha_{T}. Analogously, we denote t~=αT(τ~)\tilde{t}=\alpha_{T}(\tilde{\tau}).

The tangent space of 𝒬(τ~)\mathcal{Q}(\tilde{\tau}) at αQ𝒬(τ~)\alpha_{Q}\in\mathcal{Q}(\tilde{\tau}) is given by

TαQ𝒬(τ~)={ναQC0([τ0,τ1],TQ)\displaystyle T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau})=\Big{\{}\nu_{\alpha_{Q}}\in C^{0}([\tau_{0},\tau_{1}],TQ)\mid (3)
αQ=πTQναQ,ναQ(τ~)TαQ(τ~)Q,\displaystyle\alpha_{Q}=\pi_{TQ}\circ\nu_{\alpha_{Q}},~{}\nu_{\alpha_{Q}}(\tilde{\tau})\in T_{\alpha_{Q}(\tilde{\tau})}\partial Q,
ναQ is piecewise C2 and has only one singularity at τ~},\displaystyle\nu_{\alpha_{Q}}\text{ is piecewise }C^{2}\text{ and has only one singularity at $\tilde{\tau}$}\Big{\}},

where πTQ:TQQ\pi_{TQ}:TQ\to Q is the natural projection. In order to incorporate the constraint distribution, we define the following subspace at each αQ𝒬(τ~)\alpha_{Q}\in\mathcal{Q}(\tilde{\tau}),

Δ𝒬(τ~)(αQ)={ναQTαQ𝒬(τ~)ναQ:[τ0,τ1]ΔQ}.\Delta_{\mathcal{Q}(\tilde{\tau})}(\alpha_{Q})=\left\{\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau})\mid\nu_{\alpha_{Q}}:[\tau_{0},\tau_{1}]\to\Delta_{Q}\right\}.

As usual, we denote T𝒬(τ~)=αQ𝒬(τ~)TαQ𝒬(τ~)T\mathcal{Q}(\tilde{\tau})=\bigsqcup_{\alpha_{Q}\in\mathcal{Q}(\tilde{\tau})}T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}) and Δ𝒬(τ~)=αQ𝒬(τ~)Δ𝒬(τ~)(αQ)\Delta_{\mathcal{Q}(\tilde{\tau})}=\bigsqcup_{\alpha_{Q}\in\mathcal{Q}(\tilde{\tau})}\Delta_{\mathcal{Q}(\tilde{\tau})}(\alpha_{Q}).

Let TαQ𝒬(τ~)={ϕαQ:TαQ𝒬(τ~)ϕαQ is linear and continuous}T_{\alpha_{Q}}^{\prime}\mathcal{Q}(\tilde{\tau})=\{\phi_{\alpha_{Q}}:T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau})\to\mathbb{R}\mid\phi_{\alpha_{Q}}\text{ is linear}\newline \text{ and continuous}\} be the topological dual of TαQ𝒬(τ~)T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}). Since 𝒬(τ~)\mathcal{Q}(\tilde{\tau}) is an infinite dimensional manifold, its topological cotangent bundle is too large to formulate mechanics. For that reason, we will restrict ourselves to the vector subbundle where the Legendre transform of the Lagrangian lie, i.e., we consider a vector subbundle T𝒬(τ~)T𝒬(τ~)T^{\star}\mathcal{Q}(\tilde{\tau})\subset T^{\prime}\mathcal{Q}(\tilde{\tau}) such that 𝔽LναQT𝒬(τ~))\mathbb{F}L\circ\nu_{\alpha_{Q}}\in T^{\star}\mathcal{Q}(\tilde{\tau})) for each ναQTαQ𝒬(τ~)\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}), where 𝔽L:TQTQ\mathbb{F}L:TQ\to T^{*}Q is the Legendre transform of LL.

Lemma 2

For each αQ𝒬(τ~)\alpha_{Q}\in\mathcal{Q}(\tilde{\tau}), the vector space

TαQ𝒬(τ~)={παQC0([τ0,τ1],TQ)\displaystyle T_{\alpha_{Q}}^{\star}\mathcal{Q}(\tilde{\tau})=\big{\{}\pi_{\alpha_{Q}}\in C^{0}([\tau_{0},\tau_{1}],T^{*}Q)\mid (4)
αQ=πTQπαQ,παQ(τ~)TαQ(τ)Q,\displaystyle\alpha_{Q}=\pi_{T^{*}Q}\circ\pi_{\alpha_{Q}},~{}\pi_{\alpha_{Q}}(\tilde{\tau})\in T_{\alpha_{Q}(\tau)}^{*}\partial Q,
παQ is piecewise C2 and has only one singularity at τ~},\displaystyle\pi_{\alpha_{Q}}\text{ is piecewise }C^{2}\text{ and has only one singularity at $\tilde{\tau}$}\big{\}},

where πTQ:TQQ\pi_{T^{*}Q}:T^{*}Q\to Q is the natural projection, is a vector subspace of the topological dual of TαQ𝒬(τ~)T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}) by means of the following L2L^{2}-dual pairing:

παQ,ναQ=τ0τ1παQ(τ)ναQ(τ)𝑑τ,\langle\pi_{\alpha_{Q}},\nu_{\alpha_{Q}}\rangle=\int_{\tau_{0}}^{\tau_{1}}\pi_{\alpha_{Q}}(\tau)\cdot\nu_{\alpha_{Q}}(\tau)\,d\tau,

where \cdot represents the pairing between TQT^{*}Q and TQTQ. Furthermore, this pairing is nondegenerate.

Observe that, in general,

{𝔽LναQTαQ𝒬(τ~)ναQTαQ𝒬(τ~)}TαQ𝒬(τ~),\left\{\mathbb{F}L\circ\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}^{\star}\mathcal{Q}(\tilde{\tau})\mid\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau})\right\}\subsetneq T_{\alpha_{Q}}^{\star}\mathcal{Q}(\tilde{\tau}),

as the Lagrangian is possibly degenerate. As a straightforward consequence of the previous lemma, the vector bundle

T𝒬(τ~)=αQ𝒬(τ~)TαQ𝒬(τ~)𝒬(τ~),παQαQ,T^{\star}\mathcal{Q}(\tilde{\tau})=\bigsqcup_{\alpha_{Q}\in\mathcal{Q}(\tilde{\tau})}T_{\alpha_{Q}}^{\star}\mathcal{Q}(\tilde{\tau})\to\mathcal{Q}(\tilde{\tau}),\quad\pi_{\alpha_{Q}}\mapsto\alpha_{Q},

is a vector subbundle of the topological cotangent bundle of 𝒬(τ~)\mathcal{Q}(\tilde{\tau}).

In the same vein, for each ναQTαQ𝒬(τ~)\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}) and παQTαQ𝒬(τ~)\pi_{\alpha_{Q}}\in T_{\alpha_{Q}}^{\star}\mathcal{Q}(\tilde{\tau}), the iterated bundles are given by

TναQ(T𝒬(τ~))={δναQC0([τ0,τ1],T(TQ))\displaystyle T_{\nu_{\alpha_{Q}}}(T\mathcal{Q}(\tilde{\tau}))=\Big{\{}\delta\nu_{\alpha_{Q}}\in C^{0}([\tau_{0},\tau_{1}],T(TQ))\mid
ναQ=πT(TQ)δναQ,δναQ(τ~)TναQ(τ~)(TQ),\displaystyle\nu_{\alpha_{Q}}=\pi_{T(TQ)}\circ\delta\nu_{\alpha_{Q}},~{}\delta\nu_{\alpha_{Q}}(\tilde{\tau})\in T_{\nu_{\alpha_{Q}}(\tilde{\tau})}(T\partial Q),
δναQ is piecewise C2 and has only one singularity at τ~},\displaystyle\delta\nu_{\alpha_{Q}}\text{ is piecewise $C^{2}$ and has only one singularity at }\tilde{\tau}\Big{\}},

where πT(TQ):T(TQ)TQ\pi_{T(TQ)}:T(TQ)\to TQ is the natural projection, and

TπαQ(T𝒬(τ~))={δπαQC0([τ0,τ1],T(TQ))\displaystyle T_{\pi_{\alpha_{Q}}}(T^{\star}\mathcal{Q}(\tilde{\tau}))=\Big{\{}\delta\pi_{\alpha_{Q}}\in C^{0}([\tau_{0},\tau_{1}],T(T^{*}Q))\mid
παQ=πT(TQ)δπαQ,δπαQ(τ~)TπαQ(τ~)(TQ),\displaystyle\pi_{\alpha_{Q}}=\pi_{T(T^{*}Q)}\circ\delta\pi_{\alpha_{Q}},~{}\delta\pi_{\alpha_{Q}}(\tilde{\tau})\in T_{\pi_{\alpha_{Q}}(\tilde{\tau})}(T^{*}\partial Q),
δπαQ is piecewise C2 and has only one singularity at τ~},\displaystyle\delta\pi_{\alpha_{Q}}\text{ is piecewise $C^{2}$ and has only one singularity at }\tilde{\tau}\Big{\}},

where πT(TQ):T(TQ)TQ\pi_{T(T^{*}Q)}:T(T^{*}Q)\to T^{*}Q is the natural projection. In particular, we consider the constrained iterated bundle,

ΔT𝒬(τ~)(ναQ)=\displaystyle\Delta_{T\mathcal{Q}(\tilde{\tau})}(\nu_{\alpha_{Q}})= {δναQTναQ(T𝒬(τ~))\displaystyle\left\{\delta\nu_{\alpha_{Q}}\in T_{\nu_{\alpha_{Q}}}(T\mathcal{Q}(\tilde{\tau}))\mid\right. (5)
dπTQδναQC0([τ0,τ1],ΔQ)}.\displaystyle\left.d\pi_{TQ}\circ\delta\nu_{\alpha_{Q}}\in C^{0}([\tau_{0},\tau_{1}],\Delta_{Q})\right\}.

4 Nonholonomic implicit Lagrangian mechanics with collisions

Given a path α=(αT,αQ)Ω(Q,τ~)\alpha=(\alpha_{T},\alpha_{Q})\in\Omega(Q,\tilde{\tau}), the associated curve is defined as

qα:[t0,t1]Q,tqα(t)=(αQαT1)(t).q_{\alpha}:[t_{0},t_{1}]\to Q,\quad t\mapsto q_{\alpha}(t)=\left(\alpha_{Q}\circ\alpha_{T}^{-1}\right)(t). (6)

Similarly, given ναQTαQ𝒬(τ~)\nu_{\alpha_{Q}}\in T_{\alpha_{Q}}\mathcal{Q}(\tilde{\tau}) and παQTαQ𝒬(τ~)\pi_{\alpha_{Q}}\in T_{\alpha_{Q}}^{\prime}\mathcal{Q}(\tilde{\tau}), we set

vα:[t0,t1]TQ,tvα(t)=(ναQαT1)(t),pα:[t0,t1]TQ,tpα(t)=(παQαT1)(t).\begin{array}[]{ll}v_{\alpha}:[t_{0},t_{1}]\to TQ,&t\mapsto v_{\alpha}(t)=\left(\nu_{\alpha_{Q}}\circ\alpha_{T}^{-1}\right)(t),\\ p_{\alpha}:[t_{0},t_{1}]\to T^{*}Q,&t\mapsto p_{\alpha}(t)=\left(\pi_{\alpha_{Q}}\circ\alpha_{T}^{-1}\right)(t).\end{array}

It is clear that πTQvα=πTQpα=qα\pi_{TQ}\circ v_{\alpha}=\pi_{T^{*}Q}\circ p_{\alpha}=q_{\alpha}.

By regarding Ω(Q,τ~)\Omega(Q,\tilde{\tau}) as a trivial vector bundle over 𝒬(τ~)\mathcal{Q}(\tilde{\tau}) with the projection onto the second factor, the Lagrange–d’Alembert–Pontryagin action functional,

𝕊:Ω(Q,τ~)×𝒬(τ~)(T𝒬(τ~)T𝒬(τ~)),\mathbb{S}:\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}\left(T\mathcal{Q}(\tilde{\tau})\oplus T^{\star}\mathcal{Q}(\tilde{\tau})\right)\to\mathbb{R},

where ×𝒬(τ~)\times_{\mathcal{Q}(\tilde{\tau})} denotes the fibered product over 𝒬(τ~)\mathcal{Q}(\tilde{\tau}), is defined as

𝕊(α,ναQ,παQ)=t0t1(L(vα(t))+pα(t)(q˙α(t)vα(t)))𝑑t\displaystyle\mathbb{S}\left(\alpha,\nu_{\alpha_{Q}},\pi_{\alpha_{Q}}\right)=\int_{t_{0}}^{t_{1}}\left(L(v_{\alpha}(t))+p_{\alpha}(t)\cdot\left(\dot{q}_{\alpha}(t)-v_{\alpha}(t)\right)\right)dt
=τ0τ1(L(ναQ(τ))+παQ(τ)(αQ(τ)αT(τ)ναQ(τ)))αT(τ)𝑑τ.\displaystyle=\int_{\tau_{0}}^{\tau_{1}}\left(L\left(\nu_{\alpha_{Q}}(\tau)\right)+\pi_{\alpha_{Q}}(\tau)\cdot\left(\frac{\alpha_{Q}^{\prime}(\tau)}{\alpha_{T}^{\prime}(\tau)}-\nu_{\alpha_{Q}}(\tau)\right)\right)\alpha_{T}^{\prime}(\tau)\,d\tau.

The equality between the first and the second expressions can be easily checked by considering the change of variable t=αT(τ)t=\alpha_{T}(\tau). By recalling that the energy of the system, E:TQTQE:TQ\oplus T^{*}Q\to\mathbb{R}, is given by

E(vq,pq)=pqvqL(vq),(vq,pq)TQTQ,E(v_{q},p_{q})=p_{q}\cdot v_{q}-L(v_{q}),\quad(v_{q},p_{q})\in TQ\oplus T^{*}Q, (7)

the action functional may be rewritten as

𝕊(α,ναQ,παQ)=t0t1(pα(t)q˙α(t)E(vα(t),pα(t))dt\displaystyle\mathbb{S}\left(\alpha,\nu_{\alpha_{Q}},\pi_{\alpha_{Q}}\right)=\int_{t_{0}}^{t_{1}}\left(p_{\alpha}(t)\cdot\dot{q}_{\alpha}(t)-E(v_{\alpha}(t),p_{\alpha}(t)\right)dt
=τ0τ1(παQ(τ)αQ(τ)αT(τ)E(ναQ(τ),παQ(τ)))αT(τ)𝑑τ.\displaystyle=\int_{\tau_{0}}^{\tau_{1}}\left(\pi_{\alpha_{Q}}(\tau)\cdot\frac{\alpha_{Q}^{\prime}(\tau)}{\alpha_{T}^{\prime}(\tau)}-E\left(\nu_{\alpha_{Q}}(\tau),\pi_{\alpha_{Q}}(\tau)\right)\right)\alpha_{T}^{\prime}(\tau)\,d\tau.
Definition 1 (Hamilton–d’Alembert–Pontryagin principle)

A path

c=((αT,αQ),ναQ,παQ)Ω(Q,τ~)×𝒬(τ~)(Δ𝒬(τ~)T𝒬(τ~))\texttt{c}=((\alpha_{T},\alpha_{Q}),\nu_{\alpha_{Q}},\pi_{\alpha_{Q}})\in\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}\left(\Delta_{\mathcal{Q}(\tilde{\tau})}\oplus T^{\star}\mathcal{Q}(\tilde{\tau})\right)

is stationary (or critical) for the action functional 𝕊\mathbb{S} if it satisfies

d𝕊(c)(δc)=0,d\mathbb{S}(\texttt{c})(\delta\texttt{c})=0,

for every variation δc=((δαT,δαQ),δναQ,δπαQ)TαΩ(Q,τ~)×ΔT𝒬(τ~)(ναQ)×TπαQ(T𝒬(τ~))\delta\texttt{c}=\left((\delta\alpha_{T},\delta\alpha_{Q}),\delta\nu_{\alpha_{Q}},\delta\pi_{\alpha_{Q}}\right)\in T_{\alpha}\Omega(Q,\tilde{\tau})\times\Delta_{T\mathcal{Q}(\tilde{\tau})}(\nu_{\alpha_{Q}})\times T_{\pi_{\alpha_{Q}}}(T^{\star}\mathcal{Q}(\tilde{\tau})) such that δαT(τ0)=δαT(τ1)=0\delta\alpha_{T}(\tau_{0})=\delta\alpha_{T}(\tau_{1})=0, δαQ(τ0)=δαQ(τ1)=0\delta\alpha_{Q}(\tau_{0})=\delta\alpha_{Q}(\tau_{1})=0 and

dπTQδναQ=dπTQδπαQ=δαQ.d\pi_{TQ}\circ\delta\nu_{\alpha_{Q}}=d\pi_{T^{*}Q}\circ\delta\pi_{\alpha_{Q}}=\delta\alpha_{Q}. (8)
Theorem 1

A (local) curve

(α,ναQ,παQ)\displaystyle(\alpha,\nu_{\alpha_{Q}},\pi_{\alpha_{Q}})\simeq
(αT,αQ,νQ,πQ)Ω(Q,τ~)×𝒬(τ~)(Δ𝒬(τ~)T𝒬(τ~))\displaystyle(\alpha_{T},\alpha_{Q},\nu_{Q},\pi_{Q})\in\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}\left(\Delta_{\mathcal{Q}(\tilde{\tau})}\oplus T^{*}\mathcal{Q}(\tilde{\tau})\right)

is critical for the action functional 𝕊\mathbb{S} if and only if it satisfies the implicit Euler–Lagrange equations,

{πQαTLq(αQ,νQ)ΔQ(αQ),E(αQ,νQ,πQ)=0,πQ=Lv(αQ,νQ),νQ=αQαTΔQ(αQ),\left\{\begin{array}[]{ll}\displaystyle\pi_{Q}^{\prime}-\alpha_{T}^{\prime}\frac{\partial L}{\partial q}(\alpha_{Q},\nu_{Q})\in\Delta_{Q}^{\circ}(\alpha_{Q}),&\displaystyle E^{\prime}(\alpha_{Q},\nu_{Q},\pi_{Q})=0,\\ \displaystyle\pi_{Q}=\frac{\partial L}{\partial v}(\alpha_{Q},\nu_{Q}),&\displaystyle\nu_{Q}=\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}\in\Delta_{Q}(\alpha_{Q}),\end{array}\right.

on [τ0,τ~)(τ~,τ1][\tau_{0},\tilde{\tau})\cup(\tilde{\tau},\tau_{1}], together with the conditions for the elastic impact,

πQ(τ~+)πQ(τ~)(TQΔQ)=(TQ)+ΔQ,\displaystyle\pi_{Q}(\tilde{\tau}^{+})-\pi_{Q}(\tilde{\tau}^{-})\in(T\partial Q\cap\Delta_{Q})^{\circ}=(T\partial Q)^{\circ}+\Delta_{Q}^{\circ},
E(αQ(τ~),νQ(τ~),πQ(τ~))=E(αQ(τ~+),νQ(τ~+),πQ(τ~+)),\displaystyle E(\alpha_{Q}(\tilde{\tau}^{-}),\nu_{Q}(\tilde{\tau}^{-}),\pi_{Q}(\tilde{\tau}^{-}))=E(\alpha_{Q}(\tilde{\tau}^{+}),\nu_{Q}(\tilde{\tau}^{+}),\pi_{Q}(\tilde{\tau}^{+})),

where the annihilitaros are with respect to TQTQ.

Proof: Let TQQ×VTQ\simeq Q\times V be a trivialization of the tangent bundle of QQ, and consider the induced trivializations of the cotangent bundle of QQ, TQQ×VT^{*}Q\simeq Q\times V^{*}, as well as of the iterated bundles T(TQ)Q×V×V×VT(TQ)\simeq Q\times V\times V\times V and T(TQ)Q×V×V×VT(T^{*}Q)\simeq Q\times V^{*}\times V\times V^{*}. Locally, we may write α(αT,αQ,αT,αQ)\alpha^{\prime}\simeq(\alpha_{T},\alpha_{Q},\alpha_{T}^{\prime},\alpha_{Q}^{\prime}), ναQ(αQ,νQ)\nu_{\alpha_{Q}}\simeq(\alpha_{Q},\nu_{Q}) and παQ(αQ,πQ)\pi_{\alpha_{Q}}\simeq(\alpha_{Q},\pi_{Q}) for some αQ,νQ:[τ0,τ1]V\alpha_{Q}^{\prime},\nu_{Q}:[\tau_{0},\tau_{1}]\to V and πQ:[τ0,τ1]V\pi_{Q}:[\tau_{0},\tau_{1}]\to V^{*}. Moreover, the variations locally read δα(αT,αQ,δαT,δαQ)\delta\alpha\simeq(\alpha_{T},\alpha_{Q},\delta\alpha_{T},\delta\alpha_{Q}), δναQ(αQ,νQ,βQ,δνQ)\delta\nu_{\alpha_{Q}}\simeq(\alpha_{Q},\nu_{Q},\beta_{Q},\delta\nu_{Q}) and δπαQ(αQ,πQ,γQ,δπQ)\delta\pi_{\alpha_{Q}}\simeq(\alpha_{Q},\pi_{Q},\gamma_{Q},\delta\pi_{Q}) for some δαQ,βQ,γQ,δνQ:[τ0,τ1]V\delta\alpha_{Q},\beta_{Q},\gamma_{Q},\delta\nu_{Q}:[\tau_{0},\tau_{1}]\to V, δπQ:[τ0,τ1]V\delta\pi_{Q}:[\tau_{0},\tau_{1}]\to V^{*}. In fact, equation (8) ensures that δαQ=βQ=γQ\delta\alpha_{Q}=\beta_{Q}=\gamma_{Q}. Moreover, by locally regarding ΔQ(q)V\Delta_{Q}(q)\subset V for each qQq\in Q, the conditions ναQΔ𝒬(τ~)\nu_{\alpha_{Q}}\in\Delta_{\mathcal{Q}(\tilde{\tau})} and δναQΔT𝒬(τ~)(ναQ)\delta\nu_{\alpha_{Q}}\in\Delta_{T\mathcal{Q}(\tilde{\tau})}(\nu_{\alpha_{Q}}) read νQ(τ)ΔQ(αQ(τ))\nu_{Q}(\tau)\in\Delta_{Q}(\alpha_{Q}(\tau)) and δαQ(τ)ΔQ(αQ(τ))\delta\alpha_{Q}(\tau)\in\Delta_{Q}(\alpha_{Q}(\tau)) for each τ[τ0,τ1]\tau\in[\tau_{0},\tau_{1}], respectively. At last, the condition δναQ(τ~)TναQ(τ~)(TQ)\delta\nu_{\alpha_{Q}}(\tilde{\tau})\in T_{\nu_{\alpha_{Q}}(\tilde{\tau})}(T\partial Q) yields the local condition δαQ(τ~)W\delta\alpha_{Q}(\tilde{\tau})\in W, where WVW\subset V is a subspace of co-dimension one such that TQQ×WT\partial Q\simeq\partial Q\times W.

As a result, the variation of the action functional reads

d𝕊(α,ναQ,παQ)(δα,δναQ,δπαQ)\displaystyle d\mathbb{S}(\alpha,\nu_{\alpha_{Q}},\pi_{\alpha_{Q}})\left(\delta\alpha,\delta\nu_{\alpha_{Q}},\delta\pi_{\alpha_{Q}}\right)\simeq
d𝕊(αT,αQ,νQ,πQ)(δαT,δαQ,δνQ,δπQ)=\displaystyle d\mathbb{S}(\alpha_{T},\alpha_{Q},\nu_{Q},\pi_{Q})(\delta\alpha_{T},\delta\alpha_{Q},\delta\nu_{Q},\delta\pi_{Q})=
τ0τ1(LqδαQ+LvδνQ+δπQ(αQαTνQ)\displaystyle\int_{\tau_{0}}^{\tau_{1}}\left(\frac{\partial L}{\partial q}\cdot\delta\alpha_{Q}+\frac{\partial L}{\partial v}\cdot\delta\nu_{Q}+\delta\pi_{Q}\cdot\left(\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}-\nu_{Q}\right)\right.
+πQ(δαQαTαQδαT(αT)2δνQ))αTdτ\displaystyle\hskip 22.76219pt\left.+\pi_{Q}\cdot\left(\frac{\delta\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}-\frac{\alpha_{Q}^{\prime}\delta\alpha_{T}^{\prime}}{(\alpha_{T}^{\prime})^{2}}-\delta\nu_{Q}\right)\right)\alpha_{T}^{\prime}\,d\tau
+τ0τ1(L+πQ(αQαTνQ))δαT𝑑τ,\displaystyle\hskip 22.76219pt+\int_{\tau_{0}}^{\tau_{1}}\left(L+\pi_{Q}\cdot\left(\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}-\nu_{Q}\right)\right)\delta\alpha_{T}^{\prime}\,d\tau,

where the Lagrangian, as well as its partial derivatives, are evaluated at (αQ,νQ)\left(\alpha_{Q},\nu_{Q}\right). After splitting the integration domain, [τ0,τ1]{τ~}=[τ0,τ~)(τ~,τ1][\tau_{0},\tau_{1}]-\{\tilde{\tau}\}=[\tau_{0},\tilde{\tau})\cup(\tilde{\tau},\tau_{1}], as well as integrating by parts on each sub-interval, we may rewrite the previous expression as

d𝕊(α,ναQ,παQ)(δα,δναQ,δπαQ)\displaystyle d\mathbb{S}(\alpha,\nu_{\alpha_{Q}},\pi_{\alpha_{Q}})\left(\delta\alpha,\delta\nu_{\alpha_{Q}},\delta\pi_{\alpha_{Q}}\right)\simeq
τ0τ~((αTLqπQ)δαQddτ(LπQνQ)δαT\displaystyle\int_{\tau_{0}}^{\tilde{\tau}}\left(\left(\alpha_{T}^{\prime}\frac{\partial L}{\partial q}-\pi_{Q}^{\prime}\right)\cdot\delta\alpha_{Q}-\frac{d}{d\tau}(L-\pi_{Q}\cdot\nu_{Q})\delta\alpha_{T}\right.
+αT(LvπQ)δνQ+δπQ(αQαTνQ))dτ\displaystyle\hskip 22.76219pt\left.+\alpha_{T}^{\prime}\left(\frac{\partial L}{\partial v}-\pi_{Q}\right)\cdot\delta\nu_{Q}+\delta\pi_{Q}\cdot(\alpha_{Q}^{\prime}-\alpha_{T}^{\prime}\nu_{Q})\right)d\tau
+τ~τ1((αTLqπQ)δαQddτ(LπQνQ)δαT\displaystyle+\int_{\tilde{\tau}}^{\tau_{1}}\left(\left(\alpha_{T}^{\prime}\frac{\partial L}{\partial q}-\pi_{Q}^{\prime}\right)\cdot\delta\alpha_{Q}-\frac{d}{d\tau}(L-\pi_{Q}\cdot\nu_{Q})\delta\alpha_{T}\right.
+αT(LvπQ)δνQ+δπQ(αQαTνQ))dτ\displaystyle\hskip 22.76219pt\left.+\alpha_{T}^{\prime}\left(\frac{\partial L}{\partial v}-\pi_{Q}\right)\cdot\delta\nu_{Q}+\delta\pi_{Q}\cdot(\alpha_{Q}^{\prime}-\alpha_{T}^{\prime}\nu_{Q})\right)d\tau
+[πQδαQ+(LπQνQ)δαT]τ=τ0τ=τ~\displaystyle+\Big{[}\pi_{Q}\cdot\delta\alpha_{Q}+(L-\pi_{Q}\cdot\nu_{Q})\delta\alpha_{T}\Big{]}_{\tau=\tau_{0}}^{\tau=\tilde{\tau}^{-}}
+[πQδαQ+(LπQνQ)δαT]τ=τ~+τ=τ1.\displaystyle+\Big{[}\pi_{Q}\cdot\delta\alpha_{Q}+(L-\pi_{Q}\cdot\nu_{Q})\delta\alpha_{T}\Big{]}_{\tau=\tilde{\tau}^{+}}^{\tau=\tau_{1}}.

Since the previous expression vanishes for free variations (δα,δνQ,δπQ)(\delta\alpha,\delta\nu_{Q},\delta\pi_{Q}) such that δαQΔQ(αQ)\delta\alpha_{Q}\in\Delta_{Q}(\alpha_{Q}), δαT(τ0)=δαT(τ1)=0\delta\alpha_{T}(\tau_{0})=\delta\alpha_{T}(\tau_{1})=0 and δαQ(τ0)=δαQ(τ1)=0\delta\alpha_{Q}(\tau_{0})=\delta\alpha_{Q}(\tau_{1})=0, we obtain the desired equations and impact conditions. \square

By using the change of variable t=αT(τ)t=\alpha_{T}(\tau), we have q˙α=αQ/αT\dot{q}_{\alpha}=\alpha_{Q}^{\prime}/\alpha_{T}^{\prime} and p˙α=πQ/αT\dot{p}_{\alpha}=\pi_{Q}^{\prime}/\alpha_{T}^{\prime}. Then, the implicit Euler–Lagrange equations for a (local) curve

(vα,pα)(qα,v,p):[t0,t1]TQTQ(v_{\alpha},p_{\alpha})\simeq(q_{\alpha},v,p):[t_{0},t_{1}]\to TQ\oplus T^{*}Q

take the form

{p˙Lq(qα,v)ΔQ(qα),E˙(qα,v,p)=0,p=Lv(qα,v),v=q˙αΔQ(qα),\left\{\begin{array}[]{ll}\displaystyle\dot{p}-\frac{\partial L}{\partial q}(q_{\alpha},v)\in\Delta_{Q}^{\circ}(q_{\alpha}),&\displaystyle\dot{E}(q_{\alpha},v,p)=0,\vspace{0.1cm}\\ \displaystyle p=\frac{\partial L}{\partial v}(q_{\alpha},v),&\displaystyle v=\dot{q}_{\alpha}\in\Delta_{Q}(q_{\alpha}),\end{array}\right. (9)

on [t0,t~)(t~,t1]\left[t_{0},\tilde{t}\right)\cup\left(\tilde{t},t_{1}\right]. Similarly, the conditions for the elastic impact read

p(t~+)p(t~)(TQΔQ)=(TQ)+ΔQ,\displaystyle p\left(\tilde{t}^{+}\right)-p\left(\tilde{t}^{-}\right)\in(T\partial Q\cap\Delta_{Q})^{\circ}=(T\partial Q)^{\circ}+\Delta_{Q}^{\circ}, (10)
E(qα(t~),v(t~),p(t~))=E(qα(t~+),v(t~+),p(t~+)),\displaystyle E\left(q_{\alpha}\left(\tilde{t}^{-}\right),v\left(\tilde{t}^{-}\right),p\left(\tilde{t}^{-}\right)\right)=E\left(q_{\alpha}\left(\tilde{t}^{+}\right),v\left(\tilde{t}^{+}\right),p\left(\tilde{t}^{+}\right)\right),
v(t~+)=q˙α(t~+)ΔQ.\displaystyle v\left(\tilde{t}^{+}\right)=\dot{q}_{\alpha}\left(\tilde{t}^{+}\right)\in\Delta_{Q}. (11)

Energy balance: It may be shown that the conservation of the energy along the solutions, E˙(qα,v,p)=0\dot{E}(q_{\alpha},v,p)=0, is redundant, as it may be obtained from the remaining equations.

For unconstrained systems, i.e., ΔQ=TQ\Delta_{Q}=TQ, the Hamilton–d’Alembert–Pontryagin principle reduces to the Hamilton–Pontryagin principle, and the implicit Euler–Lagrange equations of motion read as

p˙=Lq(qα,v),p=Lv(qα,v),v=q˙α.\dot{p}=\frac{\partial L}{\partial q}(q_{\alpha},v),\qquad p=\frac{\partial L}{\partial v}(q_{\alpha},v),\qquad v=\dot{q}_{\alpha}.

5 Nonholonomic implicit Hamiltonian mechanics with collisions

The results in the previous section may be obtained in the Hamiltonian side as well. Namely, given a (possibly degenerate) Hamiltonian, H:TQH:T^{*}Q\to\mathbb{R}, the Hamilton–d’Alembert–Pontryagin action functional,

𝔖:Ω(Q,τ~)×𝒬(τ~)T𝒬(τ~),\mathfrak{S}:\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}T^{\star}\mathcal{Q}(\tilde{\tau})\to\mathbb{R},

is defined as

𝔖(α,παQ)=t0t1(pα(t)q˙α(t)H(pα(t)))𝑑t\displaystyle\mathfrak{S}\left(\alpha,\pi_{\alpha_{Q}}\right)=\int_{t_{0}}^{t_{1}}\big{(}p_{\alpha}(t)\cdot\dot{q}_{\alpha}(t)-H(p_{\alpha}(t))\big{)}\,dt
=τ0τ1(παQ(τ)αQ(τ)αT(τ)H(παQ(τ)))αT(τ)𝑑τ.\displaystyle=\int_{\tau_{0}}^{\tau_{1}}\left(\pi_{\alpha_{Q}}(\tau)\cdot\frac{\alpha_{Q}^{\prime}(\tau)}{\alpha_{T}^{\prime}(\tau)}-H\left(\pi_{\alpha_{Q}}(\tau)\right)\right)\alpha_{T}^{\prime}(\tau)\,d\tau.

Recall that from this point of view the energy of the system is simply given by the Hamiltonian.

Definition 2 (Variational principle in the phase space)

A path

c=((αT,αQ),παQ)Ω(Q,τ~)×𝒬(τ~)T𝒬(τ~)\texttt{c}=((\alpha_{T},\alpha_{Q}),\pi_{\alpha_{Q}})\in\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}T^{\star}\mathcal{Q}(\tilde{\tau})

such that αQΔ𝒬(τ~)(αQ)\alpha_{Q}^{\prime}\in\Delta_{\mathcal{Q}(\tilde{\tau})}(\alpha_{Q}) is stationary (or critical) for the action functional 𝔖\mathfrak{S} if it satisfies d𝔖(c)(δc)=0d\mathfrak{S}(\texttt{c})(\delta\texttt{c})=0 for every variation δc=((δαT,δαQ),δπαQ)TαΩ(Q,τ~)×TπαQ(T𝒬(τ~))\delta\texttt{c}=\left((\delta\alpha_{T},\delta\alpha_{Q}),\delta\pi_{\alpha_{Q}}\right)\in T_{\alpha}\Omega(Q,\tilde{\tau})\times T_{\pi_{\alpha_{Q}}}(T^{\star}\mathcal{Q}(\tilde{\tau})) such that δαT(τ0)=δαT(τ1)=0\delta\alpha_{T}(\tau_{0})=\delta\alpha_{T}(\tau_{1})=0, δαQ(τ0)=δαQ(τ1)=0\delta\alpha_{Q}(\tau_{0})=\delta\alpha_{Q}(\tau_{1})=0 and

dπTQδπαQ=δαQΔ𝒬(τ~)(αQ).d\pi_{T^{*}Q}\circ\delta\pi_{\alpha_{Q}}=\delta\alpha_{Q}\in\Delta_{\mathcal{Q}(\tilde{\tau})}(\alpha_{Q}). (12)
Theorem 2

A (local) curve

(α,παQ)(αT,αQ,πQ)Ω(Q,τ~)×𝒬(τ~)T𝒬(τ~)(\alpha,\pi_{\alpha_{Q}})\simeq(\alpha_{T},\alpha_{Q},\pi_{Q})\in\Omega(Q,\tilde{\tau})\times_{\mathcal{Q}(\tilde{\tau})}T^{*}\mathcal{Q}(\tilde{\tau})

such that αQΔ𝒬(τ~)(αQ)\alpha_{Q}^{\prime}\in\Delta_{\mathcal{Q}(\tilde{\tau})}(\alpha_{Q}) is critical for the action functional 𝔖\mathfrak{S} if and only if it satisfies the implicit Hamilton equations,

πQ+αTHq(αQ,πQ)ΔQ(αQ),H(αQ,πQ)=0,\displaystyle\pi_{Q}^{\prime}+\alpha_{T}^{\prime}\frac{\partial H}{\partial q}(\alpha_{Q},\pi_{Q})\in\Delta_{Q}^{\circ}(\alpha_{Q}),\qquad H^{\prime}(\alpha_{Q},\pi_{Q})=0,
αQαT=Hp(αQ,πQ)ΔQ(αQ),\displaystyle\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}=\frac{\partial H}{\partial p}(\alpha_{Q},\pi_{Q})\in\Delta_{Q}(\alpha_{Q}),

on [τ0,τ~)(τ~,τ1][\tau_{0},\tilde{\tau})\cup(\tilde{\tau},\tau_{1}], together with the conditions for the elastic impact,

{πQ(τ~+)πQ(τ~+)(TQΔQ)=(TQ)+ΔQ,H(αQ(τ~),πQ(τ~))=H(αQ(τ~+),πQ(τ~+)).\left\{\begin{array}[]{l}\displaystyle\pi_{Q}(\tilde{\tau}^{+})-\pi_{Q}(\tilde{\tau}^{+})\in(T\partial Q\cap\Delta_{Q})^{\circ}=(T\partial Q)^{\circ}+\Delta_{Q}^{\circ},\vspace{2mm}\\ \displaystyle H(\alpha_{Q}(\tilde{\tau}^{-}),\pi_{Q}(\tilde{\tau}^{-}))=H(\alpha_{Q}(\tilde{\tau}^{+}),\pi_{Q}(\tilde{\tau}^{+})).\end{array}\right.

Proof: Let TQQ×VTQ\simeq Q\times V be a trivialization of the tangent bundle of QQ, and consider the induced trivializations of the cotangent bundle of QQ, TQQ×VT^{*}Q\simeq Q\times V^{*}, as well as of the iterated bundle T(TQ)Q×V×V×VT(T^{*}Q)\simeq Q\times V^{*}\times V\times V^{*}. Locally, we may write α(αT,αQ,αT,αQ)\alpha^{\prime}\simeq(\alpha_{T},\alpha_{Q},\alpha_{T}^{\prime},\alpha_{Q}^{\prime}) and παQ(αQ,πQ)\pi_{\alpha_{Q}}\simeq(\alpha_{Q},\pi_{Q}) for some αQ:[τ0,τ1]V\alpha_{Q}^{\prime}:[\tau_{0},\tau_{1}]\to V and πQ:[τ0,τ1]V\pi_{Q}:[\tau_{0},\tau_{1}]\to V^{*}. Moreover, the variations locally read δα(αT,αQ,δαT,δαQ)\delta\alpha\simeq(\alpha_{T},\alpha_{Q},\delta\alpha_{T},\delta\alpha_{Q}) and δπαQ(αQ,πQ,γQ,δπQ)\delta\pi_{\alpha_{Q}}\simeq(\alpha_{Q},\pi_{Q},\gamma_{Q},\delta\pi_{Q}) for some δαQ,γQ:[τ0,τ1]V\delta\alpha_{Q},\gamma_{Q}:[\tau_{0},\tau_{1}]\to V and δπQ:[τ0,τ1]V\delta\pi_{Q}:[\tau_{0},\tau_{1}]\to V^{*}. In fact, equation (12) ensures that δαQ=γQΔQ(αQ)\delta\alpha_{Q}=\gamma_{Q}\in\Delta_{Q}(\alpha_{Q}). Moreover, we have δαQ(τ~)W\delta\alpha_{Q}(\tilde{\tau})\in W, where WVW\subset V is a subspace of co-dimension one such that TQQ×WT\partial Q\simeq\partial Q\times W.

As a result, the variation of the action functional reads

d𝔖(α,παQ)(δα,δπαQ)\displaystyle d\mathfrak{S}(\alpha,\pi_{\alpha_{Q}})\left(\delta\alpha,\delta\pi_{\alpha_{Q}}\right)\simeq
d𝔖(αT,αQ,πQ)(δαT,δαQ,δπQ)=\displaystyle d\mathfrak{S}(\alpha_{T},\alpha_{Q},\pi_{Q})(\delta\alpha_{T},\delta\alpha_{Q},\delta\pi_{Q})=
τ0τ1(δπQαQαT+πQ(δαQαTαQδαT(αT)2)\displaystyle\int_{\tau_{0}}^{\tau_{1}}\bigg{(}\delta\pi_{Q}\cdot\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}+\pi_{Q}\cdot\left(\frac{\delta\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}-\frac{\alpha_{Q}^{\prime}\delta\alpha_{T}^{\prime}}{(\alpha_{T}^{\prime})^{2}}\right)
HqδαQHpδπQ)αTdτ\displaystyle\hskip 22.76219pt-\frac{\partial H}{\partial q}\cdot\delta\alpha_{Q}-\frac{\partial H}{\partial p}\cdot\delta\pi_{Q}\bigg{)}\,\alpha_{T}^{\prime}\,d\tau
+τ0τ1(πQαQαTH)δαT𝑑τ,\displaystyle+\int_{\tau_{0}}^{\tau_{1}}\left(\pi_{Q}\cdot\frac{\alpha_{Q}^{\prime}}{\alpha_{T}^{\prime}}-H\right)\delta\alpha_{T}^{\prime}\,d\tau,

where the Hamiltonian, as well as its partial derivatives, are evaluated at (αQ,πQ)\left(\alpha_{Q},\pi_{Q}\right). After splitting the integration domain, [τ0,τ1]{τ~}=[τ0,τ~)(τ~,τ1][\tau_{0},\tau_{1}]-\{\tilde{\tau}\}=[\tau_{0},\tilde{\tau})\cup(\tilde{\tau},\tau_{1}], as well as integrating by parts on each sub-interval, we may rewrite the previous expression as

d𝔖(α,παQ)(δα,δπαQ)\displaystyle d\mathfrak{S}(\alpha,\pi_{\alpha_{Q}})\left(\delta\alpha,\delta\pi_{\alpha_{Q}}\right)\simeq
τ0τ~((πQαTHq)δαQ+HδαT\displaystyle\int_{\tau_{0}}^{\tilde{\tau}}\bigg{(}\left(-\pi_{Q}^{\prime}-\alpha_{T}^{\prime}\frac{\partial H}{\partial q}\right)\cdot\delta\alpha_{Q}+H^{\prime}\delta\alpha_{T}
+δπQ(αQαTHp))dτ\displaystyle\hskip 22.76219pt+\delta\pi_{Q}\cdot\left(\alpha_{Q}^{\prime}-\alpha_{T}^{\prime}\frac{\partial H}{\partial p}\right)\bigg{)}\,d\tau
+τ~τ1((πQαTHq)δαQ+HδαT\displaystyle+\int_{\tilde{\tau}}^{\tau_{1}}\bigg{(}\left(-\pi_{Q}^{\prime}-\alpha_{T}^{\prime}\frac{\partial H}{\partial q}\right)\cdot\delta\alpha_{Q}+H^{\prime}\delta\alpha_{T}
+δπQ(αQαTHp))dτ\displaystyle\hskip 22.76219pt+\delta\pi_{Q}\cdot\left(\alpha_{Q}^{\prime}-\alpha_{T}^{\prime}\frac{\partial H}{\partial p}\right)\bigg{)}\,d\tau
+[πQδαQHδαT]τ=τ0τ=τ~+[πQδαQHδαT]τ=τ~+τ=τ1.\displaystyle+\Big{[}\pi_{Q}\cdot\delta\alpha_{Q}-H\delta\alpha_{T}\Big{]}_{\tau=\tau_{0}}^{\tau=\tilde{\tau}^{-}}+\Big{[}\pi_{Q}\cdot\delta\alpha_{Q}-H\delta\alpha_{T}\Big{]}_{\tau=\tilde{\tau}^{+}}^{\tau=\tau_{1}}.

Since the previous expression vanishes for free variations (δα,δπQ)(\delta\alpha,\delta\pi_{Q}) such that δαQΔQ(αQ)\delta\alpha_{Q}\in\Delta_{Q}(\alpha_{Q}), δαT(τ0)=δαT(τ1)=0\delta\alpha_{T}(\tau_{0})=\delta\alpha_{T}(\tau_{1})=0 and δαQ(τ0)=δαQ(τ1)=0\delta\alpha_{Q}(\tau_{0})=\delta\alpha_{Q}(\tau_{1})=0, we obtain the desired equations and impact conditions. \square

As for the Lagrangian equations, by means of the change of variable t=αT(τ)t=\alpha_{T}(\tau), the implicit Hamilton equations for a (local) curve

pα(qα,p):[t0,t1]TQp_{\alpha}\simeq(q_{\alpha},p):[t_{0},t_{1}]\to T^{*}Q

take the form

p˙+Hq(qα,p)ΔQ(qα),H˙(qα,p)=0,\displaystyle\dot{p}+\frac{\partial H}{\partial q}(q_{\alpha},p)\in\Delta_{Q}^{\circ}(q_{\alpha}),\qquad\dot{H}(q_{\alpha},p)=0,
q˙α=Hp(qα,p)ΔQ(qα),\displaystyle\dot{q}_{\alpha}=\frac{\partial H}{\partial p}(q_{\alpha},p)\in\Delta_{Q}(q_{\alpha}),

on [t0,t~)(t~,t1]\left[t_{0},\tilde{t}\right)\cup\left(\tilde{t},t_{1}\right]. Similarly, the conditions for the elastic impact read

p(t~+)p(t~)(TQΔQ)=(TQ)+ΔQ,\displaystyle p\left(\tilde{t}^{+}\right)-p\left(\tilde{t}^{-}\right)\in(T\partial Q\cap\Delta_{Q})^{\circ}=(T\partial Q)^{\circ}+\Delta_{Q}^{\circ},
H(qα(t~),p(t~))=H(qα(t~+),p(t~+))\displaystyle H\left(q_{\alpha}\left(\tilde{t}^{-}\right),p\left(\tilde{t}^{-}\right)\right)=H\left(q_{\alpha}\left(\tilde{t}^{+}\right),p\left(\tilde{t}^{+}\right)\right)
q˙α(t~+)ΔQ.\displaystyle\dot{q}_{\alpha}\left(\tilde{t}^{+}\right)\in\Delta_{Q}.

Energy balance: It may be shown that the conservation of the energy along the solutions, H˙(qα,p)=0\dot{H}(q_{\alpha},p)=0, is redundant, as it may be obtained from the remaining equations.

For unconstrained systems, i.e., ΔQ=TQ\Delta_{Q}=TQ, the Hamilton–d’Alembert–Pontryagin principle in the phase space reduces to the Hamilton–Pontryagin principle in the phase space, and the implicit Hamilton equations of motion read as

p˙=Hq(qα,p),q˙α=Hp(qα,p).\dot{p}=-\frac{\partial H}{\partial q}(q_{\alpha},p),\qquad\dot{q}_{\alpha}=\frac{\partial H}{\partial p}(q_{\alpha},p).

Hiperregular Lagrangians: When L:TQL:TQ\to\mathbb{R} is a hyperregular Lagrangian, i.e., when the Legendre transform 𝔽L:TQTQ\mathbb{F}L:TQ\to T^{*}Q is an isomorphism, then both the Lagrangian and the Hamiltonian approaches are equivalent. Namely, the Lagrangian LL induces the Hamiltomian H:TQH:T^{*}Q\to\mathbb{R} given by

H(pq)=E((𝔽L)1(pq),pq),pqTQ.H(p_{q})=E\left((\mathbb{F}L)^{-1}(p_{q}),p_{q}\right),\quad p_{q}\in T^{*}Q.

By recalling the local expression of the Legendre transform,

𝔽L(q,v)=(q,Lv(q,v)),\mathbb{F}L(q,v)=\left(q,\frac{\partial L}{\partial v}(q,v)\right),

it is easy to check that the implicit Euler–Lagrange equations together with the conditions for the elastic impact hold for a (local) curve (vα,pα):[t0,t1]TQTQ(v_{\alpha},p_{\alpha}):[t_{0},t_{1}]\to TQ\oplus T^{*}Q if and only if pα=𝔽L(vα)p_{\alpha}=\mathbb{F}L(v_{\alpha}) and the implicit Hamilton equations together with the conditions for the elastic impact hold for the (local) curve pα=𝔽L(vα):[t0,tq]TQp_{\alpha}=\mathbb{F}L(v_{\alpha}):[t_{0},t_{q}]\to T^{*}Q.

6 Rolling disk hitting a wall

Let us consider a disk rolling without slipping, as in (Yoshimura and Marsden, 2006a, Section 7.1). However, here we assume that there is a wall that the disk may hit (see Anahory Simoes and Colombo (2023)). The configuration space is thus given by

Q={(x,y,θ,φ)2×𝕊1×𝕊1y+Rsinφ10},Q=\{(x,y,\theta,\varphi)\in\mathbb{R}^{2}\times\mathbb{S}^{1}\times\mathbb{S}^{1}\mid y+R\sin\varphi\leq 10\},

where (x,y)(x,y) denotes the contact point of the disk with the ground, θ\theta denotes the angle of rotation and φ\varphi denotes the heading angle of the disk with respect to the xx-axis. The Lagrangian L:TQL:TQ\to\mathbb{R} is given by

L(x,y,θ,φ;vx,vy,vθ,vφ)=12m(vx2+vy2)+12(Ivθ2+Jvφ2),L(x,y,\theta,\varphi;v_{x},v_{y},v_{\theta},v_{\varphi})=\frac{1}{2}m\left(v_{x}^{2}+v_{y}^{2}\right)+\frac{1}{2}\left(I\,v_{\theta}^{2}+J\,v_{\varphi}^{2}\right),

where m,I,J+m,I,J\in\mathbb{R}^{+} are the mass and the moments of inertia of the disk, respectively. For each (vq,pq)=(x,y,θ,φ;vx,vy,vθ,vφ;px,py,pθ,pφ)TQTQ(v_{q},p_{q})=(x,y,\theta,\varphi;v_{x},v_{y},v_{\theta},v_{\varphi};p_{x},p_{y},p_{\theta},p_{\varphi})\in TQ\oplus T^{*}Q, the energy reads

E(vq,pq)=\displaystyle E(v_{q},p_{q})= pxvx+pyvy+pθvθ+pφvφ\displaystyle p_{x}\,v_{x}+p_{y}\,v_{y}+p_{\theta}\,v_{\theta}+p_{\varphi}\,v_{\varphi}
12m(vx2+vy2)12(Ivθ2+Jvφ2).\displaystyle-\frac{1}{2}m\left(v_{x}^{2}+v_{y}^{2}\right)-\frac{1}{2}\left(I\,v_{\theta}^{2}+Jv_{\varphi}^{2}\right).

Non-holonomic constraint: The non-slipping condition reads vx=Rvθcosφ,v_{x}=R\,v_{\theta}\cos\varphi, vy=Rvθsinφv_{y}=R\,v_{\theta}\sin\varphi, where R+R\in\mathbb{R}^{+} is the radius of the disk, thus yielding following non-holonomic constraint:

ΔQ=span{θ+Rcosφx+Rsinφy,φ}.\Delta_{Q}=\operatorname{span}\{\partial_{\theta}+R\cos\varphi\,\partial_{x}+R\sin\varphi\,\partial_{y},\partial_{\varphi}\}.

The annihilator is easily seen to be

ΔQ=span{dxRcosφdθ,dyRsinφdθ}.\Delta_{Q}^{\circ}=\operatorname{span}\{dx-R\cos\varphi\,d\theta,dy-R\sin\varphi\,d\theta\}.

On the other hand, the boundary of the configuration manifold is given by

Q={(x,y,θ,φ)2×𝕊1×𝕊1y+Rsinφ=10},\partial Q=\{(x,y,\theta,\varphi)\in\mathbb{R}^{2}\times\mathbb{S}^{1}\times\mathbb{S}^{1}\mid y+R\sin\varphi=10\},

whose tangent bundle reads

TQ=span{x,θ,φRcosφy}.T\partial Q=\operatorname{span}\{\partial_{x},\partial_{\theta},\partial_{\varphi}-R\cos\varphi\,\partial_{y}\}.

Therefore, its annihilator reads

(TQ)=span{dy+Rcosφdφ}.(T\partial Q)^{\circ}=\operatorname{span}\{dy+R\cos\varphi\,d\varphi\}.

Dynamical equations: The implicit Euler–Lagrange equations with collisions for a curve

(x,y,θ,φ;vx,vy,vθ,vφ;px,py,pθ,pφ):[t0,t1]TQTQ(x,y,\theta,\varphi;v_{x},v_{y},v_{\theta},v_{\varphi};p_{x},p_{y},p_{\theta},p_{\varphi}):[t_{0},t_{1}]\to TQ\oplus T^{*}Q

given in (9) read

{Rp˙xcosφ+Rp˙ysinφ+p˙θ=0,p˙φ=0,vx=Rvθcosφ,vy=Rvθsinφ,px=mvx,py=mvy,pθ=Ivθ,pφ=Jvφ,vx=x˙,vy=y˙,vθ=θ˙,vφ=φ˙,\left\{\begin{array}[]{ll}R\,\dot{p}_{x}\cos\varphi+R\,\dot{p}_{y}\sin\varphi+\dot{p}_{\theta}=0,&\dot{p}_{\varphi}=0,\\ v_{x}=R\,v_{\theta}\cos\varphi,&v_{y}=R\,v_{\theta}\sin\varphi,\\ p_{x}=m\,v_{x},&p_{y}=m\,v_{y},\\ p_{\theta}=I\,v_{\theta},&p_{\varphi}=J\,v_{\varphi},\\ v_{x}=\dot{x},&v_{y}=\dot{y},\\ v_{\theta}=\dot{\theta},&v_{\varphi}=\dot{\varphi},\\ \end{array}\right.

on [t0,t1]{t~}[t_{0},t_{1}]-\left\{\tilde{t}\right\}.

Conditions for the impact: The impact condition at t=t~t=\tilde{t} given in (10) reads

{px+px=λ1,py+py=λ0+λ2,pθ+pθ=λ1Rcosφλ2Rsinφ,pφ+pφ=λ0Rcosφ,\left\{\begin{array}[]{l}p_{x}^{+}-p_{x}^{-}=\lambda^{1},\\ p_{y}^{+}-p_{y}^{-}=\lambda^{0}+\lambda^{2},\\ p_{\theta}^{+}-p_{\theta}^{-}=-\lambda^{1}\,R\cos\varphi-\lambda^{2}\,R\sin\varphi,\\ p_{\varphi}^{+}-p_{\varphi}^{-}=\lambda^{0}\,R\cos\varphi,\end{array}\right.

where we denote px+=px(t~+)p_{x}^{+}=p_{x}\left(\tilde{t}^{+}\right), etc., and λ0,λ1,λ2\lambda^{0},\lambda^{1},\lambda^{2}\in\mathbb{R} are the Lagrange multipliers. Similarly, the condition (11) reads

{vx+=λ3Rcosφ,vθ+=λ3,vy+=λ3Rsinφ,vφ+=λ4\left\{\begin{array}[]{l}v_{x}^{+}=\lambda^{3}\,R\cos\varphi,\quad v_{\theta}^{+}=\lambda^{3},\\ v_{y}^{+}=\lambda^{3}\,R\sin\varphi,\quad v_{\varphi}^{+}=\lambda^{4}\end{array}\right.

where vx+=vx(t~+)v_{x}^{+}=v_{x}\left(\tilde{t}^{+}\right), etc., and λ3,λ4\lambda^{3},\lambda^{4}\in\mathbb{R} are the Lagrange multipliers.

For instance, when the disk hits the wall orthogonally, i.e., when φ(t~)=π/2\varphi\left(\tilde{t}\right)=\pi/2, the only admissible solution of the impact equations is

px+=px=0,py+=py,pθ+=pθ,pφ+=pφ.\begin{array}[]{ll}p_{x}^{+}=p_{x}^{-}=0,\quad p_{y}^{+}=-p_{y}^{-},\quad p_{\theta}^{+}=-p_{\theta}^{-},\quad p_{\varphi}^{+}=p_{\varphi}^{-}.\end{array}

References

  • Anahory Simoes and Colombo (2023) Anahory Simoes, A. and Colombo, L. (2023). Hamel equations and quasivelocities for nonholonomic systems with inequality constraints. arXiv preprint arXiv:2303.17920.
  • Brogliato (1999) Brogliato, B. (1999). Nonsmooth mechanics, volume 3. Springer.
  • Clark and Bloch (2019) Clark, W. and Bloch, A. (2019). The bouncing penny and nonholonomic impacts. In 2019 IEEE 58th Conference on Decision and Control (CDC), 2114–2119. IEEE.
  • Colombo et al. (2022) Colombo, L., de León, M., Irazú, M.E.E., and López-Gordón, A. (2022). Geometric hamilton-jacobi theory and integrability for nonholonomic and forced hybrid systems. arXiv preprint arXiv:2211.06252.
  • Cortés and Vinogradov (2006) Cortés, J. and Vinogradov, A.M. (2006). Hamiltonian theory of constrained impulsive motion. Journal of mathematical physics, 47(4).
  • Fetecau et al. (2003) Fetecau, R.C., Marsden, J.E., Ortiz, M., and West, M. (2003). Nonsmooth lagrangian mechanics and variational collision integrators. SIAM Journal on Applied Dynamical Systems, 2(3), 381–416. 10.1137/S1111111102406038. URL https://doi.org/10.1137/S1111111102406038.
  • Goodman and Colombo (2020) Goodman, J.R. and Colombo, L.J. (2020). On the existence and uniqueness of poincaré maps for systems with impulse effects. IEEE Transactions on Automatic Control, 65(4), 1815–1821. 10.1109/TAC.2019.2941446.
  • Haddad et al. (2006) Haddad, W.M., Chellaboina, V., and Nersesov, S.G. (2006). Impulsive and hybrid dynamical systems: stability, dissipativity, and control. Princeton University Press.
  • Ibort et al. (1998) Ibort, A., de León, M., Lacomba, E.A., Marrero, J.C., de Diego, D.M., and Pitanga, P. (1998). Geometric formulation of mechanical systems subjected to time-dependent one-sided constraints. Journal of Physics A: Mathematical and General, 31(11), 2655.
  • Lacomba and Tulczyjew (1990) Lacomba, E. and Tulczyjew, W. (1990). Geometric formulation of mechanical systems with one-sided constraints. Journal of Physics A: Mathematical and General, 23(13), 2801.
  • Westervelt et al. (2018) Westervelt, E.R., Grizzle, J.W., Chevallereau, C., Choi, J.H., and Morris, B. (2018). Feedback control of dynamic bipedal robot locomotion. CRC press.
  • Yoshimura and Marsden (2006a) Yoshimura, H. and Marsden, J. (2006a). Dirac structures in Lagrangian mechanics Part I: Implicit Lagrangian systems. J. Geom. Phys., 57(1), 133–156.
  • Yoshimura and Marsden (2006b) Yoshimura, H. and Marsden, J.E. (2006b). Dirac structures in lagrangian mechanics part ii: Variational structures. Journal of Geometry and Physics, 57(1), 209–250.