Sie sind auf Seite 1von 36

Home Search Collections Journals About Contact us My IOPscience

The optimal rolling of a sphere, with twisting but without slipping

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2014 Sb. Math. 205 157

(http://iopscience.iop.org/1064-5616/205/2/157)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 150.216.68.200
This content was downloaded on 30/09/2014 at 08:47

Please note that terms and conditions apply.


Sbornik : Mathematics 205:2 157–191 Matematicheskiı̆ Sbornik 205:2 3–38
DOI 10.1070/SM2014v205n02ABEH004370

The optimal rolling of a sphere,


with twisting but without slipping

I. Yu. Beschastnyı̌

Abstract. The problem of a sphere rolling on the plane, with twisting but
without slipping, is considered. It is required to roll the sphere from one
configuration to another in such a way that the minimum of the action is
attained. We obtain a complete parametrization of the extremal trajecto-
ries and analyse the natural symmetries of the Hamiltonian system of the
Pontryagin maximum principle (rotations and reflections) and their fixed
points. Based on the estimates obtained for the fixed points we prove upper
estimates for the cut time, that is, the moment of time when an extremal
trajectory loses optimality. We consider the problem of re-orienting the
sphere in more detail; in particular, we find diffeomorphic domains in the
pre-image and image of the exponential map which are used to construct
the optimal synthesis.
Bibliography: 15 titles.

Keywords: optimal control, geometric methods, symmetries, rolling of


surfaces.

§ 1. Introduction
We consider a mechanical system consisting of a sphere rolling on a plane with
twisting but without slipping. At every moment of time the state of such a system
is characterized by a point on the plane and the orientation of the sphere in space.
It is required to roll the sphere from a given initial state to a given terminal state in
such a way that the minimum of the action is attained. No slipping means that the
contact point of the sphere and the plane has zero instantaneous velocity; presence
of twisting means that the angular velocity vector of the sphere can have arbitrary
direction.
This problem is a natural modification of the problem of the optimal rolling
of a sphere on a plane without twisting or slipping, which was first stated by
Hammersley [1]. Later Arthurs and Walsh [2] obtained equations for extremal
trajectories in terms of quaternions. Jurdjevic [3] conducted a qualitative study
of these trajectories and showed that for optimal rolling the contact point of the
sphere and the plane moves over Euler elastics. Sachkov initiated the study of
This research was supported by the grant of the Government of the Russian Federation for
state support of scientific research (contract no. 14.B25.31.0029) and by the Russian Foundation
for Basic Research (grant nos. 12-01-00913 and 13-01-91162-GFEN_a).
AMS 2010 Mathematics Subject Classification. Primary 49K15; Secondary 70B10, 93B27.


c 2014 Russian Academy of Sciences (DoM), London Mathematical Society, Turpion Ltd.
158 I. Yu. Beschastnyı̌

whether the trajectories obtained are optimal. He analysed the symmetries of


the Hamiltonian system of the Pontryagin maximum principle and obtained upper
estimates for the cut time (the moment of time at which extremal trajectories
lose optimality) [4]. In [5], [6], in collaboration with Mashtakov, he studied the
asymptotics of extremal trajectories and the Maxwell time for a sphere rolling over
sinusoids of small amplitude.
The investigation into the case without twisting is still not complete, and is prob-
ably one of the most difficult left-invariant problems in optimal control [5]. This
fact is mainly related to the complicated structure of the extremal trajectories,
which are described by elliptic Jacobi functions. Similar difficulties also occur in
many other well-known problems of this type: the sub-Riemannian problem both
on the group of motions of the plane [7], [8] and on the Engel group [9], the gen-
eralized Didona problem [10], [11], etc. It was shown in [12] that, by contrast with
the aforementioned problems, extremal trajectories in the problem of the optimal
rolling of a sphere with twisting, but without slipping, are described by elementary
functions. In particular, in this case the point of contact between the sphere and the
plane is moving over a sinusoid or a straight line. Due to a simple parametrization
of the geodesics, many properties can be given an intuitive mechanical or geometric
interpretation, and therefore the problem is of special interest from the viewpoint
of mathematical control theory and sub-Riemannian geometry. An example of
a mechanical system with twisting is a billiard ball. In the future the solution of
this problem will be able to be applied to study similar systems.
In this paper we obtain a complete parametrization of the extremal trajectories
and begin to study their optimality. We analyse continuous and discrete symmetries
of the Hamiltonian system of the Pontryagin maximum principle and the corres-
ponding Maxwell points, that is, the points into which several extremal trajectories
arrive at the same moment of time with the same value of the functional. As is
well known, beyond such a point an extremal trajectory is not optimal [11], and
therefore the corresponding time gives an upper estimate for the cut time. We con-
sider the subproblem of returning the sphere to its initial point with a pre-assigned
orientation in more detail. In this case we prove that the exponential map is
a two-sheeted covering of certain domains in a fibre of the cotangent bundle into
the corresponding domains in SO(3). We consider a metric on the group of rotations
of three-dimensional space related to the return problem and its basic properties.

§ 2. The statement of the problem


In three-dimensional space R3 we choose a fixed right orthonormal frame
(e1 , e2 , e3 ) such that the plane on which the sphere is rolling is spanned by (e1 , e2 )
and e3 is directed into the upper half-space. We also choose a moving right orthonor-
mal frame (e′1 , e′2 , e′3 ) fixed at the centre of the sphere. Then the orientation of the
sphere is given by the rotation matrix SO(3) ∋ R : (e′1 , e′2 , e′3 ) 7→ (e1 , e2 , e3 ), and its
position by the coordinates of the centre (x, y) ∈ R2 in the basis (e1 , e2 ). As control
parameters we take the components of the angular velocity vector of the sphere in


the fixed frame Ω = (u2 , −u1 , u3 ) ∈ R3 . Then the kinematics of the system is
The optimal rolling of a sphere 159

defined by the equations


 
0 −u3 −u1
ẋ = u1 , ẏ = u2 , Ṙ = R u3 0 −u2  , (2.1)
u1 u2 0
(x, y) ∈ R2 , R ∈ SO(3), (u1 , u2 , u3 ) ∈ R3 .

As the minimized functional J[u] we choose the quadratic functional

1 t1 2
Z
J[u] = (u1 + u22 + u23 ) dt, (2.2)
2 0
which, up to a constant factor, is the integral of the rotational energy of the sphere.
It is required to roll the sphere from an initial state Q0 to a terminal state Q1 in
such a way that the minimum of the functional J[u] is attained.
This problem can be stated in a natural way as a sub-Riemannian left-invariant
problem on the Lie group G = R2 × SO(3). We recall from [13] that a sub-
Riemannian manifold (M, ∆, g) is by definition a smooth manifold M with a distri-
bution ∆ on which a Riemannian metric g is defined. A Lipschitzian curve is said
to be admissible if it is tangent to the distribution ∆ at almost all of its points. The
sub-Riemannian problem is stated as follows: on the sub-Riemannian manifold it is
required to find a shortest (in the sense of the metric g) admissible curve connecting
two given points Q0 and Q1 .
The group G can be represented as a subgroup of GL(6) using the matrices
 

 R 0 

 
Q= .

 1 0 x 

 0 0 1 y 
0 0 1

We introduce the left-invariant frame on G


∂ ∂
e1 = , e2 = , Vi (R) = RA
ei , i = 1, 2, 3,
∂x ∂y

where A
ei is the following basis of the Lie algebra so(3):
     
0 0 0 0 0 1 0 −1 0
Ae1 = 0 0 −1 , Ae2 =  0 0 0 , e3 = 1
A 0 0 .
0 1 0 −1 0 0 0 0 0

Then equations (2.1) define a control system on the group G and can be written in
the form

Q̇ = u1 X1 (Q) + u2 X2 (Q) + u3 X3 (Q), Q ∈ G, (u1 , u2 , u3 ) ∈ R3 , (2.3)

where
X1 = e1 − V2 , X2 = e2 + V1 , X3 = V3 .
160 I. Yu. Beschastnyı̌

The vector fields Xi define the distribution ∆ = span{X1 , X2 , X3 } ⊂ T G. If u(t) is


a measurable locally bounded map, then a solution of system (2.3) is an admissible
curve.
We can define an inner product ⟨ · , · ⟩ on the distribution ∆ as follows:
1
⟨QA, QB⟩Q = ⟨A, B⟩Id = − tr(AB), Q ∈ G, A, B ∈ ∆,
2
where tr A is the trace of the matrix A. Then the length of an admissible curve
Q(t), t ∈ [0, t1 ], is expressed in standard fashion
Z t1 q Z t1 q
l(Q) = ⟨Q̇, Q̇⟩ dt = u21 + u22 + u23 dt. (2.4)
0 0

Using the Cauchy-Bunyakovskiı̌ inequality it is easy to show that the length


functional (2.4) and its minimization are equivalent to the action (2.2) and its min-
imization. This method is often used in the study of extremal curves in Riemannian
and sub-Riemannian geometry [14].
Since the distribution ∆ and the metric g are left-invariant, left translations
can be used to shift Q0 to the identity element, so we can assume without loss of
generality that Q0 = (0, 0, Id).
Thus, we obtain the following left-invariant sub-Riemannian optimal control
problem on the group G = R2 × SO(3):
Q̇ = u1 X1 (Q) + u2 X2 (Q) + u3 X3 (Q), (2.5)
2 3
Q = (x, y, R) ∈ G = R × SO(3), (u1 , u2 , u3 ) ∈ R , (2.6)
Q(0) = Q0 = (0, 0, Id), Q(t1 ) = Q1 = (x1 , y1 , R1 ), (2.7)
Z t1 q
l(Q) = u21 + u22 + u23 dt → min. (2.8)
0

The value of the minimizing length functional (2.8) is independent of the para-
metrization of the curve Q(t); therefore we can assume that it has constant velocity,
that is, u21 + u22 + u23 ≡ const. Moreover, it is clear from (2.5) and (2.2) that if
a control u(t), t ∈ [0, t1 ], takes the sphere from the state Q0 to Q1 in time t1 , then
the control u′ (t) = ku(kt), where k is some positive number, takes Q0 to Q1 in
time t1 /k. Then Q(t) goes to Q(kt). This means we can assume without loss of
generality that u21 + u22 + u23 ≡ 1.
Since the multiplication table in the Lie algebra L = span{e1 , e2 , V1 , V2 , V3 } has
the form
ad ei = 0, [V1 , V2 ] = V3 , [V2 , V3 ] = V1 , [V3 , V1 ] = V2 ,
for the vector fields X1 , X2 , X3 we have
[X1 , X2 ] = X3 , [X1 , X3 ] = −V1 , [X2 , X3 ] = −V2 .
Then it is evident that span{X1 , X2 , X3 , [X1 , X3 ], [X2 , X3 ]} = L, and it follows from
the Rashevskiı̌-Chow theorem [13] that the system is completely controllable, that
is, any two points Q0 , Q1 can be connected by an admissible curve. The existence
of optimal trajectories in problem (2.5)–(2.8) follows from Filippov’s theorem [13].
The optimal rolling of a sphere 161

§ 3. Extremal trajectories
To find optimal trajectories for problem (2.5)–(2.7), (2.2) we apply Pontryagin’s
maximum principle (PMP) in an invariant formulation [13]. The Hamiltonian of
the maximum principle is the family of functions on the cotangent bundle T ∗ G
ν
hνu = λ, u1 X1 (Q) + u2 X2 (Q) + u3 X3 (Q) + (u21 + u22 + u23 ),

2
λ ∈ TQ∗ G, u = (u1 , u2 , u3 ) ∈ R3 , ν ∈ R.
We introduce the following basis Hamiltonians:
hi (λ) = ⟨λ, Xi (Q)⟩, i = 1, 2, 3,
h4 (λ) = −⟨λ, V1 (Q)⟩, h5 (λ) = −⟨λ, V2 (Q)⟩.
Since the basis Hamiltonians are functions that are linear on every fibre, they can
be regarded as coordinates in the cotangent space. The cotangent space has dimen-
sion five, and therefore, apart from the three Hamiltonians hi , i = 1, 2, 3, generated
by the vector fields Xi , it is necessary to introduce two more basis Hamiltonians,
h4 and h5 , generated by the two independent vector fields V1 and V2 .
The Hamiltonian of the PMP takes the form
ν
hνu = u1 h1 + u2 h2 + u3 h3 + (u21 + u22 + u23 ). (3.1)
2
Theorem 1 (Pontryagin maximum principle). Let u e(t) and Q(t),
e t ∈ [0, t1 ], be
an optimal control and the corresponding optimal trajectory in problem (2.5)–(2.7),

(2.2). Then a nontrivial Lipschitzian curve λt ∈ TQ(t) e G and a number ν ∈ R exist
such that
(ν, λt ) ̸= 0
and the following conditions hold:

− →
− →
− →

λ̇t = h νue(t) (λt ) = u
e1 (t) h 1 (λt ) + u
e2 (t) h 2 (λt ) + u
e3 (t) h 3 (λt ), (3.2)
hνue(t) = max hνu (λt ) for almost all t ∈ [0, t1 ], (3.3)
u∈R3
ν 6 0,


where h νue(t) is the Hamiltonian vector field on the cotangent bundle T ∗ G corres-
ponding to the Hamiltonian hνue(t) .
The curve λt is called an extremal, and its projection onto the base Q(t) =
π(λ(t)) an extremal trajectory. In the coordinates hi , the Hamiltonian system (3.2)
is written in the form
Q̇ = u1 X1 (Q) + u2 X2 (Q) + u3 X3 (Q),
ḣ1 = −u2 h3 − u3 h4 ,
ḣ2 = u1 h3 − u3 h5 ,
(3.4)
ḣ3 = u1 h4 + u2 h5 ,
ḣ4 = −u1 h3 + u3 h5 ,
ḣ5 = −u2 h3 − u3 h4 .
162 I. Yu. Beschastnyı̌

In the calculations which follow we use an isomorphism between R3 and the




Lie algebra so(3). Namely, with every vector A ∈ R3 we can associate a matrix
e ∈ so(3) by the following rule:
A
   
a1 0 −a3 a2


A = a2  , e =  a3
A 0 −a1  .
a3 −a2 a1 0
This isomorphism has the following properties [3].
1. Let A,
e Be ∈ so(3) correspond to → − → − e∼
A , B ∈ R3 (in this case we write A


= A,
e ∼ →
− →
− →

e and A × B denote the matrix commutator and the
B = B ). Let [A,
e B]
vector cross-product in R3 , respectively. Then
e ∼ →
− → −
[A,
e B] = A × B. (3.5)
2. Let R ∈ SO(3). Then
e −1 ∼ →

RAR = RA. (3.6)
3.1. Abnormal extremal trajectories. We consider the case ν = 0. In this
case the Hamiltonian (3.1) has the form
h0u = u1 h1 + u2 h2 + u3 h3
and attains its maximum with respect to ui when hi ≡ 0, i = 1, 2, 3. Then the
vertical subsystem of the Hamiltonian system (3.4) (that is, the restriction of (3.4)
to a fibre of T ∗ G) takes the form
0 = −u3 h4 ,
0 = −u3 h5 ,
0 = u1 h4 + u2 h5 ,
ḣ4 = u3 h5 ,
ḣ5 = −u3 h4 .
Since all the hi cannot vanish in the abnormal case, we obtain that u3 ≡ 0. Con-
sequently, h4 = const, h5 = const, and therefore u1 = const and u2 = const as
well.
Substituting the controls into (2.1) and recalling that (u1 , u2 , u3 ) = (−Ω2 , Ω1 , Ω3 )
we obtain the parametrization of abnormal trajectories

x = −Ω2 t, y = Ω1 t, R = eΩt , (3.7)


e

where Ω
e is the skew-symmetric matrix corresponding to the angular velocity vector


Ω with component Ω3 = 0:
 
0 0 Ω2
e = 0
Ω 0 −Ω1  .
−Ω2 Ω1 0
Consequently, in the abnormal case the angular velocity vector is a constant hor-
izontal vector and the sphere is rolling uniformly along a straight line without
twisting.
The optimal rolling of a sphere 163

Proposition 1. 1) All abnormal extremal trajectories of constant velocity have the


form (3.7).
2) Any abnormal extremal trajectory Q(t), t ∈ [0, t1 ], is optimal for any t1 > 0.

Proof. The proof of part 1) was given before the statement of this proposition.
2) Let Q(t), t ∈ [0, t1 ], be an abnormal extremal trajectory and let t1 > 0. We
now prove that Q(t) is optimal.
By part 1) abnormal curves of constant velocity satisfy the conditions

u1 = ẋ = const, u2 = ẏ = const, u3 ≡ 0.

The minimum of the functionals


Z t1 Z t1
J1 = (u21 + u22 ) dt, J2 = u23 dt
0 0

is attained on these curves, and so the minimum of the functional


Z t1
J[u] = (u21 + u22 + u23 ) dt = J1 + J2
0

is too. The proposition is proved.

3.2. Normal trajectories. We now consider the case ν = −1. The Hamiltonian
(3.1) is equal to

1
hu = u1 h1 + u2 h2 + u3 h3 − (u21 + u22 + u23 )
2

and it attains its maximum with respect to u for ui = hi , i = 1, 2, 3; as its maximum


the Hamiltonian is equal to

h21 + h22 + h23


H = max3 hu = = const.
u∈R 2

The vertical subsystem of the Hamiltonian system (3.4) for normal extremals
has the form
ḣ1 = −h3 (h2 + h4 ),
ḣ2 = h3 (h1 − h5 ),
ḣ3 = h1 h4 + h2 h5 , (3.8)
ḣ4 = −h3 (h1 − h5 ),
ḣ5 = −h3 (h2 + h4 ).

This system has two obvious first integrals, ω1 = −(h2 + h4 ) and ω2 = h1 − h5 .


Consequently, solving this system can be reduced to solving a system of the third
order. Using the fact that (u1 , u2 , u3 ) = (−Ω2 , Ω1 , Ω3 ) we rewrite the Hamiltonian
164 I. Yu. Beschastnyı̌

system of the maximum principle (3.4) in the following form:


ẋ = −Ω2 , (3.9)
ẏ = Ω1 , (3.10)
Ṙ = RΩ,
e (3.11)
Ω̇1 = ω2 Ω3 , (3.12)
Ω̇2 = −ω1 Ω3 , (3.13)
Ω̇3 = ω1 Ω2 − ω2 Ω1 , (3.14)
ω̇1 = 0, ω̇2 = 0.
In accordance with (3.5), we can write the system (3.12)–(3.14) concisely in one
of two equivalent ways:

−̇ →
− e˙ = [e
Ω =→

ω ×Ω ⇐⇒ Ω ω , Ω],
e (3.15)
where →−
ω = (ω1 , ω2 , 0) is a constant horizontal vector and ω
e is the corresponding


skew-symmetric matrix. Equations (3.15) show that the angular velocity vector Ω


rotates uniformly about some constant horizontal vector ω in the moving frame.
The modulus of the angular velocity of the sphere is equal to
v
u 3

− uX √
|Ω| = t u2i = 2H = const.
i=1

Consequently, the solutions of (3.15) are circles on a sphere of radius 2H
contained in planes perpendicular to the vector →−
ω (Fig. 1). Since the same extremal
trajectory corresponds to different values of H, but with different parametrizations,
we can assume without loss of generality that H = 1/2.

Figure 1. The general solution of system (3.15).


− →
− →

We set ω = |→−
ω | and Ω 0 = Ω (0) = (Ω01 , Ω02 , Ω03 ). If Ω 0 = λ→

ω for some λ ∈ R or

−̇ →

ω = 0, then we see that Ω = 0, and, consequently, Ω is a constant vector. In this
The optimal rolling of a sphere 165

case the equations for the extremal trajectories have the same form as (3.7); the
only difference is that Ω3 is not necessarily equal to zero, and therefore the sphere
is rolling along straight lines with or without twisting. Therefore the abnormal
extremal trajectories are normal, that is, they are not strictly abnormal.
We integrate equations (3.15) for ω ̸= 0. We obtain
 2
ω1 + ω22 cos ωt ω1 ω2 ω2

(1 − cos ωt) sin ωt   
   ω2 ω2 ω  Ω01
Ω1 

Ω2  =  ω1 ω2 2 2   0
ω2 + ω1 cos ωt ω1  Ω2  . (3.16)

2
(1 − cos ωt) 2
− sin ωt
Ω3  ω ω ω  Ω03
 
ω2 ω1
− sin ωt sin ωt cos ωt
ω ω
Using equations (3.16) we integrate equations (2.1) for x and y:
Ω03 ω1 (Ω01 ω2 − Ω02 ω1 )ω1 ω2 (Ω02 ω2 + Ω01 ω1 )
x= (1 − cos ωt) + sin ωt − t,
ω2 ω3 ω2 (3.17)
Ω0 ω2 (Ω0 ω2 − Ω02 ω1 )ω2 ω1 (Ω02 ω2 + Ω01 ω1 )
y = 3 2 (1 − cos ωt) + 1 3
sin ωt + t.
ω ω ω2
We will now obtain an expression for the matrix R. To do this we consider
equation (3.15). It is easy to see that its solution is given by

Ω(t)
e e 0 e−teω .
= eteω Ω

Substituting Ω
e into (2.1) we obtain

e 0 e−teω .
Ṙ(t) = R(t)eteω Ω
We make the change of variables
Z(t) = R(t)eteω .
Differentiating Z(t) with respect to t we obtain

Ż(t) = Ṙ(t)eteω + R(t)eteω ω e 0 e−teω eteω + R(t)eteω ω


e = R(t)eteω Ω e = R(t)eteω (Ω
e0 + ω
e ).
As a result we have the differential equation
Ż(t) = Z(t)(e
ω+Ω
e 0 ).

As is well known, its solution is the matrix exponential

Z(t) = Z(0)et(eω+Ω0 ) .
e

Substituting in the original variables and taking the fact that Z(0) = Id into account
we obtain the expression for the matrix R

R(t) = et(eω+Ω0 ) e−teω , (3.18)


e

which shows that the motion of the sphere consists of the composition of two uni-
form rotations: about the vectors →

ω and →− ⃗
ω + Ω.
Consequently, we have proved the following proposition.
166 I. Yu. Beschastnyı̌



Proposition 2. If ω ̸= 0 and Ω ̸= λ→ −ω , then the normal extremal trajectories
parametrized by arc length are described by the equations
Ω03 ω1 (Ω0 ω2 − Ω02 ω1 )ω1 ω2 (Ω02 ω2 + Ω01 ω1 )
x= 2
(1 − cos ωt) + 1 3
sin ωt − t,
ω ω ω2 (3.19)
Ω0 ω2 (Ω0 ω2 − Ω02 ω1 )ω2 ω1 (Ω02 ω2 + Ω01 ω1 )
y = 3 2 (1 − cos ωt) + 1 sin ωt + t,
ω ω3 ω2
R(t) = et(eω+Ω0 ) e−teω . (3.20)
e

In the remaining cases the normal extremal trajectories are described by the equa-
tions
x = −Ω2 t, y = Ω1 t, R = eΩt , (3.21)
e

where Ωe is the skew-symmetric matrix corresponding to an arbitrary unit vector →



Ω.
A parametrization of normal extremal trajectories was also obtained in [12]. It
can be derived from equations (3.19) and (3.20) as a special case with Ω03 = 0.
3.3. Rotations as symmetries of the Hamiltonian system. To simplify the
calculations that follows, we show that using rotations in the plane (x, y) we can
choose the vector →
−ω to be (ω, 0, 0). In this case, the complicated equations for the
normal geodesics (3.19) are transformed to the simple form
x = −ω −1 (Ω02 sin ωt + Ω03 cos ωt − Ω03 ), y = Ω01 t. (3.22)
Proposition 3. Rotation about the axis OΩ3 in the space (Ω1 , Ω2 , Ω3 ) generates
a one-parameter symmetry group {Φβ , β ∈ S 1 } of the whole system (3.9)–(3.14),
which is defined as follows:
Φβ : {Ω es , Qs : s ∈ [0, t]} 7→ {Ω
e s, ω e βs , ω
esβ , Qβs : s ∈ [0, t]},
e β = eβA3 Ω
Ω e s e−βA3 , e β = eβA3 ω
ω e e−βA3 , (3.23)
s
 β   
xs cos β − sin β xs
= , (3.24)
ysβ sin β cos β ys
Rsβ = eβA3 Rs e−βA3 . (3.25)
Proof. We now prove that equations (3.12)–(3.14) are invariant under the one-
parameter group Φβ :
β
e˙ = [e
Ω ωsβ , Ω
eβ] ⇐⇒ e˙ s e−βA3
eβA3 Ω
s s

= eβA3 ω es e−βA3 eβA3 Ω


e s e−βA3 − eβA3 Ωe s e−βA3 eβA3 ω
es e−βA3
⇐⇒ eβA3 Ωe˙ s e−βA3 = eβA3 (eωs Ωes − Ω es )e−βA3
e sω

⇐⇒ e˙ s = [e
Ω ωs , Ω
e s ].

That equations (3.11) are invariant can be proved in a similar fashion:

Ṙsβ = Rsβ Ω

s ⇐⇒ eβA3 Ṙs e−βA3 = eβA3 Rs e−βA3 eβA3 Ω
e s e−βA3
⇐⇒ eβA3 Ṙs e−βA3 = eβA3 (Rs Ω
e s )e−βA3 ⇐⇒ Ṙs = Rs Ω
e s.
The optimal rolling of a sphere 167

In the case of equations (3.9), (3.10) we use equation (3.6). It follows from (3.23)
that

e s e−βA3
e βs = eβA3 Ω →
−β →

Ω ⇐⇒ Ω s = eβA3 Ω s
 β   
Ω1s cos β − sin β 0 Ω1s
⇐⇒ Ωβ  =  sin β cos β 0 Ω2s  .
2s
Ωβ3s 0 0 1 Ω3s

We substitute the expressions for Ωβ1s , Ωβ2s into (3.9), (3.10):

ẋβs −Ωβ2s
         
cos β sin β ẋs cos β sin β −Ω2s
= ⇐⇒ =
ẏsβ Ωβ1s − sin β cos β ẏs − sin β cos β Ω1s
   
ẋs −Ω2s
⇐⇒ = .
ẏs Ω1s

The proposition is proved.


It is clear from formulae (3.23) that under the action of the symmetry Φβ the


vectors →−
ω and Ω are rotated through the angle β in the horizontal plane. Obviously,
there exists a value of β such that the horizontal vector →
−ω can be reduced to the

− β
form ω = (ω, 0, 0). In this case we arrive at equations (3.22), which show that
the sphere is rolling along an arc of a sinusoid whose axis is perpendicular to the
vector →−
ω.
Proposition 4. The optimality of the normal extremal trajectories (3.19), (3.20)
is preserved under the symmetry Φβ defined by the system (3.23)–(3.25).
Proof. Consider the family Γ of all normal geodesics connecting Q0 and Q1 . Under
the action of the symmetry Φβ , the point Q0 remains fixed, and the point Q1 goes
to the point Qβ1 = Φβ (Q1 ). Consequently, Φβ is a bijective map from the family
Γ into the family Γβ of all normal geodesics connecting the points Q0 and Qβ1 .
The symmetry Φβ leaves the length functional (2.8) invariant; therefore if γt is
a shortest trajectory of the family Γ, then γtβ = Φβ (γt ) is a shortest trajectory of
the family Γβ . The proposition is proved.
In view of Proposition 4, in what follows we can confine ourselves to studying
the optimality of trajectories for which → −
ω = (ω, 0, 0).
At the end of this subsection we give a summary of the results we have obtained
and a description of all the extremal trajectories (Fig. 2). Number 1 in Fig. 2
denotes the most general case when the sphere is rolling along a sinusoid. In this


motion, by Proposition 2 we have → −
ω ̸= 0, Ω ̸= λω, and the trajectory is described


by equations (3.19), (3.20). If Ω ⊥ → −
ω , then the sinusoid degenerates into a segment
of a straight line passing through the point (0, 0) (case 2). For → −ω = (ω, 0, 0) the

− →
− 0
condition Ω ⊥ ω means that Ω1 ≡ 0.
When ω = 0 an extremal trajectory is described by equations (3.21). Its proj-
ection onto the plane is a straight line going out from the origin (case 3), or a point
(the sphere stays on the spot and twists).
168 I. Yu. Beschastnyı̌

Figure 2. Projections of extremal trajectories onto the plane (x, y).

3.4. A representation of the orientation of the sphere by a quaternion


of unit length. In the calculations which follow, rather than a rotation matrix R
it is more convenient to use the corresponding quaternion q. In this subsection we
rewrite equations (3.20) in quaternion form.
Let H = {q = q0 + iq1 + jq2 + kq3 : q0 , . . . , q3 ∈ R} be the
p algebra of quaternions.
The length of q ∈ H is defined by the formula |q| = q02 + q12 + q22 + q32 . The
quaternion q = q0 − iq1 − jq2 − kq3 is called the conjugate quaternion of q. The
element inverse to q is given by the following formula:

q
q −1 = .
|q|2

Consider the unit sphere S 3 = {q ∈ H : |q| = 1} and the space of imaginary


quaternions I = {q ∈ H : q0 = 0}, which is naturally identified with the Euclidean
space R3 . Any quaternion q ∈ S 3 defines a rotation of a vector a ∈ I by the
operator Rq :

Rq : a 7→ qaq −1 ∈ I.

To every rotation operator Rq there correspond two different quaternions q and −q,
and therefore the map p : q 7→ Rq defines a two-sheeted covering S 3 over SO(3).
The covering map is a homomorphism [15].
The matrix R of a rotation about an arbitrary nonzero vector → −a ∈ R3 , →
−a =
(a1 , a2 , a3 ), through an angle β has a corresponding unit quaternion

β a1 i + a2 j + a3 k β
q = cos + sin . (3.26)
2 |→

a| 2
The optimal rolling of a sphere 169

In particular, since →

ω = (ω, 0, 0), the quaternions

− →

|→

ω + Ω 0 |t (ω + Ω01 )i + Ω02 j + Ω03 k |→

ω + Ω 0 |t
cos + →
− sin ,
2 |→
−ω + Ω 0| 2
ωt ωt
cos − i sin
2 2

correspond to the matrices et(eω+Ω0 ) and e−teω involved in (3.20). Substituting the
e

expressions we have obtained into (3.20) we obtain a law for changing the compo-
nents of the quaternion q corresponding to the matrix R:

− →

|Ω0 + →
−ω |t ωt Ω01 + ω |Ω0 + → −ω |t ωt
q0 = cos cos + → − sin sin ,
2 2 →

|Ω0 + ω | 2 2

− →

Ω01 + ω tω |Ω0 + →−ω |t |Ω0 + → −ω |t ωt
q1 = → − cos sin − cos sin ,


|Ω0 + ω | 2 2 2 2

− (3.27)
1

ωt ωt
 →

| Ω 0 + ω |t
q2 = → − Ω02 cos − Ω03 sin sin ,


|Ω0 + ω | 2 2 2


|Ω0 + → −
 
1 0 ωt 0 ωt ω |t
q3 = → − Ω cos + Ω sin sin .


|Ω0 + ω |
3
2 2
2 2

Of the two quaternions ±q corresponding to the matrix R, here we have given


the quaternion with the initial conditions q(0) = 1.

§ 4. Symmetries of the exponential map


By definition, a symmetry of a map f : M → N is a pair of maps g : M → M
and g ′ : N → N such that the following diagram is commutative:

M
f
/N
g g′
 
M
f
/ N.

In what follows we shall denote the pair of maps g and g ′ by the one symbol g.
The exponential map is defined to be a map Exp(λ, t) taking a covector in the
cotangent space λ ∈ {λ ∈ TQ∗ 0 G : H(λ) = 1/2} and a moment of time t ∈ R+ to the
end-point of the corresponding normal extremal trajectory. In the problem under
consideration, this is the map

Exp : S 2 × R2 × R+ → R2 × SO(3),

or

− −
Exp : ( Ω , →
ω , t) 7→ (x, y, R)
which is defined by equations of the form (3.7) if the projection of the trajectory onto
the plane (x, y) is a straight line, and (3.22), (3.20) if the projection is a sinusoid.
170 I. Yu. Beschastnyı̌

In this section we consider the symmetries of the exponential map and the Hamil-


tonian system (3.9)–(3.14) that preserve the condition | Ω | = const. As shown ear-
lier, the Hamiltonian system has one continuous symmetry Φβ . Later we shall show
that the exponential map also has seven discrete symmetries εi , i = 1, . . . , 7. For
brevity, we use the notation N = S 2 × R2 × R+ in what follows.
4.1. Discrete symmetries in the pre-image of the exponential map.
Proposition 5. Equations (3.12)–(3.14) have the following symmetries:

ε1 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (Ω1 (t − s), −Ω2 (t − s), Ω3 (t − s)),


ε2 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (Ω1 (t − s), Ω2 (t − s), −Ω3 (t − s)),
ε3 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (Ω1 (s), −Ω2 (s), −Ω3 (s)),
ε4 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (−Ω1 (s), Ω2 (s), Ω3 (s)), (4.1)
ε5 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (−Ω1 (t − s), −Ω2 (t − s), Ω3 (t − s)),
ε6 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (−Ω1 (t − s), Ω2 (t − s), −Ω3 (t − s)),
ε7 : (Ω1 (s), Ω2 (s), Ω3 (s)) 7→ (−Ω1 (s), −Ω2 (s), −Ω3 (s)),

and →

ωi = →

ω.
To prove this we only need to substitute (4.1) into (3.12)–(3.14).
The transformations (4.1) can also be written in the matrix form

ε1 : Ω
es →
7 −I2 Ω
e t−s I2 ,

ε2 : Ω
es →
7 −I3 Ω
e t−s I3 ,

ε3 : Ω
es →
7 I1 Ω
e s I1 ,

ε4 : Ω
es →
7 −I1 Ω
e s I1 , (4.2)

ε5 : Ω
es →
7 I3 Ω
e t−s I3 ,

ε6 : Ω
es →
7 I2 Ω
e t−s I2 ,

ε7 : Ω
es →
7 −Ω
e s,

where
   
1 0 0 −1 0 0
I1 = 0 −1 0  = eπA1 ∼
= i, I2 =  0 1 0  = eπA2 ∼
= j,
0 0 −1 0 0 −1
 
−1 0 0
I3 =  0 −1 0 = eπA3 ∼
= k.
0 0 1

The discrete symmetries commute and form a group with the following multipli-
cation table:
The optimal rolling of a sphere 171

ε1 ε2 ε3 ε4 ε5 ε6 ε7
ε1 Id ε3 ε2 ε5 ε4 ε7 ε6
2 1 6 7 4
ε Id ε ε ε ε ε5
ε3 Id ε7 ε6 ε5 ε4
ε4 Id ε1 ε2 ε3
ε5 Id ε3 ε2
6
ε Id ε1
ε7 Id

The group {Id, ε1 , . . . , ε7 } ∼


= Z2 × Z2 × Z2 is the symmetry group of a paral-
lelepiped.
The action of the symmetries εi is shown in Fig. 3. It is clear from (4.1) that the
symmetries ε1 –ε3 preserve the orbits of the system (3.12)–(3.14) (Fig. 3, (a)). The
symmetry ε4 is reflection with respect to the plane passing through the origin, per-
pendicular to the vector → −ω (Fig. 3, (b)), and the other symmetries are compositions
of these (Fig. 3, (c)).

Figure 3. Action of the discrete symmetries in the pre-image of the exponen-


tial map: (a) action of the symmetries ε1 –ε3 ; (b) action of the symmetry ε4 ;
(c) action of the symmetries ε4 –ε7 .

4.2. Discrete symmetries in the image of the exponential map.


Proposition 6. The discrete transformations εi generate the following symmetries
of extremal trajectories:
 
1 2
xs = xt−s − xt ,
 xs = xt − xt−s ,

1 1 2 2
ε : ys = yt − yt−s , ε : ys = yt − yt−s ,
−1
Rs = I3 (Rt )−1 Rt−s I3 ,

 1 
 2
Rs = I2 (Rt ) Rt−s I2 ,
 
3 4
xs = −xs ,
 xs = xs ,

ε3 : ys3 = ys , ε4 : ys4 = −ys ,
Rs = I1 es(eω−Ω0 ) e−seω I1 ,

 3 
 4
Rs = I1 Rs I1 ,
e
172 I. Yu. Beschastnyı̌

 

 x5s = xt−s − xt , 
x6s = xt − xt−s ,
 
5 6
ys = yt−s − yt , ys = yt−s − yt ,

 

 
5 6
ε : Rs5 = I3 eteω e−t(eω−Ω0 ) ε : Rs6 = I2 eteω e−t(eω−Ω0 )
e e
 
×e(t−s)(eω−Ω0 ) ×e(t−s)(eω−Ω0 )

 e 
 e

 

×e−(t−s)eω I3 , ×e−(t−s)eω I2 ,

 


7
xs = −xs ,

7
ε : ys7 = −ys ,
Rs = es(eω−Ω0 ) e−seω .

 7 e

Proof. Since we use the same method in all cases, we will only prove the assertion
for the symmetry ε1 .
Equations (3.9), (3.10) imply

d
ẋ1s = (xt−s − xt ) = −ẋt−s = Ω2 (t − s) = −Ω12 (s), (4.3)
ds
d
ẏs1 = (yt − yt−s ) = ẏt−s = Ω1 (t − s) = Ω11 (s). (4.4)
ds
We now verify that equation (3.11) is invariant under ε1 :

d
Ṙs1 = (I2 (Rt )−1 Rt−s I2 ) = −I2 (Rt )−1 Ṙt−s I2 = −I2 (Rt )−1 Rt−s Ωt−s I2
ds
= I2 (Rt )−1 Rt−s I2 (−I2 Ωt−s I2 ) = Rs1 Ω1s .

The proposition is proved.

Figure 4. Action of the symmetries ε1 –ε7 on the projections of extremal


trajectories onto the horizontal plane.

The symmetry ε2 is the reflection of the arc of a sinusoid with respect to the
centre of the chord connecting the beginning and end-point of a trajectory. The
The optimal rolling of a sphere 173

symmetries ε3 , ε4 are the reflections of the arc with respect to the coordinate axes
OY and OX, respectively. All the other symmetries are compositions of these
reflections (Fig. 4).
Explicit expressions for the Rsi in terms of Rs have been obtained only for the
first three, ε1 –ε3 . Therefore in what follows we confine ourselves to studying only
these three reflections.
4.3. Symmetries of the exponential map. A rotation
Φβ : λt 7→ λβt
(see (3.23)–(3.25)) is a symmetry of the Hamiltonian system; therefore its action in
T ∗ G decomposes naturally into a direct sum of actions in N and G:
Φβ : N → N, (λ, t) 7→ (λβ , t),

− − →
− −β →
−β →

λ = (Ω,→
ω ), λβ = ( Ω β , →
ω ), Ω = eβA3 Ω , →

ω β = eβA3 →

ω,
where λ = λt |t=0 , and
Φβ : G → G, Q 7→ Qβ ,
Q = (x, y, R), Qβ = (xβ , y β , Rβ ),
β
    
x cos β − sin β x
= , Rβ = eβA3 Re−βA3 .
yβ sin β cos β y

The action of the reflections εi , i = 1, 2, 3, in N is determined by the restriction


of their action to the vertical components of extremals at the initial moment of
time s = 0:
ε1 : ( Ω e 10 , ω
e , t) 7→ (Ω
e 0, ω e 1 , t) = (−I2 Ω
e t I2 , ω
e , t),
2 2 2
ε : (Ω e , t) 7→ (Ω
e 0, ω e 0, ω e v) = (−I3 Ω e t I3 , ω
e , t),
ε3 : ( Ω e 30 , ω
e , t) 7→ (Ω
e 0, ω e 3 , t) = (I1 Ω
e 0 I1 , ω
e , t).
The action of the reflections εi in G is determined by their action on extremal
trajectories at the terminal moment of time s = t:
ε1 : (xt , yt , Rt ) 7→ (x1t , yt1 , Rt1 ) = (−xt , yt , I2 (Rt )−1 I2 ),
ε2 : (xt , yt , Rt ) 7→ (x2t , yt2 , Rt2 ) = (xt , yt , I3 (Rt )−1 I3 ),
ε3 : (xt , yt , Rt ) 7→ (x3t , yt3 , Rt3 ) = (−xt , yt , I1 Rt I1 ).
We find that the action of rotations and reflections is defined in the pre-image
of the exponential map
Φβ , εi : N → N, (λ, t) 7→ (λβ , t), (λi , t),
Φβ , εi : G → G, Qt 7→ Qβt , Qit .

Proposition 7. The maps Φβ and εi are symmetries of the exponential map.


Proof. The proof follows directly from the definitions of the actions Φβ , εi in N , G
and from equations (3.7), (3.22), (3.20).
174 I. Yu. Beschastnyı̌

§ 5. Maxwell points and times


A Maxwell point is defined as a point Q into which two different extremal trajec-
tories arrive simultaneously with the same value of the functional and time (Fig. 5).
It is well known that in an analytic sub-Riemannian problem a normal extremal tra-
jectory cannot be optimal beyond such a point [11]. The moment of time at which
an extremal trajectory arrives at a Maxwell point is called a Maxwell time tmax .

Figure 5. A Maxwell point.

The following proposition answers the question of when two extremal trajectories
are different in the problem under consideration.

Proposition 8. 1. Let Qs , Q′s be two extremal trajectories such that Ω3 ̸≡ 0 and



− →

Ω′3 ̸≡ 0. They coincide if and only if (→

ω , Ω 0 ) = (→

ω ′ , Ω ′0 ).
2. Let Qs , Q′s be two abnormal extremal trajectories. They coincide if and only

−′ →

if Ω 0 = Ω 0 .

Proof. We have the following chain of equivalent statements:



− →

Qs ≡ Q′s ⇐⇒ xs ≡ x′s , ys ≡ ys′ , Rs ≡ Rs′ ⇐⇒ Ω s ≡ Ω ′s . (5.1)

Suppose that Ω3 (s) ̸= Ω3 (0) for some s > 0, that is, the trajectory is not
abnormal. Then we have two cases.

−̇ →

1. →

ω = 0. Then it follows from (3.15) that Ω = 0. Therefore, Ω = const, and

−′
by the last assertion of the chain (5.1) we have Ω = const. Consequently,


ω ′ = 0.

− →

2. ω ̸= 0. In this case we obtain that the vector Ω s is rotating uniformly

− →
− →

about ω with velocity ω. But since we must have Ω s ≡ Ω ′s , we find that

−′
Ω s must also be rotating about → −
ω with the same angular velocity in the
same direction. Therefore, ω = →

− ′ −
ω , and the first part of the assertion is
proved.

− →

Suppose that Ω s ≡ Ω 0 is a constant horizontal vector, that is, the extremal


trajectory is abnormal. Then → −ω = λ Ω for an arbitrary λ; but, as follows from
(3.7), the trajectory is independent of ω. Therefore the last assertion of the chain

− →

(5.1) is sufficient and is equivalent to Ω 0 = Ω ′0 . The proposition is proved.
The optimal rolling of a sphere 175

In what follows we shall assume that q = q0 +q1 i+q2 j +q3 k ∈ S 3 , q(0) = 1, is the
quaternion corresponding to the orientation matrix R, and χ is the angle between
the chord connecting (0, 0) and (x, y) determined from the following equations for
(x, y) ̸= (0, 0):
p
x = ρ cos χ, y = ρ sin χ, ρ = x2 + y 2 .

5.1. Fixed points in the image of the exponential map and the Maxwell
time for a sphere rolling along a straight line. By Proposition 1, if the
angular velocity vector is horizontal for a sphere rolling along a straight line, then
the trajectory is optimal and there are no Maxwell points on it. If, however, the
angular velocity vector is not horizontal, then the following proposition holds.


Proposition 9. If the sphere is rolling along a straight line in such a way that Ω
is not horizontal, then t = 2π is a Maxwell time. If the sphere is twisting on the
spot, then t = π is a Maxwell time.
Proof. Let Qt be the trajectory described by equations (3.21) with Ω03 ̸= 0. If
Ω01 = 0 and Ω02 = 0, then the sphere stays on the spot and twists; otherwise it rolls
along a straight line.
Consider a trajectory Q′t such that → − ′ ′ ′
ω ′ = 0, Ω10 = Ω01 , Ω20 = Ω02 , Ω30 = −Ω03 . For
a moment of time t to be a Maxwell time, we must have Q′t = Qt . It follows from
(3.21) that the projection of Q′t onto the plane (x, y) coincides with the projection
of Qt , that is, xt = x′t and yt = yt′ for any t. Consequently, the proof reduces to
verifying that q(2π) = ±q ′ (2π) in the first case, and q(π) = ±q ′ (π) in the second,
e′
where q and q ′ are the quaternions corresponding to the matrices eΩ0 t and eΩ0 t ,
e

t t t t
q(t) = cos + (Ω01 i + Ω02 j + Ω03 k) sin , q ′ (t) = cos + (Ω01 i + Ω02 j − Ω03 k) sin .
2 2 2 2
The proposition is proved.
5.2. Fixed points in the image of the exponential map for the case of
rolling along a sinusoid. In what follows we assume that the sphere is rolling


along a sinusoid, that is, the conditions →

ω ̸= 0 and Ω 0 ̸= λ→−ω hold. Then the
exponential map is defined by equations (3.22) and (3.27). In what follows we use
the following lemma.
Lemma 1 (see [4]). Let q ∈ S 3 ⊂ H. Then
1. q 2 = 1 ⇐⇒ q = ±1 ⇐⇒ Im q = 0;
2. q 2 = −1 ⇐⇒ Re q = 0.
We now find the fixed points in the image of the exponential map under the
action of the symmetries Φβ ◦ εi .
Proposition 10. 1. If (x, y) ̸= (0, 0), then the condition Φβ ◦ ε1 (Q) = Q is equiv-
alent to the disjunction

xq1 + yq2 = 0 or q = ±(i cos χ + j sin χ).

2. If (x, y) = (0, 0), then Φβ ◦ ε1 (Q) = Q holds for β = 0.


176 I. Yu. Beschastnyı̌

Proof. The definition of the actions of Φβ and εi (§ 4.3) implies that

eβA3 I2 R−1 I2 e−βA3 = R ⇐⇒ I2 e−βA3 R−1 I2 e−βA3 = R ⇐⇒ (ReβA3 I2 )2 = Id


   2
β β
⇐⇒ q cos + sin k j = ±1
2 2
  2
β β
⇐⇒ q cos j − sin i = ±1.
2 2

It follows from Lemma 1 that


  2
β β
q cos j − sin i = ±1
2 2
   
β β 
β β
Re q cos j − sin i = 0, −q2 cos + q1 sin = 0,
 2 2 2 2 
⇐⇒  ⇐⇒  (5.2)
 
β β
 
β β
q = ± cos j − sin i .

q cos j − sin i = ±1
2 2 2 2

Suppose that (x, y) = (ρ cos χ, ρ sin χ) ̸= (0, 0). Then


    
cos β − sin β −x x
= ⇐⇒ β = 2χ + π.
sin β cos β y y

Consequently, we see that Φβ ◦ ε1 (Q) = Q is equivalent to the equations



q1 cos χ + q2 sin χ = 0,
q = ±(sin χi + cos χj).

Suppose that (x, y) = (0, 0); then at the first moment of time when the sphere
returns to the origin, q2 will be equal to zero. Indeed, we transform (3.22) to the
form
 
2 ωt ωt ωt
x(t) = − sin Ω02 cos − Ω03 sin . (5.3)
ω 2 2 2

The function x(t) vanishes when

ωt ωt ωt
sin = 0 or Ω02 cos − Ω03 sin = 0,
2 2 2

but, as is evident from (3.27), when the second equation holds, q2 is equal to zero.
In this case it follows from the first equation (5.2)

β β
−q2 cos + q1 sin = 0
2 2

that β = 2πk, k ∈ Z. The proposition is proved.


The optimal rolling of a sphere 177

Proposition 11. The condition Φβ ◦ ε2 (Q) = Q is equivalent to the disjunction

q3 = 0 or q = ±k.

Proof. By the definition of the action of the symmetry ε2 , this reflection takes the
endpoint of the sinusoid to itself, and therefore β = 0. Next, we have

−1 2 2 Re(qk) = 0,
I3 R I3 = R ⇐⇒ (RI3 ) = Id ⇐⇒ (qk) = ±1 ⇐⇒
qk = ±1

q3 = 0,
⇐⇒
q = ±k
The proposition is proved.
Proposition 12. 1. If (x, y) ̸= (0, 0), then the condition Φβ ◦ ε3 (Q) = Q is equiv-
alent to the disjunction
( (
q1 x + q2 y = 0, q1 y − q2 x = 0,
or
q3 = 0 q0 = 0.

2. If (x, y) = (0, 0), then Φβ ◦ ε3 (Q) = Q is equivalent to the disjunction

q3 = 0 or q0 = 0,

where β = 0 in the first case, and β = π in the second.


Proof. The definitions of the action of Φβ , εi imply that

eβA3 I2 RI1 e−βA3 = R ⇐⇒ RI1 e−βA3 = I1 e−βA3 R


   
β β β β
⇐⇒ q i cos + j sin = ± i cos + j sin q
2 2 2 2
   
β β β β
 1q sin − q2 cos k + q3 cos j − sin i = 0,
⇐⇒   2 2  2 2
 β β β β
q0 cos i + sin j − q1 cos − q2 sin = 0
2 2 2 2
 
β β β β
q1 sin − q2 cos = 0, −q1 cos − q2 sin = 0,
 
⇐⇒ 2 2 or 2 2 (5.4)
q = 0 q = 0.
3 0

Suppose that (x, y) ̸= (0, 0). Then similarly to the proof of Proposition 10 we
see that β = 2χ + π. Substituting this value into (5.4) we arrive at the equations
given in the statement of this proposition.
If (x, y) = (0, 0), then q2 = 0. Therefore (5.4) is transformed to the form
 
β β
q1 sin = 0, −q1 cos = 0,
 
2 or 2
q = 0 q = 0.
3 0

Hence it is clear that β = 0 in the first case, and β = π in the second. The
proposition is proved.
178 I. Yu. Beschastnyı̌

§ 6. Maxwell times for the case of a sphere rolling along a sinusoid


The Maxwell time corresponding to the symmetry ε2 is found from the equation
q3 = 0, that is,


|Ω0 + →

 
1 0 ωt 0 ωt ω |t

− Ω3 cos + Ω2 sin sin = 0.


|Ω0 + ω | 2 2 2

Therefore,


ωt ωt |Ω0 + →

ω |t
Ω03 cos + Ω02 sin = 0 or
sin = 0.
2 2 2


We denote the angle between the projection of Ω onto the plane Ω2 OΩ3 and the
axis OΩ2 by ψ,

Ω2 (t) Ω3 (t)
cos ψ(t) = p , sin ψ(t) = p .
1 − (Ω1 (t))2 1 − (Ω1 (t))2

An extremal trajectory has a Maxwell point corresponding to the symmetry


Φβ ◦ ε2 only at a moment of time satisfying the second equation. Now, the first
equation holds for t = 2(πk − ψ0 )/ω. Then ψ(t) = ψ0 + tω. But from (4.1) we
see that under the action of the symmetry ε2 the angle ψ0 is transformed into
ψ02 = −ψ(t) = ψ0 − 2πk = ψ0 . That is, the trajectory is transformed into itself
and, consequently, the moment of time t is not a Maxwell time.
The second equation defines the Maxwell time


t2max = →− ,
|Ω0 + →

ω|

which has the following property. If we substitute t2max into (3.20), then we obtain

− →

R = e(2π/| Ω 0 + ω |)eω ,


that is, this is the rotation about the vector →−ω through the angle 2π/| Ω 0 + → −
ω |.


Hence it is evident that if we fix ω , then all the trajectories for which the values


|Ω0 + → −
ω | coincide have the same component R at the moment of time t2max . But


different points Ω 0 taken on the same circle that is a solution of (3.15) correspond


to such trajectories, since when Ω 0 rotates about →−
ω , the angle between them does
not change.
We now find the solutions of the equation xq1 + yq2 = 0 which describe the fixed
points of the symmetry Φβ ◦ ε1 . Substituting x(t), y(t), q1 (t), and q2 (t) we obtain


Ω01 |Ω0 + →−
 
ωt ωt ω |t
Ω02 cos − Ω03 sin →
− sin
2 2 →

|Ω0 + ω | 2

− →

|Ω0 + →− |Ω0 + →

 0 
2 ωt Ω1 + ω ωt ω |t ω |t ωt
− sin cos sin − cos sin = 0.
ω 2 |→ −
Ω0 + →−
ω| 2 2 2 2
The optimal rolling of a sphere 179

Consider the equation corresponding to the first factor,


ωt ωt
Ω02 cos − Ω03 sin = 0. (6.1)
2 2
The first positive moment of time when (6.1) holds is described by the expression

π(2k − 1) − 2ψ0
tR = ,
ω
where k ∈ Z is chosen in such a way that the time tR is minimal for a given
trajectory. It is evident from (5.3) that then x(tR ) = 0. However, this moment of
time is not a Maxwell time, since the trajectory is transformed into itself under the
action of the corresponding symmetry.
Indeed, as in the case of the symmetry ε2 , from (4.1) we have

ψ01 = π(2l − 1) − ψ(tR ) = ψ0 + 2π(l − k) = ψ0 ,

where (
1 if ψ0 ∈ [0, π),
l=
2 if ψ0 ∈ [π, 2π).
It follows from Proposition 10 that the trajectory we have obtained should be
rotated through the angle β = π + 2χ, but since x(tR ) = 0, we have χ = π/2 and
β = 2π, and so we conclude that the trajectory remains fixed.
In spite of what we have said above, the time tR plays an important role in the
problem of re-orientation, which is discussed in the next section. In fact, this is
the moment of the first return of the sphere to the origin when it is rolling along
a segment of a straight line. In order not to be concerned with the choice of k in
the expression for tR , it is convenient to introduce a new variable
π
ψ′ = ψ + mod π.
2
In this case, tR is equal to
2(π − ψ0′ )
tR = . (6.2)
ω

§ 7. The problem of re-orienting the sphere


In this section we consider the subproblem of returning the sphere to the origin
with a given orientation. We have to find a control that takes the sphere from the
initial state Q0 = (0, 0, Id) to a terminal state Q1 = (0, 0, R) in such a way that the
action functional attains its minimum.
7.1. The restriction of the exponential map. As before, we consider the case


ω = (ω, 0, 0). We also assume that the sphere is not twisting on the spot. Then
by equations (3.22) the sphere can return to its starting point only if it is rolling
along a segment of a straight line, that is, when Ω1 ≡ 0.
It follows from equations (5.3) that the period of the sinusoid is equal to 2π/ω.
Therefore the sphere first crosses the straight line x = 0 at time tR , and then at
180 I. Yu. Beschastnyı̌

time t = 2π/ω. It is easy to verify that t2max < 2π/ω. This follows from the obvious
inequality
1 1
√ < . (7.1)
ω2 + 1 ω
Therefore a re-orientation can be optimal only if the sphere returns to the origin
at time tR .
Since the return time is fixed, the pre-image of the restricted exponential map is
a semicylinder R+ × S 1 defined by the values of (ω, ψ0′ ). As follows from equations
(3.27), the orientation quaternion q(tR ) takes the form
√ 2 
ω +1
q0 = − cos (π − ψ0′ ) cos ψ0′
ω
√ 2 
ω ω +1
+√ sin (π − ψ0 ) sin ψ0′ ,

ω2 + 1 ω
√ 2 
ω ω +1
q1 = − √ sin (π − ψ0′ ) cos ψ0′
ω2 + 1 ω (7.2)
√ 2 
ω +1
− cos (π − ψ0′ ) sin ψ0′ ,
ω
q2 = 0,
√ 2 
1 ω +1
q3 = √ sin (π − ψ0 ) sin(π + ψ − ψ0′ ).

ω2 + 1 ω

These formulae define a coordinate representation of the exponential map restricted


to R+ × S 1 , the image of which is the surface defined by the equations x = 0, y = 0,
q2 = 0, |q| = 1. Consequently, the exponential map is reduced to the map

Exp′ : R+ × S 1 → S 2 ,
Exp′ (ω, ψ0′ ) = q(tR ). (7.3)

We further exclude from the semicylinder domains to which obviously non-


optimal trajectories correspond. Since at time tR the sphere returns to the ori-
gin, in order for the trajectory to be optimal it is necessary that the following
condition holds:
tR 6 t2max .

Proposition 13. 1) Suppose that ψ0′ ̸= 0 and let

π − ψ0′
ω=p ′ .
ψ0 (2π − ψ0′ )

The following assertions are equivalent:


• tR < t2max ⇐⇒ ω > ω;
• tR = t2max ⇐⇒ ω = ω;
• tR > t2max ⇐⇒ ω < ω.
2) Suppose that ψ0′ = 0. Then tR > t2max .
The optimal rolling of a sphere 181

Proof. Part 2) follows from inequality (7.1). The proof of part 1) is based on
comparing the quantities

2(π − ψ0′ ) 2π
and √ ,
ω 1 + ω2
which reduces to solving quadratic inequalities of the type
 ′ ′

2 ψ0 (2π − ψ0 )
ω − 1 > 0.
(π − ψ0′ )2

Since a vector cannot have negative length, we obtain the desired inequalities. The
proposition is proved.
For ψ0′ ̸= 0, by Proposition 13 a trajectory can be optimal only if

π − ψ0′
ω>p ′ , (7.4)
ψ0 (2π − ψ0′ )

and for ψ0′ = 0 the inequality tR 6 t2max cannot hold for any →−ω.
The map (w, ψ0 ) 7→ (w′ , ψ0′ ) defines a two-sheeted covering
   
π π π 3π
R+ × − , ∪ , → R+ × (0, π),
2 2 2 2
( π
ψ0′ = ψ0 + mod π,
2

ω = ω,

since the restrictions of this map to every sheet


  
π π
(ω, ψ0 ) : ω ∈ R+ , ψ0 ∈ − , ,
2 2
  
π 3π
(ω, ψ0 ) : ω ∈ R+ , ψ0 ∈ ,
2 2

are obviously the diffeomorphisms


( π ( π
ψ0′ = ψ0 + , ψ0′ = ψ0 − ,
mod+ : 2 mod− : 2
ω ′ = ω, ω ′ = ω.

We now consider a domain D in the space (ω, ψ0′ ) that characterizes the extremal
trajectories for which the inequality (7.4) is strict,

π − ψ0′
 
′ 2 ′
D = (ω, ψ0 ) ∈ R : ω > p ′ , 0 < ψ0 < π , (7.5)
ψ0 (2π − ψ0′ )

and the domains D+ = (mod+ )−1 D and D− = (mod− )−1 D. If D is regarded as


the interior of the epigraph of the function ω(ψ0′ ) (Fig. 6), then D+ and D− are
the translates of D by π/2 to the left and right, respectively.
182 I. Yu. Beschastnyı̌

Figure 6. The domain D.

The main result of this section is the proof of the fact that the following maps
are diffeomorphisms:

Exp′ : D+ → Q′+ , Exp′ : D− → Q′− ,

where the Q′± are the quarters of the sphere S 2 for which q1 > 0 and q3 > 0 or
q3 < 0, respectively.
Below we shall use the following theorem frequently.
Theorem 2 (see [8]). Let F : X → Y be a differentiable map between manifolds X
and Y . If
1) the map F is nondegenerate, that is, its Jacobian does not vanish,
2) F is proper, that is, the pre-image of a compact set is a compact set,
3) X and Y are connected,
4) Y is simply connected,
then F is a diffeomorphism.
We introduce an auxiliary change of variables α = α(ω, ψ0′ ), θ = θ(ω, ψ0′ ) by
means of the expressions

ω ω2 + 1
α = arccos √ , θ= (π − ψ0′ ). (7.6)
2
ω +1 ω

Let  
2 π
Π= (α, θ) ∈ R : 0 < α < , 0 < θ < π
2
and let f be the map defined by equations (7.6).
Lemma 2. The map f : D → Π is a diffeomorphism.
Proof. We apply Theorem 2. It is obvious that D is connected and Π is simply
connected. We claim that f (D) ⊂ Π. Since

dα 1
=− 2 < 0,
dω ω +1
The optimal rolling of a sphere 183

the function α(ω) is decreasing and its supremum andp infimum must be attained on
the boundary of the set D, that is, at ω = (π − ψ0′ )/ ψ0′ (2π − ψ0′ ) and ω → +∞,
respectively. We have
inf α(ω) = lim α(ω) = 0,
D ω→+∞
π − ψ0′
 
π
sup α(ω) = sup arccos = .
D ′
ψ0 ∈(0,π/2) π 2

Since the function α(ω) is strictly decreasing, we have 0 < α(ω, ψ0′ ) < π/2 for all
(ω, ψ0′ ) ∈ D.
Now, √
∂θ ω2 + 1
= − < 0,
∂ψ0′ ω
and so for any fixed ω the function θ = θ(ω, ψ0′ ) is a strictly decreasing function
with respect to ψ0′ , and therefore its supremum and infimum are attained on ∂D.
We have
inf θ = inf θ(ω, π) = 0,
D ω∈D
sup θ = sup θ(ω, 0) = lim θ(ω, 0) = π.
D ω∈D ω→+∞

Since the function θ(ω, ψ0′ ) is strictly decreasing for ω = const, we find that
0 < θ(ω, ψ0′ ) < π for all (ω, ψ0′ ) ∈ D.
The map f is nondegenerate on D, since
∂α ∂θ ∂θ ∂α 1
− = √ ̸= 0.
∂ω ∂ψ0′ ∂ω ∂ψ0′ ω ω2 + 1
Thus, to prove the lemma it remains to show that the map f is proper, that
is, the pre-image of a compact set is a compact set. Let Y be a compact set
in Π, and X = f −1 (Y ) its pre-image in D. We will prove that X is compact.
To do this we take any sequence {xi } ⊂ X and prove that it has a subsequence
converging in X. Let {xik } be a subsequence that has a limit (possibly, infinite).
To simplify the notation we denote it by {xi }. Since Y is compact, the sequence
{yi = f (xi )} is separated from the boundary of the set Π; therefore, ε < θi < π − ε
and ε < αi < π/2 − ε for some ε > 0.
1) Suppose that lim xi = ∞. Then ωi → ∞ and therefore, αi → 0, a contra-
diction.
2) Suppose thatplim xi = x ∈ ∂D. If (ψ0′ )i → π, then θi → 0, a contradiction. If
ωi → (π − (ψ0′ )i )/ (ψ0′ )i (2π − (ψ0′ )i ), then θi → π, a contradiction.
Consequently, the map f is proper and is a diffeomorphism by Theorem 2. The
lemma is proved.
Since f is a diffeomorphism, we can invert formulae (7.6) and substitute ω(α, θ)
and ψ0′ (α, θ) into (7.2):
q0 = cos θ cos(θ cos α) + cos α sin θ sin(θ cos α),
q1 = cos α sin θ cos(θ cos α) − cos θ sin(θ cos α), (7.7)
q3 = ± sin α sin θ.
184 I. Yu. Beschastnyı̌

7.2. Diffeomorphisms defined by the exponential map. There exist a pair


of maps g+ : Π → Q′+ and g− : Π → Q′− that transform the open rectangle Π into
the two quarters of a two-dimensional sphere
Q′+ = (q0′ , q1′ , q3′ ) ∈ R3 : |q ′ | = 1, q1′ > 0, q3′ > 0 ,

(7.8)
Q′− = (q0′ , q1′ , q3′ ) ∈ R3 : |q ′ | = 1, q1′ > 0, q3′ < 0 .


These maps are defined in the standard way and are the diffeomorphisms
q0′ = cos θ,
q1′ = cos α sin θ,
q3′ = ± sin α sin θ,
where q3′ is taken with plus for g+ , and with minus for g− .
Consider the following change of variables:
 ′
q1 cos−1 q0′
 ′
q1 cos−1 q0′
 
′ ′
q0 = q0 cos p + q1 sin p ,
1 − q0′2 1 − q0′2
 ′
q1 cos−1 q0′
 ′
q1 cos−1 q0′
 
(7.9)
q1 = −q0′ sin p + q1

cos p ,
1 − q0′2 1 − q0′2
q3 = q3′ .
Lemma 3. The map h defined by (7.9) maps Q′+ and Q′− diffeomorphically to
themselves.
Proof. The sets Q′+ and Q′− are connected and simply connected. It is clear from
(7.9) that the map h is a rotation in the plane q0′ Oq1′ , and, consequently, q3′ does
not change sign. Therefore the pre-image of each of the quarters is contained
in the upper or lower half of the sphere. Therefore, to prove h(Q′+ ) ⊂ Q′+ and
h(Q′− ) ⊂ Q′− we only need to show that q1 > 0.
Now, we set cos α = k. Then it follows from formulae (7.7) that
q1 (k, θ) = k sin θ cos(θk) − cos θ sin(θk)
1 
= k sin(θ + θk) + k sin(θ − θk) − sin(θk + θ) + sin(θ − θk)
2
1 
= (1 + k) sin(θ(1 − k)) − (1 − k) sin(θ(1 + k)) .
2
We differentiate the function we have obtained with respect to θ:
∂q1 1 − k2
cos(θ(1 − k)) − cos(θ(1 + k)) = (1 − k 2 ) sin θ sin θk.

=
∂θ 2
∂q1
Thus, > 0 on Π, and q1 (α, θ) only has an infimum on ∂Π for θ = 0. It
∂θ
follows from (7.7) that q1 (α, 0) = 0, and therefore q1 (α, θ) > 0 on Π. Consequently,
h(Q′+ ) ⊂ Q′+ .
The Jacobian of the map h is equal to
p
(1 − q0′2 − q1′2 )( 1 − q0′2 − q0′ cos−1 q0′ )
Jh = .
(1 − q0′2 )3/2
The optimal rolling of a sphere 185

It vanishes when

1 − q0′2 − q1′2 = 0, which is equivalent to q3′ = 0,

or q
1 − q0′2 − q0′ cos−1 q0′ = 0.

The set {(q0′ , q1′ , q3′ ) ∈ R3 : q3′ = 0} does not intersect either Q′+ or Q′− . Consider
p
the function s(q0′ ) = 1 − q0′2 − q0′ cos−1 q0′ . Its derivative is equal to

ds
= − cos−1 q0′ .
dq0′

Thus, the function attains extremum at the points qs = {q ′ : q0′ = ±1}, and
max s(q0′ ) = s(−1) = π and min s(q0′ ) = s(1) = 0. But qs ∈ / Q′+ and qs ∈/ Q′− ,
and therefore s(q0′ ) ̸= 0 for any q ′ ∈ Q′+ or q ′ ∈ Q′− . Consequently, the map h is
nondegenerate.
We will show that the map h : Q′+ → Q′+ is proper. The proof that h : Q′− → Q′−
is proper is similar.
Let Y ⊂ Q+ be compact. We will prove that X = h−1 (Y ) is also compact. As
in the proof of Lemma 2, we take any sequence {xi } ∈ X and choose a subsequence
{xik } that has a limit. To simplify the notation we denote it by {xi }. Since Y is
compact, (q1 )i > ε and (q3 )i > ε for some ε > 0.
Suppose that lim xi ∈ ∂Q′+ . If (q1′ )i → 0, then (q1 )i → 0, which is a contra-
diction. If (q3′ )i → 0, then (q3 )i → 0, again a contradiction. Consequently, h is
a proper map, and it is a diffeomorphism of Q′+ onto itself by Theorem 2. Similar
arguments imply that h : Q′− → Q′− is also a diffeomorphism. The lemma is proved.
The maps h ◦ g+ and h ◦ g− are defined in coordinates by equations (7.7). There-
fore the following theorem holds.
Theorem 3. The restricted exponential map (7.3) defines the following diffeomor-
phisms:
Exp′ : D+ → Q′+ and Exp′ : D− → Q′− .
Proof. It follows from Lemmas 2 and 3 that the following diagrams are commuta-
tive:
f ◦mod+ f ◦mod−
D+ /Π D− /Π

Exp′ g+ Exp′ g−
   
Q′+ o Q′− o
h h
Q′+ Q′− .

Since mod± , f , g± and h are diffeomorphisms, it follows that Exp′ is also a diffeo-
morphism. The theorem is proved.
This theorem shows that equations (7.2) have a unique solution for any point
q ∈ Q′+ or q ∈ Q′− . Indeed, suppose that for some q ∈ Q′+ there exist two solutions
of (7.2). Then both points belong to D+ , since D− ∩ D+ = ∅, and points in ∂D±
are mapped into ∂Q± ; this follows from the definition of the set Π and (7.7). But
this contradicts the fact that Exp′ is a diffeomorphism between D+ and Q+ .
186 I. Yu. Beschastnyı̌


7.3. Re-orienting the sphere for an arbitrary ω . The results obtained in the
preceding subsections can be combined into the following theorem.

Theorem 4. Let R1 ∈ SO(3) be the matrix corresponding to the rotation of the


sphere about a unit vector →−a = (a1 , a2 , a3 ) with a21 + a22 ̸= 1 and |a3 | ̸= 1 through
some angle. Then there exist two extremal trajectories taking the sphere from the
state Q0 = (0, 0, Id) to Q1 = (0, 0, R1 ) such that tR < t2max .

Proof. By Theorem 3 the exponential map factorized by rotations defines the two
diffeomorphisms Exp′ : D+ → Q′+ and Exp′ : D− → Q′− , where Q′+ and Q′− are the
two quarters of the sphere defined in (7.8).
By (3.26) the equations q1′ = 0, q2′ = 0 imply that a1 = 0, a2 = 0. But since the
vector → −
a is a unit vector, it follows that |a3 | = 1. Similarly the equations q3′ = 0,

q2 = 0 imply that |a1 | = 1. Since one orientation matrix has two corresponding
quaternions ±q, for → −ω = (ω, 0, 0) there exist two extremal trajectories approaching
the point Q1 = (0, 0, R1 ) if R1 is the matrix of a rotation about a vector → −
a =

(a1 , 0, a3 ) such that a1 ̸= 1 and a3 ̸= 1. Therefore the map Exp is a two-sheeted
covering.
We now consider the case when the vector → −
ω ∈ R2 is arbitrary. Under the
β →

action of the symmetry Φ the vector ω is rotated through the angle β, and
the orientation quaternion q± = ±(q0 + q1 i + q2 j + q3 k) goes to the quaternion
β
q± = ±(q0 + (q1 cos β − q2 sin β)i + (q1 sin β + q2 cos β)j + q3 k); this is easy to see
using (3.25). Then by (3.26) the vector → −a is rotated through the angle β, that is,
β
q is the rotation quaternion corresponding to the matrix of a rotation about the
vector → −
a β = (a1 cos β − a2 sin β, a1 sin β + a2 cos β, a3 ) through some angle. Since
Φ is a symmetry of the exponential map, Exp′ is also a two-sheeted covering in
β

this case. The theorem is proved.

Suppose that R1 is a rotation through an angle ϕ about a unit vector → −a =


2 2
(a1 , a2 , a3 ) with a1 + a2 ̸= 1 and |a3 | =
̸ 1. Then the orientation of the sphere at the
moment of time t1 is defined by the pair of quaternions
 
ϕ ϕ
q± = ±(q0 + q1 i + q2 j + q3 k) = ± cos + (a1 i + a2 j + a3 k) sin .
2 2

In order to find the conjugate variables to which the desired optimal trajectory
corresponds, we act on each of the quaternions by the rotation Φβ , where β is
determined from the relations
q1 q2
cos β = ± p 2 , sin β = ∓ p .
q1 + q22 q12 + q22

Then the terminal orientation is defined by the quaternion


q
β
q± = ±q0 + q1β i ± q3 k, q1β = q12 + q22 .

In particular, this implies that the vector →



ω is perpendicular to the vertical plane


passing through the vector a .
The optimal rolling of a sphere 187

The values of ω and ψ0′ are found from equations (7.2), that is,
√ 2 
ω +1
− cos (π − ψ0′ ) cos ψ0′
ω
√ 2 
ω ω +1
+√ sin (π − ψ0 ) sin ψ0′ = ±q0 ,

ω2 + 1 ω
√ 2 
ω ω +1
−√ sin (π − ψ0′ ) cos ψ0′ (7.10)
ω2 + 1 ω
√ 2 
ω +1
− cos (π − ψ0′ ) sin ψ0′ = q1β ,
ω
√ 2 
1 ω +1 ′
−√ sin (π − ψ0 ) = |q3 |.
ω2 + 1 ω

By Theorem 3 we obtain a pair of solutions (ω, ψ0′ ), which define two extremal
trajectories satisfying the boundary conditions. We choose the values of (ω, ψ0′ ) to
which the shorter trajectory corresponds by comparing the times tR using formula
(6.2). Next, the values of ψ0 and →
−ω are determined as follows:
π →

ψ0 = ψ0′ − sign(q3 ) , ω = (ω cos β, ω sin β, 0),
2
and they define the desired control and optimal trajectory. Thus, if a matrix R1 sat-
isfies the hypothesis of Theorem 4, then solving the problem of optimally re-orient-
ing the sphere is reduced to solving equations (7.9) in the domains D+ and D− .
7.4. The induced metric on SO(3). The length functional (2.4) defines a met-
ric on R2 × SO(3), known as the Carnot-Carathéodory metric. In this case the
distance between points Q0 , Q1 ∈ R2 × SO(3) is the infimum of the lengths of
admissible curves connecting these points.
Let S(R0 , R1 ) denote the set of all extremal curves connecting Q0 = (0, 0, R0 )
with Q1 = (0, 0, R1 ). Then the Carnot-Carathéodory metric induces the following
metric on SO(3):
d(R0 , R1 ) = inf l(Q), (7.11)
Q∈S(R0 ,R1 )

where l(Q) is the length (2.4) of the curve Q(t). We describe some properties of
this metric: maximally distant pairs and the diameter (the supremum of the dis-
tances between pairs of points). This metric is left-invariant by construction. Con-
sequently, we can seek the most distant point from R0 = Id. This problem reduces
to finding an optimal trajectory with the maximal value of tR = 2(π − ψ0′ )/ω.
Consider the closure D (see (7.5)) in the space (ω, ψ0′ ). It follows from Theorem 4
that all the rotations about a unit vector → −a = (a1 , a2 , a3 ) with |a3 | ̸= 1 and
a1 + a2 ̸= 1 correspond to the variables (ω, ψ0′ ) ∈ D. The rotations about →
2 2 −
a =

(a1 , a2 , a3 ) with a3 = 0 p
correspond to the variables (ω, ψ0 ) in ∂D. Indeed, we
substitute ω = (π − ψ0′ )/ ψ0′ (2π − ψ0′ ) into (7.2) and find that when it returns to
the origin the sphere turns out to be in the state

q0 = cos ψ0′ , q1 = sin ψ0′ , q2 = 0, q3 = 0.


188 I. Yu. Beschastnyı̌

Proposition 14. If the terminal state of the sphere in the re-orientation problem
is R1 = eθA1 , θ ∈ [0, 2π),
p then rolling along a segment of a straight line is optimal
only if ω = (π − ψ0′ )/ ψ0′ (2π − ψ0′ ) and ψ0′ 6 π/2.
Proof. The quaternion corresponding to the matrix eθA1 has the form
 
cos θ sin θ
q=± +i .
2 2
The fact that (ω, ψ0′ ) ∈ ∂D is obvious, since for (ω, ψ0′ ) ∈ D the component a3
is nonzero by Theorem 4. Then, as indicated in § 7.3, we need to solve the two
systems
  
θ ′ θ ′ θ
cos = cos ψ0 , cos = − cos ψ0 , cos = cos(π − ψ0′ ),

 
 

2 2 2
and ⇐⇒
θ θ θ
sin = sin ψ0′ sin = sin ψ0′ sin = sin(π − ψ0′ ).

 
 

2 2 2
For any θ each of these systems has a unique solution ψ0′ . Consequently, we
have a pair of extremal trajectories, from which we p need to choose the one with
the smaller value of tR . We substitute ω = (π − ψ0′ )/ ψ0′ (2π − ψ0′ ) into (6.2) and
obtain on ∂D the expression
q
tR (ψ0′ ) = 2 ψ0′ (2π − ψ0′ ). (7.12)

The function tR (ψ0′ ) is an increasing function on [0, π], and therefore the smaller of
the two values will be optimal: either ψ0′ , or π − ψ0′ . This implies the assertion of
the proposition. The proposition is proved.
The hypothesis of Proposition 14 and the fact that the function tR is increasing
imply that for an optimal
√ trajectory on ∂D the time √ tR takes its maximal √ value
at ψ0′ = π/2, ω = 1/ 3 and it is equal to tR = π 3. We claim that π 3 is the
maximal value for all optimal trajectories. √
First we verify that if |a3 | = 1, then the cut time satisfies tcut < π 3. Indeed, it
follows from the description of the set D that its image does not contain points with
q1 = 0 or q2 = 0. Therefore the sphere cannot get into the state eθA3 moving along
a segment of a straight line. Consequently, twisting√on the spot must be optimal.
But it follows from Proposition 9 that tcut 6 π < π 3 for such a motion.
We now consider the case |a3 | ̸= 1. The fact that tR is decreasing (see formula
(6.2)) both with respect to ω and with respect
√ to ψ0′ , together √
with the fact that

the maximum time is attained at ω = 1/ 3, imply that √ tR 6 π 3 for ω > 1/ 3.
Consequently, it only remains to verify the case ω < 1/ 3, which is shown in Fig. 7.

Proposition 15. In √ the problem of re-orienting the sphere for ω < 1/ 3 the cut
time satisfies tcut < π 3.
Proof. By Proposition 12 the set q0 = 0 is contained in the Maxwell set. Con-
sequently, if a trajectory crosses this set at a moment of time t, then it is not
√ q(t, ω) going out from the point q = 1; we
optimal later. Consider the trajectories
claim that for ω < 1/3√ at time t = π 3 the corresponding trajectories cross the set
q0 = 0, that is, q0 (π 3, ω) < 0.
The optimal rolling of a sphere 189


Figure 7. The domain D. In the non-coloured subdomain, tR 6 π 3.

From (3.27) we have


 p   p 
t ωt ω t ωt
q0 (t, ω) = cos 2
ω + 1 cos + √ sin 2
ω + 1 sin .
2 2 2
ω +1 2 2

We differentiate this function with respect to ω and find that at time t = π 3
√   √ p
ωπ 3 √ p

∂q0 1 2
π 3 2
=− sin π 3 1 + ω cos ω +1
∂ω t=π√3 2(1 + w2 )3/2 2 2
 √ 
π 3p 2
− 2 sin ω +1 .
2

For ω < 1/ 3 we have

ωπ 3
sin > 0,
√ 2
 √ p
√ p
  
2
π 3p 2 π 3
π 3 1 + ω cos ω + 1 − 2 sin ω2 + 1
2 2
 √ p
√ p

π 3
6 π 3 1 + ω 2 cos ω 2 + 1 < 0.
2

√ the function q0 (π 3, ω)√is increasing in √
Consequently,
√ ω. And since the equation
q0 (π 3, 1/ 3) = 0 holds, for √ ω < 1/ 3 we have q0 (π 3, ω) < 0. Consequently,
the cut time is smaller than π 3. The proposition is proved.
Thus, we have proved the following theorem.
Theorem 5. In the induced metric (7.11) on SO(3) the points furthest apart √ are
Id and eπ(a1 A1 +a2 A2 ) and the distance between them is d(Id, eπ(a1 A1 +a2 A2 ) ) = π 3.
The states eπ(a1 A1 +a2 A2 ) correspond to the sphere turned onto opposite pole.
By comparison, on SO(3) with the metric generated by the Killing form, the
furthest points from Id with diameter equal to π are rotations about any vector
through an angle √ π. Therefore the ratio of the diameters in this metric and the
induced metric is 3.
190 I. Yu. Beschastnyı̌

§ 8. Conclusion
In this paper we have studied the problem of the optimal rolling of a sphere on
a plane with twisting but no slipping. We have obtained a complete parametrization
of the extremal trajectories, analysed the symmetries of the exponential map, found
their fixed points, and used them to obtain estimates for the cut time.
We have also considered the problem of re-orienting the sphere: we have found
the terminal states of the sphere for which the parameters of an optimal trajectory
are uniquely determined from equations (7.10). We have described some properties
of the left-invariant metric on the group of rotations of three-dimensional space
related to this problem.
Further studies could be directed to solving the optimal control problem for an
arbitrary terminal position of the sphere Q1 . To do this, first it would be neces-
sary to consider the symmetries of the Hamiltonian system ε4 –ε7 and to construct
the corresponding symmetries of the exponential map, find their fixed points, and
compare the new Maxwell times with those that have already been found in this
paper.
The author thanks Yu. L. Sachkov and A. A. Agrachev for posing the problem
and for some useful discussions during this work.

Bibliography

[1] J. M. Hammersley, “Oxford commemoration ball”, Probability, statistics and


analysis, London Math. Soc. Lecture Note Ser., vol. 79, Cambridge Univ. Press,
Cambridge–New York 1983, pp. 112–142.
[2] A. M. Arthurs and G. R. Walsh, “On Hammersley’s minimum problem for a rolling
sphere”, Math. Proc. Cambridge Philos. Soc. 99:3 (1986), 529–534.
[3] V. Jurdjevic, “The geometry of the plate-ball problem”, Arch. Rational Mech. Anal.
124:4 (1993), 305–328.
[4] Yu. L. Sachkov, “Maxwell strata and symmetries in the problem of optimal rolling
of a sphere over a plane”, Mat. Sb. 201:7 (2010), 99–120; English transl. in
Sb. Math. 201:7 (2010), 1029–1051.
[5] A. P. Mashtakov and Yu. L. Sachkov, “Extremal trajectories and the asymptotics
of the Maxwell time in the problem of the optimal rolling of a sphere on a plane”,
Mat. Sb. 202:9 (2011), 97–120; English transl. in Sb. Math. 202:9 (2011),
1347–1371.
[6] A. P. Mashtakov, “Asymptotics of extremal curves in the problem of the rolling of
a ball on a plane”, Proc. Int. Conf. on Mathematical Control Theory and Mechanics
(Suzdal’ 2009), Sovrem. Mat. Fundam. Napravl., vol. 42, Peoples’ Friendship
University of Russia, Moscow 2011, pp. 158–165. (Russian)
[7] I. Moiseev and Yu. L. Sachkov, “Maxwell strata in sub-Riemannian problem on
the group of motions of a plane”, ESAIM Control Optim. Calc. Var. 16:2 (2010),
380–399.
[8] Yu. L. Sachkov, “Conjugate and cut time in the sub-Riemannian problem on the
group of motions of a plane”, ESAIM Control Optim. Calc. Var. 16:4 (2010),
1018–1039.
[9] A. A. Ardentov and Yu. L. Sachkov, “Extremal trajectories in a nilpotent
sub-Riemannian problem on the Engel group”, Mat. Sb. 202:11 (2011), 31–54;
English transl. in Sb. Math. 202:11 (2011), 1593–1615.
The optimal rolling of a sphere 191

[10] Yu. L. Sachkov, “Exponential map in the generalized Dido problem”, Mat. Sb.
194:9 (2003), 63–90; English transl. in Sb. Math. 194:9 (2003), 1331–1359.
[11] Yu. L. Sachkov, “Complete description of the Maxwell strata in the generalized
Dido problem”, Mat. Sb. 197:6 (2006), 111–160; English transl. in Sb. Math. 197:6
(2006), 901–950.
[12] S. Popov, Extremal trajectories in the problem of the rolling of a sphere on a plane
with twisting, Diploma Thesis, Moscow State University 2009. (Russian)
[13] A. A. Agrachev and Yu. L. Sachkov, Geometric control theory, Fizmatlit, Moscow
2005. (Russian)
[14] M. M. Postnikov, Lectures in geometry. Semester V. Riemannian geometry,
Faktorial, Moscow 1998; English transl. Geometry VI. Riemannian geometry,
Encyclopaedia Math. Sci., vol. 91, Springer-Verlag, Berlin 2001.
[15] V. I. Arnol’d, Geometry of complex numbers, quaternions, and spins, Moscow
Centre for Continuous Mathematical Education, Moscow 2002. (Russian)

Ivan Yu. Beschastnyı̌ Received 28/OCT/13


Institute of program systems Translated by E. KHUKHRO
of the Russian Academy of Sciences,
Pereslavl’-Zalesskiı̌
E-mail: i.beschastnyi@gmail.com

Das könnte Ihnen auch gefallen