0% found this document useful (0 votes)
21 views13 pages

A Note On Hybrid Routh Reduction For Time-Dependent

This document discusses Routh reduction for hybrid time-dependent mechanical systems. It provides general conditions for when it is possible to reduce symmetries in such systems, extending previous results. It also illustrates the method using the example of a billiard with moving walls.

Uploaded by

pepoahmed12
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
21 views13 pages

A Note On Hybrid Routh Reduction For Time-Dependent

This document discusses Routh reduction for hybrid time-dependent mechanical systems. It provides general conditions for when it is possible to reduce symmetries in such systems, extending previous results. It also illustrates the method using the example of a billiard with moving walls.

Uploaded by

pepoahmed12
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 13

A note on Hybrid Routh reduction for time-dependent

Lagrangian systems
L. J. Colomboa , M. E. Eyrea Irazúb and E. Garcı́a-Toraño Andrésc
a
arXiv:2003.07484v1 [math-ph] 17 Mar 2020

Instituto de Ciencias Matemáticas, Consejo Superior de Investigaciones Cientı́ficas


Calle Nicolás Cabrera 13-15, Cantoblanco, 28049, Madrid, Spain
b
Departamento de Matemática, Universidad Nacional de La Plata
Calle 1 y 115, La Plata 1900, Buenos Aires, Argentina
c
Departamento de Matemática, Universidad Nacional del Sur
Av. Alem 1253, 8000 Bahı́a Blanca, Argentina

Abstract
This note discusses Routh reduction for hybrid time-dependent mechanical systems. We give
general conditions on whether it is possible to reduce by symmetries a hybrid time-dependent La-
grangian system extending and unifying previous results for continuous-time systems. We illustrate
the applicability of the method using the example of a billiard with moving walls.

Keywords: Symmetries, cyclic coordinates, hybrid mechanical systems, conserved quantities,


cosymplectic reduction.

2010 Mathematics Subject Classification: 70S10, 37J15, 70H03.

1 Introduction
One of the first instances of symmetry reduction for mechanical systems can be found in the pioneering
works of Routh in the second half of the 19th century. Routh’s procedure, which is Lagrangian in
nature, deals with the so-called cyclic or ignorable variables (variables on which the Lagrangian does not
depend explicitily): cyclic variables lead to conserved momenta, and these allow to construct a reduced
Lagrangian function, nowadays known as the Routhian. The Euler-Lagrange equations for the Routhian
involve fewer variables than the un-reduced Euler-Lagrange equations, and the solutions of the dynamical
equations for the Routhian, together with a precribed value of momenta for the un-reduced system, can
be used to reconstruct solutions of the original Lagrangian system.
Since then, a number of papers have been devoted to the geometrization and generalization of this
reduction technique (e.g. [10, 14, 18]) but, to the best of our knowledge, the hybrid analogue has not
been fully discussed in the literature. An hybrid scheme for Routh reduction for autonomous hybrid
Lagrangian systems with cyclic variables is found in [3]; this technique (or a variant of it, the so-called
functional Routh reduction) has been successfully applied to the study control strategies of certain bipedal
walkers, see e.g. [4, 15]. This work attempt to go one step further and discuss Routh reduction for non-
autonomous hybrid systems. We remark that some of the techniques outlined here have been recently
applied to the study of periodic orbits in reduced hybrid Lagrangian systems [6, 7, 8, 13].
L. Colombo was partially supported by I-Link Project (Ref: linkA20079) from CSIC, Ministerio de Economia, Industria y
Competitividad (MINEICO, Spain) under grant MTM2016- 76702-P; “Severo Ochoa Programme for Centres of Excellence”
in R&D (SEV-2015-0554). The project that gave rise to these results received the support of a fellowship from “La Caixa’
Foundation (ID 100010434). M.E. Eyrea Irazú was partially supported by CONICET Argentina.

1
The material is organized as follows. Section 2 contains some preliminary results on the geometry of
time-dependent Lagrangian systems. Section 3 discusses the notion of hybrid time-dependent mechanical
system with symmetry. The reduction scheme is proposed in Section 4. Finally, Section 5 contains an
illustrative example: the time-dependent system described by a rough billiard with moving walls.

2 Time-dependent hybrid systems and symmetries


We start recalling some basic facts about time-dependant mechanics. Our starting point is a time-
dependent Lagrangian L : R × T Q → R; we will denote by FL : R × T Q → R × T ∗ Q the Legendre
transformation associated with L, which is the map (t, q, q̇) 7→ (t, q, p = ∂L/∂ q̇).
We will assume that the Lagrangian is hyperregular, i.e. that FL is a diffeomorphism between R × T Q
and R × T ∗ Q (this is always the case for mechanical Lagrangians). One can then work out the velocities
q̇ in terms of (t, q, p) using the inverse of FL and define the Hamiltonian function H : R × T ∗ Q → R as
H(t, q, p) = hp, q̇(t, q, p)i − L(q, q̇(t, q, p)).

Among the several geometric approaches to non-autonomous mechanics [12], we will be using one
based on the notion of cosymplectic manifold (see Appendix for details). We first describe the procedure
for Hamiltonian mechanics. To describe the dynamics of a non-autonomous Hamiltonian system, one
begins by considering the manifold R × T ∗ Q equipped with the canonical cosymplectic structure
Ω = dq ∧ dp, η = dt,
with Reeb vector field T = ∂t . Given a Hamiltonian H(t, q, p), the Hamiltonian vector field XH is the
vector field on R × T ∗ Q defined by
iXH Ω = dH − T(H)η, iXH η = 0.
The evolution vector field corresponding to the Hamiltonian H, denoted by ZH , is defined by
∂ ∂H ∂ ∂H ∂
ZH = T + XH = + − ,
∂t ∂p ∂q ∂q ∂p
and its integral curves are solutions of the Hamilton equations q̇ = ∂p H, ṗ = −∂q H, where the “∂t ”
component of ZH sets the evolution parameter of the solutions to be the “time” t.
In the Lagrangian picture, under the assumption of hyperregularity discussed above, a similar pro-
cedure can be used. The manifold R × T Q can be endowed with the following cosymplectic structure
 
∂L
ΩL = FL∗ (dq ∧ dp) = dq ∧ d , η = dt, (1)
∂ q̇
which depends on the Lagrangian (note that we will be using the same symbol η = dt for two different
1-forms on different manifolds). One then constructs the energy function EL : R × T Q → R given by
EL (t, q, q̇) = hFL(t, q, q̇), q̇i − L(t, q, q̇),
and obtains the Hamiltonian vector field associated to the cosymplectic structure (1) and Hamiltonian
EL . This leads to an evolution vector field, that we denote ZL , given by
∂ ∂ ∂
ZL = + q̇ + Γ(t, q, q̇) ,
∂t ∂q ∂ q̇
where Γ(t, q, q̇) is obtained by isolating q̈ from the standard Euler-Lagrange equations
 
d ∂L ∂L
− = 0.
dt ∂ q̇ ∂q

Finally, the well-known equivalence between the Lagrangian and Hamiltonian dynamics in the hyper-
regular case is achieved via FL.

2
Proposition 1. The tangent map of FL maps ZL onto ZH , i.e. (T FL)(ZL ) = XH . In particular, the
flow of ZL is mapped onto the flow of ZH .

Proof: The evolution vector field ZH is characterized by iXH Ω = dH − T(H)η and iXH η = 1. Now:

(FL)∗ (iZH Ω) = (FL)∗ (dH − T(H)η) = (FL)∗ (dH) − (FL)∗ (T(H)η)


= d((FL)∗ H) − T(FL(H))η = d(EL ) − T(EL )η
= iZL ΩL .

This means iZL ΩL = (FL)∗ (iZH Ω) = i(FL−1 )∗ ZH (FL∗ Ω) = i(FL−1 )∗ ZH ΩL . In a similar way one shows
that iZL η = i(FL−1 )∗ ZH η. This implies ZL = (FL−1 )∗ ZH , that is, (FL)∗ ZL = ZH . 

3 Simple Hybrid time-dependent Mechanical Systems and Sym-


metries
Roughly speaking, the term hybrid system refers to a dynamical system which exhibits both continuous
and discrete behavior. In the literature, one finds slightly different definitions of hybrid system depending
on the specific class of applications of interest. For simplicity, and following [2] and [3], we will restrict
ourselves to the so-called simple hybrid mechanical systems in Lagrangian form. We will extend the
definition in order to include time dependence on both the Lagrangian and the switching surface (see
below).

Definition 1. A simple hybrid time-dependent Lagrangian system is a tuple L = (Q, L, S, R) where

(i) Q is a differentiable manifold,


(ii) L : R×T Q → R is a time-dependent Lagrangian (recall that we will always assume hyperregularity),
(iii) S is an embedded submanifold of R × T Q with co-dimension one called the switching surface
(sometimes referred to as the guard ),

(iv) R : S → R × T Q is a smooth map called the reset map (often called the impact map).

The triple (Q, S, R) alone is referred to as an hybrid manifold.

In a similar way one defines a hybrid time-dependent Hamiltonian system as a tuple H = (Q, H, SH , RH ),
where H : R × T ∗ Q → R is a time-dependent Hamiltonian function and the elements of the hybrid man-
ifold are now defined on T ∗ Q instead of T Q.
We have recalled in Proposition 1 that in the hyperregular case there is an equivalence between the
Lagrangian and the Hamiltonian descriptions of a mechanical system. We will now extend this equivalence
to the hybrid setting (among other things, this clarifies the relation between the results in [2] and [3]).
We need the following definition:
Definition 2. A hybrid flow for L is a tuple χL = (Λ, J , C ), where

• Λ = {0, 1, 2, ...} ⊆ N is a finite (or infinite) indexing set,


• J = {Ii }i∈Λ a set of intervales, called hybrid intervals where Ii = [τi , τi+1 ] if i, i + 1 ∈ Λ and
IN −1 = [τN −1 , τN ] or [τN −1 , τN ) or [τN −1 , ∞) if |Λ| = N , N finite, with τi , τi+1 , τN ∈ R and
τi ≤ τi+1 ,
• C = {ci }i∈Λ is a collection of solutions for the vector field ZL specifying the continous-time dy-
namics, i.e., c˙i = ZL (ci (t)) for all i ∈ Λ, and such that for each i, i + 1 ∈ Λ,

3
(i) ci (τi+1 ) ∈ S,
(ii) R(ci (τi+1 )) = ci+1 (τi+1 ).

Analogously, one defines the notion of hybrid flow χH for a hybrid time-dependent Hamiltonian
system H . The relation between both the Lagrangian and the Hamiltonian hybrid flows is given by
the following result, where for clarity in the exposition we will use the notation L = (Q, L, S, R),
H = (Q, H, SH , RH ).
Proposition 2. If χL = (Λ, J , C ) is a hybrid flow for L , SH = FL(S), and RH is defined in such a
way that FL ◦ R = RH ◦ FL |S , then χH = (Λ, J , (FL)(C )) with (FL)(C ) = {(FL)(ci )}i∈Λ .

Proof: By Proposition 1, if ci (t) is an integral curve of ZL , c̃i (t) = (FL ◦ ci )(t) is an integral curve
for XH . In this way, if we consider a solution c0 (t) with initial value c0 = (q0 , q̇0 ) defined on [τ0 , τ1 ], then
c̃0 (t) is a solution with initial value c̃0 = (q0 , p0 ) defined on [τ0 , τ1 ]. Likewise for a solution c1 (t) defined
on [τ1 , τ2 ], we get a corresponding solution c̃1 (t) defined on the same hybrid interval [τ1 , τ2 ]. Proceeding
inductively, one finds ci (t) defined on [τi , τi+1 ]. It only remains to check that c̃i (t) satisfies c̃i (τi+1 ) ∈ SH
and RH (c̃i (τi+1 )) = c̃i+1 (τi+1 ), but using the properties of FL, we have that,

(i) c̃i (τi+1 ) = (FL ◦ ci )(τi+1 ) = FL(ci (τi+1 )) and given that ci (τi+1 ) ∈ S then c̃i (τi+1 ) ∈ SH .
(ii) RH (c̃i (τi+1 )) = RH ◦ FL ◦ ci (τi+1 ) = FL ◦ R ◦ ci (τi+1 ) = FL ◦ ci+1 (τi+1 ) = c̃i+1 (τi+1 ). 

Let L = (Q, L, S, R) be a simple hybrid time-dependent Lagrangian system. The starting point for
symmetry reduction is a Lie group action ψ : G × Q → Q of some Lie group G on the manifold Q. We
will assume that all the actions satisfy some regularity conditions as to do reduction (for instance, one
can consider free and proper actions).
There is a natural lift Ψ of the action ψ to the space R × T ∗ Q, the cotangent lift action, defined by
Ψg = T ∗ ψg−1 . It enjoys the following properties [1]:

• Ψ is a cosymplectic action, meaning that Ψ∗g Ω = Ω and Ψ∗ η = η.


• It admits an Ad∗ -equivariant momentum map J : R × T ∗ Q → g∗ given by

hJ(t, q, p), ξi = hp, ξQ i, ∀ξ ∈ g,

where ξQ (q) = d(ψexp(tξ) q)/dt is the infinitesimal generator of ξ ∈ g.

Likewise, ΨT Q denotes the tangent lift action on R × T Q, defined by ΨTg Q = ψg (q, q̇).
To perform a hybrid reduction one needs to impose some compatibility conditions between the action
and the hybrid system (see e.g. [3]). By an hybrid action on the simple hybrid time-dependent Lagrangian
system L = (Q, L, S, R) we mean a Lie group action ψ : G × Q → Q such that

• L is invariant under ΨT Q , i.e. L ◦ ΨT Q = L.


• ΨT Q restricts to an action of G on S.
• R is equivariant with respect to the previous action, namely R ◦ ΨTg Q |S = ΨTg Q ◦ R.

Recall that ΨT Q admits an Ad∗ -equivariant momentum map JL : R × T Q → g∗ given by JL = J ◦ FL.


This follows directly from the invariance of L, since it implies that FL is an equivariant diffeomorphism,
i.e. FL ◦ ΨTg Q = Ψg ◦ FL
FL ◦ ΨTg Q = Ψg ◦ FL.

4
The hybrid equivalent of momentum map is the notion of hybrid momentum map introduced in [2].
In the case of R × T ∗ Q, J is an hybrid momentum map if the diagram

g∗
J J (2)
J|S

i R
R × T ∗Q S R × T ∗Q

commutes, where i : S ,→ R × T ∗ Q denotes the canonical inclusion.

4 Reduction by symmetries of simple hybrid time-dependent


Lagrangian system
Consider a simple hybrid time-dependent Lagrangian system L = (Q, L, S, R) equipped with an hybrid
action ψ. We begin analyzing the reduction of the associated hybrid time-dependent Hamiltonian system
H = (Q, H, SH , RH ) discussed in Proposition 2.
Consider a hybrid regular value µ ∈ g∗ of J : R × T ∗ Q → g∗ , which means that µ is a regular value of
both J and J |S : S → g∗ . When we combine this definition with the commuting diagram (2), we obtain
that the following diagram

i RH |S
J −1 (µ) J |−1
S (µ) J −1 (µ)

i RH
R × T ∗Q S R × T ∗Q

commutes, where J −1 (µ) and J |−1 ∗


S (µ) are embedded submanifolds of R × T Q and S, respectively.

We can apply a hybrid analog of the cosymplectic reduction Theorem [1] to the hybrid time-dependent
Hamiltonian system H (see Appendix). Note that, since L is invariant under ΨT Q , so is the Hamiltonian

H under ΨT Q . The main conclusions are:

(i) The reduced space J −1 (µ)/Gµ (with Gµ the isotropy group of µ under the coadjoint action) is a
cosymplectic manifold, and the reduced cosymplectic structure (ηµ , Ωµ ) is characterized in terms
of the submersion πµ : J −1 (µ) → J −1 (µ)/Gµ and the inclussion iµ : J −1 (µ) ,→ R × T ∗ Q by means
of the relations πµ∗ ηµ = i∗µ η and πµ∗ Ωµ = i∗µ Ω.

(ii) If we denote by Hµ the reduction of H |J −1 (µ) to J −1 (µ)/Gµ , the evolution vector field ZH projects
onto ZHµ . This is the second part of the Cosymplectic reduction Theorem of Albert [1] (see
Appendix).
(iii) J |−1
S (µ) ⊂ SH is Gµ -invariant and hence reduces to a submanifold of the reduced space which we
denote (SH )µ ⊂ J −1 (µ)/Gµ .

(iv) Again, using invariance RH reduces to a map (RH )µ : (SH )µ → J −1 (µ)/Gµ .

The reduction picture in the Lagrangian side can now be obtained from the Hamiltonian one by
adapting the scheme developed in [18] to the cosymplectic setting. The key idea is that, since L is
invarian, the Legendre transformation FL is a diffeomorphism such that:

• It is equivariant with respect to ΨT Q and ΨT Q
,

5
• Preserves the level sets of the momentum map, that is, FL(JL−1 (µ)) = J −1 (µ),
• Relates both cosymplectic structures, that is, (FL)∗ Ω = ΩL and (FL)∗ η = η (FL is sometimes
referred to as a cosymplectomorphism).

It follows that the map FL reduces to a cosymplectomorphism (FL)red between the reduced spaces.
Therefore we get the following commutative diagram of hybrid manifolds:

(R × T Q, S, R) FL
(R × T ∗ Q, SH , RH )

Red. Red.

(FL)red
(JL−1 (µ)/Gµ , Sµ , Rµ ) (J −1 (µ)/Gµ , (SH )µ , (RH )µ )

We now make use of a principal connection on the bundle Q → Q/G to make some further identifi-
cations. Let A : T Q → g∗ be the connection one form, and let us denote by Aµ (·) = hµ, A(·)i the 1-form
on Q obtained by contraction with µ ∈ g∗ . Building on the well-known results on cotangent bundle
reduction it is possible to show that there is an identification

J −1 (µ)/Gµ ' R × T ∗ (Q/G) ×Q/G Q/Gµ .



(3)

This identification is a symplectomorphism when we endow the space on the right hand side of (3) with
the symplectic structures pr∗1 ΩQ/G + pr∗2 Bµ , where pr1 and pr2 are the canonical projections, and Bµ is
the so-called magnetic term, obtained from the reduction of dAµ to Q/Gµ . For details, see [18, 21].
For the Lagrangian side, one needs a further regularity condition, sometimes referred to as G-
regularity, which is satisfied by mechanical Lagrangians. Precisely, one has the following definition [20]
(for an alternative, equivalent definition, see [18]):

Definition 3. Let L be an invariant Lagrangian on T Q and denote by ξQ the infinitesimal generator for
the associated action. We say that L is G-regular if, for each vq ∈ T Q, the map
v
JL q : g → g∗ ,
ξ 7→ JL (vq + ξQ (q)) ,

is a diffeomorphism.

In a nutshell, G-regularity amounts to regularity “with respect to the group variables”. From now on we
will assume that the Lagrangian is G-regular. In this case, there is an identification

JL−1 (µ)/Gµ ' R × T (Q/G) ×Q/G Q/Gµ .




It is possible to interpret the reduced dynamics on this space as being the Lagrangian dynamics of some
regular Lagrangian subjected to a gyroscopic force (arising from the magnetic term) if one works in
the more general class of magnetic Lagrangians [19], which in the present situation should be extended
to include time-dependent Lagrangians. The so-called magnetic Lagrangian systems are a wide class
of Lagrangian systems on which the Lagrangian function might not depend on some of the velocities,
and which may as well include a force term given by a 2-form. The framework of magnetic Lagrangian
systems is very convenient when carrying out Routh reduction, since the reduced system is not, in
general, a standard Lagrangian system. We remark that Routh reduction has been extended to magnetic
Lagrangian system, and this permits to carry out Routh reduction by stages [20].
The function which plays the role of the reduced Lagrangian is the Routhian, and it is defined as (the
reduction of) the Gµ -invariant function
Lµ = L − Aµ (4)

6
restricted to JL−1 (µ). The next diagram summarizes the situation:

R × TQ FL
R × T ∗Q

Red. Red.

FLµ
R × T ∗ (Q/G) ×Q/G Q/Gµ
 
R × T (Q/G) ×Q/G Q/Gµ

We will now focus on the particular case of cyclic coordinates, which is the one in encountered in the
reeferences motivating this note. More details and examples on the general situation can be found in [13].

Cyclic coordinates in simple hybrid time-dependent Lagrangian system The case G = S1


corresponds to the notion of cyclic coordinates (the case G = R is analogous; if G is a product one
can iterate the procedure). The reduced space JL−1 (µ)/Gµ can be identified with R × T (Q/S1 ) and the
reduced dynamics is Lagrangian with respect to the reduced Lagrangian Lµ = L − Aµ on R × T (Q/S1 ).
The reduced switching Sµ can be identified with a submanifold of R × T (Q/S1 ) and the reset map is
identified with a map Rµ : Sµ → R × T (Q/S1 ). We will use the same notations for both of them.
A case of special interest with regards to applications is when Q = S1 ×M , where M is called the shape
space and the action is simply (θ, x) 7→ (θ + α, x). This is often the situation when dealing with simple
models of bipedal walkers, see e.g. [4, 15]. From now on, we will assume we work in this setting. While
this is indeed a strong assumption, it is always the case locally, so as long as it applies to the domain of
interest of an specific problem the procedure applies. The Lagrangian has a cyclic coordinate θ, i.e. L
is a function of the form L(t, θ̇, x, ẋ). The conservation of the momentum map J = µ reads ∂L/∂ θ̇ = µ,
and one can use this relation to express θ̇ as a function of the remaining -non cyclic- coordinates and
their velocities, and the prescribed regular value of the momentum map µ. We point out that it is at this
stage that G-regularity of L is used: it guarantees that θ̇ can be worked out in terms of x, ẋ and µ. If
one chooses the cannonical flat connection on Q → Q/S1 = M , then the Routhian can be computed as
h i
Lµ (t, x, ẋ) = L(t, θ̇, x, ẋ) − µθ̇ , (5)
θ̇=θ̇(t,x,ẋ,µ)

where the notation means that we have everywhere expressed θ̇ as a function of (t, x, ẋ, µ). Note that (5)
is the classical definition of the Routhian [23]. Besides, since the connection is flat, one has no force term
in the reduced dynamics.
Let us first consider the case in which the momentum map is preserved in the collisions with the
switching surface (elastic case). We then have:

Proposition 3. In the situation above:

(i) Any solution of L = (Q, L, S, R) with momentum µ projects onto a solution of Lµ = (Q/S1 , Lµ , Sµ , Rµ ).
(ii) Any solution of Lµ = (Q/S1 , Lµ , Sµ , Rµ ) is the projection of a solution of L = (Q, L, S, R) with
momentum µ.

Collisions with the switching surface will, in general, modify the value of the momentum map (non-
elastic case). Therefore, if J = {Ii }i∈Λ is the hybrid interval (see Definition 2), the Routhian has to
be defined in each Ii taking into account the value of the momentum µi after the collision at time τi .
Note that this also has an influence in the way the reset map R and the switching S are reduced. Let us
denote: (1) µi the momentum of the system in Ii = [τi , τi+1 ], (2) Rµi the reduction of R |J −1 (µi ) , and (3)
Sµi the reduction of J |−1
S (µi ), There is a sequence of reduced simple hybrid time-dependent Lagrangian

7
systems (“coll” stands for collision):

Red.
[τ0 , τ1 ] (Q/S1 , Lµ0 , Sµ0 , Rµ0 )
coll. coll.

Red.
[τ1 , τ2 ] (Q/S1 , Lµ1 , Sµ1 , Rµ1 )
coll. coll.
Red.
(. . . ) (. . . )

The fact that the momentum will, in general, change with the collisions makes the reconstruction
procedure more challenging. If one wishes, as usual, to use a reduced solution to reconstruct the original
dynamics, one needs to compute the reduced hybrid data after each collision. This means that once
the reduced solution has been obtained between two collison events, say at t = τn and t = τn+1 , one
should reconstruct this solution to obtain the new momentum after the collision at τn+1 and use this
new momentum to build a new reduced hybrid system whose solution should be obtained until the next
collision eventy at τn+2 , and so on. As usual, the reconstruction procedure from the reduced hybrid flow
to the hybrid flow involves an integration at each stage in the previous diagram of the cyclic variable using
the solution of the reduced simple hybrid time-dependent Lagrangian system. Essentially, this accounts
to imposing the momentum constraint on the reconstructed solution.
Remark 1. Proposition 3 is easily adapted to more general scenarios. If Q is not a product, one needs
to compute the Routhian using the general expression (4) and consider the reduced simple hybrid time-
dependent Lagrangian system with a force term. If G is non-Abelian one can use the class of magnetic
Lagrangian systems to describe the reduced dynamics. Coordinate formulae for the reduced dynamics
and reconstruction in this more general case can be obtained using the techniques in [10].

5 Example: a rough billiard with moving walls


Consider a particle of mass m in the plane which is free to move inside the surface defined by a circle
whose radius varies in time (Figure 1) according to a given function f (t), i.e. x2 + y 2 = f (t). The surface

f (t)

Figure 1: A “billiard” with moving walls

of the “billiard” is assumed to be rough in such a way that the friction is proportional to the velocity (the
dissipation is of Rayleigh type, see e.g. [5]). This example falls in the category of mechanical systems
with one-sided constraints, see e.g. [9, 16, 17] for alternative approaches.
The equations of motion for the particle off the boundary are

mẍ = −cẋ, mÿ = −cẏ. (6)

8
for some constant c > 0 (a dissipation coefficient). To fit these equations into Lagrangian form, one
considers the time-dependent Lagrangian L : R × T R2 → R given by
 c  1 
2 2
L(t, x, y, ẋ, ẏ) = exp t m(ẋ + ẏ ) .
m 2

It is easy to check that the Euler-Lagrange equations for L give the desired equations of motion. The
guard is the subset of R × T R2 ' R × R2 × R2 given by

S = (R × T Q) ∩ {x2 + y 2 = f (t), (ẋ, ẏ) · (x, y) ≥ 0}.

This set S describes the situation in which the particle hits the moving boundary while heading “out-
wards” the billiard. For simplicity in the definition of the switching surface, we assume that f (t) is
increasing: this guarantees the particle only hits the boundary when the boundary is also moving out-
wards. Under the assumption of an elastic collision, the reset map

(t, x, y, ẋ− , ẏ − ) 7→ (t, x, y, ẋ+ , ẏ + )

is given by [16]:

f˙(t) − 2(xẋ− + y ẏ − ) f˙(t) − 2(xẋ− + y ẏ − )


ẋ+ = ẋ− + x, ẏ + = ẏ − + y.
f (t) f (t)

Note that, if f (t) = constant, the reset map agrees with the one given for a constraint function
h : Q → R in [3] in the case of a perfectly elastic impact. Introducing polar coordinates the Lagrangian
becomes  c m
L(t, r, θ, ṙ, θ̇) = exp t (ṙ2 + r2 θ̇2 )
m 2
for which θ is a cyclic coordinate. In this case JL is time dependent and represents a damped angular
momentum (mr2 θ̇ decays exponentially in time as a result of the friction):
c 
JL (t, r, ṙ, θ, θ̇) = exp t mr2 θ̇. (7)
m

A computation reveals that the reset map, in polar coordinates, takes the form (observe that 2(xẋ− +
y ẏ ) is nothing but 2rṙ− ):

(ṙ+ )2 = (ṙ− )2 +
!
r ˙ (f˙(t) − 2rṙ− )r
+ (f (t) − 2rṙ− ) 2ṙ− + , (8)
f (t) f (t)
θ̇+ = θ̇− .

It is understood that the “minus” square root is taken in ṙ+ (the particle bounces on the boundary
after the collision). The assumption of elastic collision implies, in particular, that the momentum map is
preserved. This is clear since r and θ̇ do not change with the collision (8).
Using (7), the Routhian takes the form

1 c  µ2  c 
Lµ (t, r, ṙ) = exp t mṙ2 − exp − t .
2 m 2mr2 m
The reduced reset map is determined by the expression (8) for ṙ+ (note that the expression drops to the
quotient since it only involves r, ṙ and f (t)). The reduced switching surface is Sµ = {r2 = f (t), ṙ > 0}.
One then obtains the following simple hybrid time-dependent Lagrangian system L = (Qred , Lµ , Sµ , Rµ ),
with Qred ' R+ parametrized by the radial coordinate r. Figures 2 and 3 show numerical results using
Python for two different values of the dissipation parameter c. The remaining parameters are the same

9
2.00 Exact 1.0 Exact
Numerical Numerical
1.75 f( t )

1.50 0.5

1.25

0.0
2

y
1.00
r

0.75
−0.5
0.50

0.25
−1.0
0.00
0 5 10 15 20 25 30 35 −1.0 −0.5 0.0 0.5 1.0
t x

Figure 2: Simulation for c = 0.25. The figure in the left corresponds with the reduced trajectory while
the figure to the right corresponds with the reconstructed sulution

for both simulations: m = 1, r(0) = 0.5590, ṙ(0) = 2.8621, θ(0) = 1.1071 (rad), θ̇(0) = −3.0400 (rad/s),
and the function f (t) equals
f (t) = 2 − exp(t/10).
The reduced dynamics corresponding to Lµ is solved numerically (dashed black line) and used to integrate
(numerically) the reconstruction equation (7)
 c  µ
θ̇ = exp − t ,
m mr2
with µ determined from the initial conditions. Since (6) admits the explicit solution
 c   c 
x(t) = x0 exp − t , y(t) = y0 exp − t ,
m m
a comparison with the analytical solution (solid teal line) can be easily obtained. Note, however, that
the impact times on which the particle bounces are obtained numerically also in this “analytical” case.

2.00 Exact Exact


Numerical Numerical
f( t )
1.0
1.75

1.50
0.5

1.25
2

0.0
y

1.00
r

0.75
−0.5
0.50

0.25 −1.0

0.00
0 10 20 30 40 −1.0 −0.5 0.0 0.5 1.0
t x

Figure 3: Simulation for c = 0.10. he figure in the left corresponds with the reduced trajectory while the
figure to the right corresponds with the reconstructed sulution

10
6 Conclusion
We have extended Routh reduction to hybrid time-dependent mechanical systems. This extension re-
covers previous results on Routh reduction for hybrid systems in an intrinsic way, and hence broadens
the applicability of the reduction procedure. The technique has been illustrated both analitically and
numerically in an illustrative example.

Appendix: Cosymplectic reduction theorem


An almost cosymplectic structure on a manifold Q of odd dimension 2n + 1 is a pair (η, Ω), where η is a
1-form and Ω is a 2-form such that η ∧ Ωn is a volume form on Q. The structure is said to be cosymplectic
if η and Ω are closed. A manifold endowed with a cosymplectic structure is referred to as a cosymplectic
manifold.
A cosymplectic structure (η, Ω) on Q induces an isomorphism of C ∞ (Q)-modules [(η,ω) : X(Q) →
1
Ω (Q) defined by
[(η,Ω) (X) = iX Ω + η(X)η,

for every vector field X ∈ X(Q). The vector field T = [−1 (η,Ω) (η) on Q is called the Reeb vector field of
(Q, η, Ω) and it can characterized by the following conditions
iT Ω = 0, η(T) = 1.

Let G be a Lie group. An action φ : G × Q → Q of a Lie group G on a cosymplectic manifold (Q, η, Ω)


is said to be cosymplectic if φg : Q → Q is a cosymplectic map for any g ∈ G, i.e. if
φ∗g η = η, φ∗g Ω = Ω.
Given a cosymplectic action on Q, a smooth map J : Q → g∗ is said to be a momentum map if the
infinitesimal generator ξQ ∈ X(Q) of the action associated with any ξ ∈ g is the Hamiltonian vector field
of the function Jξ : Q → R defined by the natural pointwise pairing Jξ (q) = hJ(q), ξi. The momentum
map is Ad∗ -equivariant (or equivariant, for short) if it is equivariant with respect to the action φ and to
the coadjoint action Ad∗ : G × g∗ → g∗ , that is
J(φg (q)) = Ad∗g−1 (J(q)).

Let µ ∈ g∗ be a weakly regular value of an equivariant momentum map J : Q → g∗ . We denote by


Gµ the isotropy group of µ with respect to the coadjoint action, i.e. Gµ = {g ∈ G : Ad∗g µ = µ}.
The action φ induces an action φ : Gµ × J −1 (µ) → J −1 (µ) of Gµ on the submanifold J −1 (µ).
Following [1] we will say that this action is quotientant if the orbit space Qµ = J −1 (µ)/Gµ admits a
smooth manifold structure and the canonical projection πµ : J −1 (µ) → Qµ is a surjective submersion.
Theorem (Cosymplectic reduction Theorem, [1]). Let φ : G × Q → Q be a cosymplectic action
of a Lie group G on a cosymplectic manifold (Q, η, Ω). Suppose that J : Q → g∗ is an Ad∗ -equivariant
momentum map associated with φ such that T(J) = 0 where T is the Reeb vector field of Q. Let µ ∈ g∗
be a weakly regular value of J such that the induced action of Gµ on J −1 (µ) is quotientant. Then,
Qµ = J −1 (µ)/Gµ is a cosymplectic manifold with cosymplectic structure (ηµ , Ωµ ) characterized by
πµ∗ ηµ = i∗µ η, πµ∗ Ωµ = i∗µ Ω,
where πµ : J −1 (µ) → Qµ is the canonical projection and iµ : J −1 (µ) ,→ Q is the canonical inclusion.
Moreover, the restriction T|J −1 (µ) of T is tangent to J −1 (µ) and πµ -projectable onto the Reeb vector
field Tµ of Qµ .

As a matter of fact, the cosymplectic reduction theorem of Albert [1] can be obtained directly from
the Marsden-Weinstein reduction theorem [22], see [11].

11
Acknowledgments
The authors are indebted to Andrea Bel for the assistance with Python, and to David Martı́n de Diego
for the fruitful discussions. EGTA and EEI are grateful to the ICMAT (Madrid) for its hospitality during
the visits which made this work possible.

References
[1] C. Albert. Le théorème de réduction de Marsden-Weinstein en géométrie cosymplectique et de
contact. J. Geom. Phys., 6(4):627–649, 1989.

[2] A. Ames and S. Sastry. Hybrid cotangent bundle reduction of simple hybrid mechanical systems
with symmetry. In American Control Conference, June 2006.
[3] A. Ames and S. Sastry. Hybrid Routhian reduction of lagrangian hybrid systems. In American
Control Conference, June 2006.

[4] A. Ames, R Gregg, and M. Spong. A geometric approach to three-dimensional hipped bipedal robotic
walking. In Decision and Control, 2007 46th IEEE Conference on, pages 5123–5130. IEEE, 2007.
[5] A. M. Bloch. Nonholonomic mechanics and control, volume 24 of Interdisciplinary Applied Mathe-
matics. Springer, New York, second edition, 2015.
[6] A. Bloch, W. Clark, and L. Colombo. Quasivelocities and symmetries in simple hybrid systems. In
2017 IEEE 56th Annual Conference on Decision and Control (CDC), pages 1529–1534, Dec 2017.
[7] L. Colombo and E. Eyrea Iraz. Symmetries and periodic orbits in simple hybrid routhian systems.
Nonlinear Analysis: Hybrid Systems, 36:100857, 2020.
[8] L. Colombo, W. Clark, and A. Bloch. Time reversal symmetries and zero dynamics for simple
hybrid hamiltonian control systems. In 2018 Annual American Control Conference (ACC), pages
2218–2223, June 2018.
[9] J. Cortés, M. de León, D. Martı́n de Diego, and S. Martı́nez. Mechanical systems subjected to gener-
alized non-holonomic constraints. R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 457(2007):651–
670, 2001.

[10] M. Crampin and T. Mestdag. Routh’s procedure for non-abelian symmetry groups. J. Math. Phys.,
49(3):032901, 28, 2008.
[11] M. de León and M. Saralegi. Cosymplectic reduction for singular momentum maps. J. Phys. A,
26(19):5033–5043, 1993.

[12] A. Echeverrı́a Enrı́quez, M. C. Muñoz Lecanda, and N. Román-Roy. Geometrical setting of time-
dependent regular systems. Alternative models. Rev. Math. Phys., 3(3):301–330, 1991.
[13] E. Eyrea Irazú. Geometric and numerical aspects of mechanical systems with magnetic terms. Ph.D.
thesis. Facultad de Ciencias Exactas, Universidad Nacional de La Plata, 2019.
[14] K. Grabowska and P. Urbański. Geometry of Routh reduction. J. Geom. Mech., 11(1):23–44, 2019.

[15] R. Gregg and M. Spong. Reduction-based control with application to three-dimensional bipedal
walking robots. In 2008 American Control Conference, pages 880–887, June 2008.
[16] A. Ibort, M. de León, E. A. Lacomba, J. C. Marrero, D. Martı́n de Diego, and P. Pitanga. Geometric
formulation of mechanical systems subjected to time-dependent one-sided constraints. J. Phys. A,
31(11):2655–2674, 1998.

12
[17] E. A. Lacomba and W. M. Tulczyjew. Geometric formulation of mechanical systems with one-sided
constraints. J. Phys. A, 23(13):2801–2813, 1990.
[18] B. Langerock, F. Cantrijn, and J. Vankerschaver. Routhian reduction for quasi-invariant La-
grangians. J. Math. Phys., 51(2):022902, 20, 2010.
[19] B. Langerock, E. Garcı́a-Toraño Andrés, and F. Cantrijn. Routh reduction and the class of magnetic
Lagrangian systems. J. Math. Phys., 53(6):062902, 19, 2012.
[20] B. Langerock, T. Mestdag, and J. Vankerschaver. Routh reduction by stages. SIGMA Symmetry
Integrability Geom. Methods Appl., 7:Paper 109, 31, 2011.

[21] J. E. Marsden, G. Misiolek, J P. Ortega, M. Perlmutter, and T. S. Ratiu. Hamiltonian reduction by


stages, volume 1913 of Lecture Notes in Mathematics. Springer, Berlin, 2007.
[22] J. E. Marsden and A. Weinstein. Reduction of symplectic manifolds with symmetry. Rep. Mathe-
matical Phys., 5(1):121–130, 1974.

[23] L. A. Pars. A treatise on analytical dynamics. John Wiley & Sons, Inc., New York, 1965.

13

You might also like