Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

ISSN 1560-3547, Regular and Chaotic Dynamics, 2012, Vol. 17, No. 2, pp. 191198. c Pleiades Publishing, Ltd.

Publishing, Ltd., 2012.


Two Non-holonomic Integrable Problems
Tracing Back to Chaplygin
Alexey V. Borisov
*
and Ivan S. Mamaev
**
Institute of Computer Science,
Udmurt State University
ul. Universitetskaya 1, Izhevsk, 426034 Russia
Received August 14, 2011; accepted November 29, 2011
AbstractThe paper considers two new integrable systems which go back to Chaplygin. The
systems consist of a spherical shell that rolls on a plane; within the shell there is a ball or
Lagranges gyroscope. All necessary rst integrals and an invariant measure are found. The
solutions are shown to be expressed in terms of quadratures.
MSC2010 numbers: 76M23, 34A05
DOI: 10.1134/S1560354712020074
Keywords: non-holonomic constraint, integrability, invariant measure, gyroscope, quadrature,
coupled rigid bodies
INTRODUCTION
Recent advances in the design of controlled devices that use one or several balls for their
propulsion (see, e.g., [3, 4, 710, 1214]) has recently evoked an increasing interest in various models
(in particular, non-holonomic ones) for rolling motion of spherical shells, including the case where
some of the shells contain intricate mechanisms inside. The classical problem of rolling motion of
a dynamically non-symmetric balanced Chaplygin ball has been suciently well investigated [11].
A newer version of it, namely the rolling of a Chaplygins ball over a sphere, has recently been
discussed in a number of papers [1, 5, 6].
Here we consider two new integrable systems that trace back to Chaplygin. His paper [2] is
essentially concerned with generalized conditions for the existence of integrals linear in velocities
for mechanical systems that consist of several spheres. These conditions are even now far from
being completely appreciated. Chaplygin discusses in detail two problems. The rst of them deals
with a system that consists of a spherical, geometrically and dynamically symmetric shell with
a homogeneous ball rolling inside; the shell itself rolls without slipping on a horizontal plane
(Fig. 1). Chaplygin established the integrability of this system by expressing the solutions in terms
of quadratures. Although his method for obtaining quadratures is quite natural (and also applies to
the problem of rolling of a body of revolution on a plane), we believe that in solving this problem
Chaplygin committed a few inaccuracies which resulted in enormously complicated and hardly
veriable formulas. Besides, Chaplygin missed one of the additional rst integrals. The complexity
of the results seems to discourage Chaplygin himself. Indeed, for each newly obtained analytical
solution he always pursued clarication of its dynamical and geometrical aspects, but not in that
case! Here we present a new approach to this system, a complete set of rst integrals and reduce
the problem to quadratures.
The second problem is concerned with rolling of a shell (with a spherical pendulum inside) on
a plane. For this problem the equations of motion were integrated by Chaplygin. We show that
a more general system which consists of a Lagranges gyroscope placed inside a rolling sphere is
also integrable. The equations can be integrated using some generalized versions of the Chaplygin
vector integrals and two additional linear integrals.
*
E-mail: borisov@rcd.ru
**
E-mail: mamaev@rcd.ru
191
192 BORISOV, MAMAEV
1. THE CHAPLYGIN SYSTEM
1.1. Equations of Motion
Consider a system that consists of two bodies: 1) a spherical shell rolling on a horizontal plane
inside which there is a spherical cavity such that its center coincides with the shells center and 2)
a ball placed into the spherical cavity of the shell (Fig. 1). Both bodies are assumed to be balanced
(i. e. the center of mass coincides with the geometrical center) and dynamically symmetric (i. e. the
tensor of inertia is spherical). We denote their moments of inertia by I and i respectively.
Fig. 1
Choose a xed reference frame Oxyz with origin in a horizontal plane through the center of the
shell and with the Ozaxis directed vertically downwards. Let us write the equations for balance of
the angular and linear momentum of the bodies. For the shell the balance of the angular momentum
relative to its center G
s
and the balance of the linear momentum read
I

= R
o
k N
o
+R
i
n N
i
, M

V = N
o
+N
i
+Mgk, (1.1)
where and V are the angular velocity of the shell and the linear velocity of its center G
s
, M is
the mass of the shell, R
o
and R
i
are its inner and outer radii, k = (0, 0, 1) is the unit vector along
the force of gravity, N
o
and N
i
are the reaction forces acting on the shell at the points of contact
with the plane Q
o
and the ball Q
i
, g is the acceleration of gravity. Similar equations for the ball
relative to its center G
b
are
i = R
b
n (N
i
), m v = N
i
+mgk, (1.2)
here and v are the balls angular velocity and the linear velocity of its center G
b
, m is the mass
of the ball, and R
b
is its radius.
We shall assume that there is no slipping at the points of contact Q
o
and Q
i
, i. e. the velocity
of the point of contact Q
o
is zero:
V +R
o
k = 0, (1.3)
and at the point Q
i
the velocities of the contacting elements of the shell and the ball coincide:
V + R
i
n = v + R
b
n. (1.4)
These equations express the (non-holonomic) constraints imposed on the system. Using them, we
eliminate the reaction forces from (1.1) and (1.2); this allows derivation of a closed system of
equations governing the evolution of the vectors , and n.
According to the denition of the vector n (see Fig. 1), we have v V = (R
i
R
b
) n, and (1.4)
gives
(R
i
R
b
) n = (R
i
R
b
) n.
We express N
o
from the second equation in the system (1.1) and

V from (1.3), whence we nd
R
o
k N
o
= R
o
k (N
i
Mgk MR
o

k).
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
TWO NON-HOLONOMIC INTEGRABLE PROBLEMS 193
This allows us to represent the equations for and in the form
I

+MR
2
o
k (

k) = J

= a N
i
, i = R
b
n N
i
, (1.5)
where J = diag(I +MR
2
o
, I +MR
2
o
, I) and a = R
i
n R
o
k =

Q
o
Q
i
is the vector connecting the
points of contact, and the reaction force N
i
can be found from the second equation (1.2) and the
constraint (1.4):
N
i
= m v +mgk, v = a R
b
n. (1.6)
Dierentiating v and simplifying, we ultimately obtain:
The equations of motion for the shell and the ball rolling without slipping on a plane can be
written as
J

+ma (

a) mR
b
a ( n) = ma
_
(R
i
R
b
) n
_
+mgR
i
n k,
i +mR
2
b
n( n) mR
b
n
_

a
_
= mR
b
n
_
(R
i
R
b
) n
_
mgR
b
nk,
(R
i
R
b
) n = (R
i
R
b
) n,
(1.7)
where a = R
i
n R
o
k, k = (0, 0, 1). The trajectory of the shells center r
s
(t) (point of
contact) is given by
r
s
= V = R
o
k . (1.8)
1.2. Invariant Measure, First Integrals, Gyroscopic Function
Equations (1.7) admit an invariant measure dddn whose density on the level surface of
the geometrical integral n
2
= 1 is as follows:
=
1

2
,

1
(n
3
) =
I +MR
2
o
+m
e
(R
i
R
o
)
2
m
e
R
2
o
+
2R
i
R
o
(1 n
3
) +
mR
2
b
i
(1 n
2
3
),

2
(n
3
) =
I +MR
2
o
+m
e
(R
i
R
o
)
2
m
e
R
2
o
+
2R
i
R
o
(1 n
3
) +
MR
2
i
I
(1 n
2
3
),
(1.9)
where, as in celestial mechanics, we refer to m
e
=
im
i+mR
2
b
as the reduced mass of the ball.
Remark. Concatenating the angular velocities into a single vector w = (, ), we can write the
rst six equations of the system (1.7) as
G w = b(w, n),
where G is a 6 6frm[o]matrix whose components depend on n and b is a six-dimensional vector.
It turns out that
det G =
1

2
,
where is a constant.
The equations of motion (1.7) admit obvious rst integrals:
the geometric integral n
2
= 1,
energy E =
1
2
(MV
2
+I
2
+mv
2
+i
2
) mg(R
i
R
b
)(k, n),
(1.10)
where V and v are expressed from (1.3) and (1.4) as follows:
V = R
o
k, v = a R
b
n.
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
194 BORISOV, MAMAEV
Moreover, Eqs. (1.7) also admit the vector integral
K = J +
R
i
R
b
i +mR
o
v k (1.11)
which is linear in the angular velocities and . (The simplest way to nd this integral is to
use (1.5) and (1.6).)
In addition to these three integrals, there are two more integrals linear in the angular velocities:
F
1
= (J, R
i
n R
o
k) IR
b
(, n) MR
o
R
i
(v k, n) +
MR
o
R
i
mR
b
(i, k),
F
2
= (R
i
R
b
, n)

2
.
(1.12)
Remark. The existence of the vector integral K was discovered by S. A. Chaplygin [2] during his
study of general mechanical systems that admit vector integrals linear in velocities. In this context,
Chaplygin treated this system as a comparatively trivial but ultimately edifying example.
For this specic system Chaplygin found the general integrals (1.10), (1.11) and the integral F
1
(but did not indicate the integral F
2
and measure). His attempts of integrating the equations by
means of various substitutions led to a collection of bulky and unwieldy expressions (which most
likely are fallacious).
Thus, the system of nine equations (1.7) admits seven rst integrals and an invariant measure
and, therefore, is integrable (by the Euler Jacobi theorem). Moreover, on a xed level of the rst
integrals
M
,
= {(, , n) | n
2
= 1, K = , F
1
=
1
, F
2
=
2
}
the energy integral can be represented as
E =

i
2

1
n
2
3
1 n
2
3
mg(R
i
R
b
)n
3
+U

(n
3
) +
1
2

2
1
+
2
2
I +i
R
2
i
R
2
b
+ (m+M)R
2
o
,

i = i
mR
2
o
_
R
2
i
R
2
b
1
_
I +i
R
2
i
R
2
b
+ (M +m)R
2
o
,
U

(n
3
) = A
0

2
1
+A
1

3
+A
2

2
3
+
B
0

2
+
B
1

2
+
C
0

2
2
.
Here A
k
are quadratic polynomials in n
3
, B
k
are cubic polynomials, and C
0
are polynomials of
degree four. The coecients of the polynomials depend in a complicated fashion on the parameters
of the system I, i, M, . . ., and for this reason we do not write them out here (they can be easily
found using any system of analytic computations, for example, Maple, Mathematica, etc.). For a
xed value of energy
E =
1
2

2
1
+
2
2
I +i
R
2
i
R
2
b
+ (m+M)R
2
o
+h
the derivative n
3
can be found from the equation
n
2
3
=
2(1 n
2
3
)

i
1
(n
3
)
_
h +mg(R
i
R
o
)n
3
U

(n
3
)
_
, (1.13)
where the gyroscopic function of the system is contained within the brackets.
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
TWO NON-HOLONOMIC INTEGRABLE PROBLEMS 195
1.3. A spherical shell with a pendulum
There is an interesting limiting case of this Chaplygin system which is equivalent to the problem
of rolling of a shell with a spherical pendulum xed at its center. To illustrate this, consider a
formal problem: an ball rolls (without slipping) over the outer side of a spherical surface which is
xed at the center of the shell (Fig. 2). Since the vector from the point of contact Q
i
to the center
of the ball G
b
is now pointed in the opposite direction, we must replace R
b
by R
b
in Eqs. (1.7).
Then we let R
i
go to zero and thereby obtain a system for which the functions (1.10) and (1.11)
remain integrals of motion and the second of the functions (1.12) simplies to
(, n) = const.
On the zero level of this integral (, n) = 0 we get a system that is equivalent to a ball with a
spherical pendulum.
Fig. 2
It is interesting to note that the rst of the integrals (1.12) becomes degenerate and naturally
reincarnates into the integral

3
= const.
Here we do not analyze this problem in detail because in the next section we consider a more
general system that includes it as a particular case.
Remark. Chaplygin himself also considered the problem of a shell with a spherical pendulum. He
dealt with the conguration depicted in Fig. 1, but assumed the inner surface of the shell to be
absolutely smooth. In this case he explicitly found all necessary integrals and the corresponding
gyroscopic function.
2. A SPHERICAL SHELL WITH LAGRANGES TOP
2.1. Equations of Motion
We now consider in greater detail a more general system which cannot be obtained as a particular
or limiting case of the Chaplygin system, namely a spherical shell rolling without slipping on a
horizontal plane with an axisymmetric top at the center of the shell (Fig. 3). Similar problems arise
in the design of control strategies for ballbots. One of the most intriguing projects here is a ballbot
for interplanetary missions (e.g. a Mars rover, [14]).
As in the previous case, we associate a xed reference frame Oxyz with a horizontal plane Oxy
passing through the center of the ball. Let G
s
be the center of mass of the shell, G
t
the center of
mass of the top and let R
t
= |G
s
G
t
| denote the distance between them. Assuming that in the tops
principal axes frame its tensor of inertia is

i = diag(i, i, i +j), we can represent the kinetic energy


of the system as
T =
1
2
(MV
2
+I
2
) +
1
2
(mv
2
+i
2
+j(, n)
2
),
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
196 BORISOV, MAMAEV
Fig. 3
Here V and are the linear velocity of the center of the shell and the shells angular velocity, M
and I stand for its mass and moment of inertia, v and are the linear velocity of the center of
mass of the top and its angular velocity, and nally m is the mass of the top.
The equations for balance of the angular momentum relative to the point G
s
and balance of the
shells linear momentum read
_
T

_
.
= I

= R
o
k N
o
,
_
T
V
_
.
= M

V = N
o
+N
t
+Mgk,
where N
o
, N
t
are the reaction forces on the shell applied at the point of contact Q
o
and the point
where the top G
s
is attached. Similar equations for the top (relative to its center of mass G
t
) are
_
T

= (i +j(, n)n)

= R
t
n N
t
,
_
T
v
_

= m v = mgk N
t
.
The no-slip condition at the point of contact Q
o
implies that
V = R
o
k ,
whereas the velocity of the center of mass of the top is
v = V +R
t
n = R
o
k +R
t
n.
The evolution of the vector n obtained from the equation R
t
n = v V reads
n = n.
Using these relations and the fact that there is an integral (, n) = const, we can get rid of the
reaction forces N
o
, N
t
(as this was done for the Chaplygin system) and thus obtain the following
result:
The equations of motion for a spherical shell with an axisymmetric top xed at its geometrical
center can be represented as
J

+mR
2
o
k (

k) mR
o
R
t
k ( n) = mR
o
R
t
k ( n),
i mR
2
t
n ( n) mR
o
R
t
n (

k)
= j(, n) nmR
2
t
n ( n) +mgR
t
n k,
n = n.
(2.1)
Here J = diag(I +MR
2
o
, I +MR
2
o
, I) and k = (0, 0, 1). The trajectory of the point of contact
of the shell with the plane is then also given by (1.8).
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
TWO NON-HOLONOMIC INTEGRABLE PROBLEMS 197
2.2. Invariant Measure, First Integrals and Gyroscopic Function
To save space, we put

I = I + (M +m)R
2
o
, J
s
= diag(

I,

I, I),
=
j

I mR
2
t
(I +mR
2
o
)
i

I +mR
2
t
(I +mR
2
o
)
.
The density of the invariant measure dddn from Eqs. (2.1) is form
(n
3
) =
(i +mR
2
t
)

I
m
2
R
2
t
R
2
o
n
2
3
.
As in the previous case, this system admits seven rst integrals:
geometrical integral n
2
= 1,
energy E =
1
2
(MV
2
+I
2
) +
1
2
(mv
2
+i
2
+j(, n)
2
) mgR
t
(n, k),
Chaplygins vector
integral
K = J+mR
o
v k,
linear integrals F
1
=
3
+(, n)n
3
,
F
2
= (, n),
where
V = R
o
k, v = R
o
k +R
t
n.
On a common level of the rst integrals
M
,
=
_
(, , n) | n
2
= 1, K = , F
1
=
1
, F
2
=
2
_
the energy of the system can be written as
E =
1
2
(, J
1
s
) +
m
2
R
2
o
R
2
t
2

I
(n
3
) n
2
3
1 n
2
3
+U

(n
3
) mgR
t
n
3
,
U

=
1
2(1 n
2
3
)
_
_
i +
mR
2
t
(I +MR
2
o
)

I
_

2
1
2(i +j)n
3

2
+ (i +j)(1 +n
2
3
)
2
2
_
.
Thus, on a level surface of the energy integral
E =
1
2
(, J
1
s
) +h
the evolution n
3
is governed by the equation
n
2
3
=
2

I(1 n
2
3
)
m
2
R
2
o
R
2
t
(n
3
)
_
h +mgR
t
n
3
U

(n
3
)
_
,
where the gyroscopic function of the system is contained within the brackets.
3. DISCUSSION
In conclusion we highlight several problems which can be solved using the results of this paper.
1. Note that the property of integrability is preserved when a ball rolls on the outer surface of a
shell (Fig. 4). This can seriously facilitate a complete analysis of stability (stabilization) of rotation
and rolling of the ball on the top of the shell. Moreover, the shell itself can be allowed to roll.
(It is interesting to note that similar arrangements are often used in circus performances where
maintaining equilibrium is the actors concern.)
2. The systems described in this paper are ubiquitous in control applications (an extensive list
of references can be found in [9]). Taking advantage of their integrability, one can develop a set of
elementary stable solutions which can be used for solving various control problems.
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012
198 BORISOV, MAMAEV
Fig. 4
ACKNOWLEDGMENTS
This research was supported by the Grant of the Government of the Russian Federation for
state support of scientic research conducted under supervision of leading scientists in Russian
educational institutions of higher professional education (contract no. 11.G34.31.0039) and the
Federal target programme Scientic and Scientic-Pedagogical Personnel of Innovative Russia,
measure 1.5 Topology and Mechanics (project code 14.740.11.0876). The work was supported
by the Grant of the President of the Russian Federation for the Leading Scientic Schools of the
Russian Federation (NSh-2519.2012.1).
REFERENCES
1. Borisov, A. V., Kilin, A. A., and Mamaev, I. S., Generalized Chaplygins Transformation and Explicit
Integration of a System with a Spherical Support, Regul. Chaotic Dyn., 2012, vol. 8, no. 2, pp. 170190.
2. Chaplygin, S. A., On Some Generalization of the Area Theorem with Applications to the Problem of
Rolling Balls, Regul. Chaotic Dyn., 2012, vol. 8, no. 2, pp. 199217 [Russian original: Mat. Sb., 1897, Vol.
20; reprinted in: Collected Works: Vol. 1, MoscowLeningrad: Gostekhizdat, 1948, pp. 2656].
3. Alves, J. and Dias, J., Design and Control of a Spherical Mobile Robot, J. Systems and Control
Engineering, 2003, vol. 217, pp. 457467.
4. Bhattacharya, S. and Agrawal, S. K., Design, Experiments and Motion Planning of a Spherical Rolling
Robot, in Proc. of the IEEE Internat. Conf. on Robotics and Automation (San Francisco, CA, April
2000), IEEE, 2000, pp. 12071212.
5. Borisov, A. V., Kilin, A. A., and Mamaev, I. S., Rolling of a Homogeneous Ball over a Dynamically
Asymmetric Sphere, Regul. Chaotic Dyn., 2011, vol. 16, no. 5, pp. 465483.
6. Borisov, A. V., Fedorov, Yu. N., and Mamaev, I. S., Chaplygin Ball over a Fixed Sphere: An Explicit
Integration, Regul. Chaotic Dyn., 2008, vol. 13, no. 6, pp. 557571.
7. Camicia, C., Conticelli, F., and Bicchi, A., Nonholonimic Kinematics and Dynamics of the Sphericle,
in Proc. of the 2000 IEEE/RSJ Internat. Conf. on Intelligent Robots and Systems (Takamatsu, Japan,
Oct. 31 Nov. 5 2000), IEEE, 2000, pp. 805810.
8. Chung, W., Nonholonomic Manipulators, Springer Tracts in Advanced Robotics, vol. 13, Berlin: Springer,
2004.
9. Crossley, V. A., A Literature Review on the Design of Spherical Rolling Robots, Pittsburgh, PA, 2006.
10. Goncharenko, I., Svinin, M., and Hosoe, S., Dynamic Model, Haptic Solution, and Human-inspired Mo-
tion Planning for Rolling-based Manipulation, J. of Computing and Information Science in Engineering,
2009, vol. 9, no. 1, 011004, 10 pp.
11. Kilin, A. A., The Dynamics of Chaplygin Ball: The Qualitative and Computer Analysis, Regul. Chaotic
Dyn., 2001, vol. 6, no. 3, pp. 291306.
12. Michaud, F. and Caron, S., Roball, the Rolling Robot, Autonomous Robots, 2002, vol. 12, pp. 211222.
13. Mukherjee, R., Minor, M. A., and Pukrushpan, J. T., Motion Planning for a Spherical Mobile Robot:
Revisiting the Classical Ball-plate Problem, J. Dyn. Syst. Meas. Control, 2002, vol. 124, pp. 502511.
14. Wilson, J. L., Mazzoleni, A. P., DeJarnette, F. R., Antol, J., Hajos, G. A., and Strickland, C. V., Design,
Analysis, and Testing of Mars Tumbleweed Rover Concepts, J. of Spacecraft and Rockets, 2008, vol. 45,
no. 2, pp. 370382.
REGULAR AND CHAOTIC DYNAMICS Vol. 17 No. 2 2012

You might also like