Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Chapter 6

Lagrangian Dynamics
Having completed our mathematical foundation, we are ready to greatly
extend your abilities in classical mechanics. Newtons equation F = p may
oer the correct description of an object (or a system of objects), however,
in some cases it provides a rather complicated formalism. The formalism
is particularly complicated when there are restrictions or constraints to the
motion of objects. For example when a particle moves in a groove, that may
be of irregular shape, Newtons formalism would require writing expressions
for the forces that conne the particle to the groove. It can be challenging
to nd expressions for such forces (which will vary in time, with the position
of the particle).
In Lagrangian mechanics, we will work with an alternative formalism
based on independent generalized coordinates that completely characterize
the possible motion of the particle. These can greatly simplify problems as
one no longer needs a detailed knowledge of all constraint forces acting in
a system before calculating the dynamics. Lagrangian mechanics is based
on Hamiltons principle, which formulates classical mechanics in a way that
we can apply the calculus of variations in the form of the Euler-Lagrange
equations. Having introduced this formalism, we will consider several problems that would be very dicult with Newtons approach, such as double
pendulums and systems of coupled objects.

6.1

Generalized coordinates

The positions of all particles in a mechanical system are completely described


by a set of generalized coordinates, which we will denote: q1 , . . ., qn . In a
system with nf degrees of freedom and nk constraints, we require at least
ng = nf nk generalized coordinates.
For example, for a particle moving on a sphere of radius a, there is a
constraint on the particle coordinates x2 + y 2 + z 2 = a2 .
In this case, nf = 3
and nk = 1. We could eliminate z by writing z = a2 x2 y 2 and
52

choose x and y as coordinates. Or, we could choose the polar and azimuthal
angles and as generalized coordinates to completely describe the particle
position. Note that generalized coordinates are not unique; we have some
choice with some choices being more convenient than others.
For now we will consider only cases where ng = nf nk . Later we will
consider how to work with any remaining constraints among the chosen q1 ,
. . ., qn .
Generalized coordinates may have units of length, or angle or something
completely dierent! For example, we will see that in the theory of small
oscillations, the generalized coordinates are conventionally chosen to have

units of mass length. Importantly, once a choice of generalized coordinates has been made, with a set of units, the units of the generalized
momentum and force are related as:
{pj } =

mass length2 1
time
{qj }

(6.1)

{Fj } =

mass length2 1
{qj }
time2

(6.2)

where {A} represents the units of A.


For example, in the most familiar case, where qj has units of length, pj
has units of momentum and Fj has units of force. If qj is an angle and is
dimensionless, then pj has units of angular momentum (masslength2 /time)
and Fj has units of torque (mass length2 /time2 ). Note, nally, that there
is no reason why dierent coordinates in the set q1 , . . ., qn need to have the
same units.

6.2

Hamiltons Principle

Hamiltons principle states that the motion of a system from time t1 to t2


is such that the action functional:
t2
t)
S [q(t)] =
L (q, q,
(6.3)
t1

is an extremum. Here, q = q1 , . . . , qn is a complete set of generalized coordinates and


L=T U
(6.4)
is the so-called Lagrangian, where T and U are the kinetic energy and potential energy, respectively. Note that the statement of an extremum does not
necessarily mean a minimum, although in almost all applications in classical
mechanics it is a minimum that occurs.
Hamiltons principle should not be taken as itself a new physical theory.
It would be more correct to say that it is more fundamental that Newtons equations. This is because Hamiltons principle can be applied to a
53

much wider range of physical phenomena, for example in quantum mechanics, electromagnetism and relativity. Hamiltons principle gives a formalism
equivalent to Newtons equations of motion for classical mechanics problems.
Setting S = 0 allows us to apply the Euler-Lagrange equations:
(
)
L
d L
=
(6.5)
dt qj
qj
Note that the quantity L/ qj = pj can be identied as the generalized
momentum and L/qj = Fj , on the right, is the generalized force. The
above equation is equivalent to Newtons second law, p j = Fj . However,
let us recall that qj is a generalized coordinate, so pj does not necessarily
have units of momentum and neither should Fj necessarily have units of
force. For example, if the generalized coordinate is an angle, , then the
corresponding generalized momentum is the angular momentum about the
axis of and the generalized force is the torque.

6.2.1

Invariance of the equations of motion

An important property of the Lagrangian is that it is not a unique quantity


and can be shifted giving no change to the equations of motion. Let us
consider the shifted Lagrangian:
d
(q, q,
L
t) = L (q, q,
t) + G (q, t)
dt

(6.6)

where G (q, t) is any function of the generalized coordinates and time. The
dierence in the action for this Lagrangian and the original Lagrangian,
L (q, q,
t), is:
t2
d

S [q(t)] S [q(t)] =
G(q, t)dt = G(q(t2 ), t2 ) G(q(t1 ), t1 )
(6.7)
dt
t1
That is, the dierence only depends on the end-point values of the path
q(t). When minimizing S over the paths q(t) there can be no variation in
the paths at the end-points (q(t1 ) = 0 and q(t2 ) = 0), so S and S are
minimized by the same path. Thus the equations of motion are invariant
under a shift that can be written as dG(q, t)/dt.
(q, q,
This can also be seen by substituting L
t) into the Euler-Lagrange
equation:
(
)
(
)
d L
d
dG
L
dG
+
=
+
(6.8)
dt qj
dt qj dt
qj
qj dt
(
)
d L
d dqj dG
L
dG
+
=
+
(6.9)
dt qj
dt dqj qj dt
qj
qj dt
where the terms involving dG/dt cancel.
54

6.2.2

Lagrangian for a free particle

Let us consider the description of a free particle using a Cartesian coordinate


system. For a single particle, the Lagrangian L(x, v, t) must be a function
only of v2 . That is, in a homogeneous space there can be no dependence
on x and, similarly, homogeneity of time requires no explicit dependence on
time. The isotropy of space requires that there is no sensitivity to direction,
thus we can only have a dependence on v2 .
Next, let us note that Galilean invariance requires that the equations of
motion are invariant under transformation to a reference frame moving with
constant velocity. Let V be the velocity of a new reference frame, in which
x = x Vt and v = v V. In order that the equations of motion are
invariant, we require:
( )
( )
d
(6.10)
L v2 = L v2 + G(x, t)
dt
The function
where, the function G(x, t) does not depend on v = x.
dG(x, t)/dt can thus only give a linear dependence on v, so the only possibility is L = 21 mv2 , where m is a constant. This gives:
(
)
1
d 1
dG
1
2
2
2

mV t mV.x = L +
(6.11)
L = m (v V) = mv +
2
2
dt 2
dt
For a system of N interacting particles,
1
L=
mn v n2 U ({xn }, {xn })
2
N

(6.12)

n=1

where U is the potential energy, which can usually be decomposed:

U=
U1 (xn ) +
U2 (xn xn )
(6.13)
n

n<n

However, it should be noted that velocity-dependent potentials can appear in the case of charged particles interacting with electromagnetic elds.

6.3

Conserved quantities

A conserved quantity A(q, q,


t) is a quantity that does not change throughout
the motion of the system. Mathematically, this statement is written:

dA
=0
(6.14)
dt
q(t)

When studying the Euler-Lagrange equation, we considered two cases where


the Lagrangian did not depend on x and y. In classical mechanics
these cases correspond to momentum conservation, where the Lagrangian
is independent on a generalized coordinate, and energy conservation, where
the Lagrangian is independent of time.
55

6.3.1

Momentum Conservation

If the Lagrangian does not depend on one or more of the generalized coordinates, we have:
L
Fj =
=0
(6.15)
qj
The Euler-Lagrange equations then give pj = Fj , that is, the generalized
momentum pj is conserved.
As an example, we can consider the motion of a particle in a uniform
gravitational eld, such as near the surface of the Earth. Let (x, y) be
coordinates parallel to the surface and let z be the height. The kinetic and
potential energy are:
)
1 (
T = m x 2 + y 2 + z 2
2
U = mgz

(6.16)

The Lagrangian is then:


)
1 (
L = T U = m x 2 + y 2 + z 2 mgz
2
Note that:
Fx =

L
=0
x

Fy =

(6.17)

L
=0
y

(6.18)

L
= my
y

(6.19)

such that px and py are conserved. Furthermore:


px =

L
= mx
x

py =

Integrating, we obtain:
x(t) = x(0) +

px
t
m

y(t) = y(0) +

py
t
m

(6.20)

Now, let us consider the motion in the z direction. In this case the Lagrangian does depend on z so the momentum pz is not conserved, but instead
given by:
( )
d L
L
d
(pz ) =
=
(6.21)
dt
dt z
z
pz = m
z = mg = Fz
(6.22)
Integrating:
z(t)
= z(0)

gt

(6.23)

1
z(t) = z(0) + z(0)t

gt2
2

(6.24)

56

As another example, let us consider a particle moving in the (x, y) plane


under the inuence of a cylindrically symmetric potential, U (x, y)
= U (),
which depends only on the particles distance from the origin = x2 + y 2 .
Working in polar coordinates (, ), the Lagrangian is:
)
1 (
L = m 2 + 2 2 U ()
(6.25)
2
The Lagrangian does not depend explicitly on , so the quantity p is conserved:
L
p =
= m2
(6.26)

p is the angular momentum about the z-axis.

6.3.2

Energy Conservation

Next, let us consider the case when the Lagrangian does not depend on time.
Consider the function:
t) =
H(q, q,

t)
pj qj L(q, q,

(6.27)

j=1

Taking the total time derivative of H:


)
n (

dH
L
L
L
=
pj qj + pj qj
qj
qj
dt
qj
qj
t

(6.28)

j=1

Note that pj = L/ qj and along the path of motion of the system the
Euler-Lagrange equations give pj = L/qj . Hence the term in brackets
vanishes and we are left with:
dH
L
=
dt
t

(6.29)

That is, H, is conserved if the Lagrangian has no explicit time dependence.


For a Lagrangian describing a set of interacting particles, of the form:
L=

1
j=1

mj r 2j U (r1 , . . . , rn )

(6.30)

the generalized momenta are pj = (L/ x j , L/ y j , L/ zj ) = mj r j . The


quantity H is then:
H=

1
j=1

mj r j + U (r1 , . . . , rn ) = T + U

(6.31)

That is, H is the total energy of the system. Note, however, that it is not
always the case that H = T + U .
57

6.4

Coordinate Systems

Although any choice of generalized coordinates will yield an equivalent set of


equations of motion, it is worth taking some care in the choice of coordinates
as some choices are simpler to work with than others. The form of the
potential energy and any constraints on the motion of the system are often
more easily represented using a particular set of generalized coordinates.
Three common systems are the Cartesian, cylindrical and spherical polar
coordinate systems.
Cartesian Coordinates
The kinetic energy is always simple to write in Cartesian coordinates. Even
if another coordinate system will be used, it can be helpful to write the kinetic energy rst in Cartesian coordinates and then transform to the desired
coordinates. For a single particle, of mass m:
)
1 (
T = m x 2 + y 2 + z 2
2

(6.32)

(
)
If the motion is conned to a plane, one can use T = 12 m x 2 + y 2 .
Cylindrical coordinates
In cylindrical coordinates (, , z), is the radial coordinate in the (x, y)
plane, z is the height above the (x, y) plane, and is the azimuthal angle:
x = cos
y = sin

x = cos sin
y = sin + cos

(6.33)
(6.34)

The kinetic energy is:


)
1 (
T = m 2 + 2 2 + z 2
2

(6.35)

(
)
Again, if the motion is conned to a plane, one can use T = 12 m 2 + 2 2 .
Spherical coordinates
In spherical coordinates (r, , ), r is the radius, is the polar angle, and
is the azimuthal angle. On the globe, one conventionally chooses = 0 to
correspond to the North pole, and = /2 to correspond to points on the
equator. In spherical polar coordinates:

y = r sin sin

x = r sin cos + r cos cos r sin sin


y = r sin sin + r cos sin + r sin cos

z = r cos

z = r cos r sin

x = r sin cos

58

(6.36)

The kinetic energy is:


)
1 (
T = m r 2 + r2 2 + r2 2 sin2
2

6.5

(6.37)

Examples

Now let us consider some example problems, to demonstrate how the Lagrangian formulation of classical mechanics can be used. Let us rst recall
the recipe for solving mechanics problems with this method:
1. Choose a convenient set of generalized coordinates, {q1 , . . . , qn }.
2. Find the kinetic energy T (q, q,
t), the potential energy U (q, q,
t) and
the Lagrangian L(q, q,
t) = T U .
3. Find the generalized momenta pj = L/ qj and the generalized forces
Fj = L/qj .
4. Use the Euler-Lagrange equations to write the equations of motion.
5. Identify any conserved quantities.

6.5.1

One-dimensional motion

Consider the motion of a particle in a 1D system with potential energy U (x).


The Lagrangian is:
1
L = T U = mx 2 U (x)
(6.38)
2
The generalized momentum is:
p=

L
= mx
x

(6.39)

The Euler-Lagrange equation is:


( )
d L
L
=
dt x
x

m
x = U (x)

(6.40)
(6.41)

which is nothing other than the statement F = ma, where a is the acceleration.
The Lagrangian does not explicitly depend on t, so the quantity H should
be conserved:
1
H = px L = mx 2 + U (x)
(6.42)
2

59

That is the total energy, H = T + U is conserved. In this example, this can


also be seen by multiplying the equation of motion by x:

(
)
0 = x m
x + U (x)
(6.43)
(
)
dx
dx
dt dU (x)
=
m
+
(6.44)
dt
dt
dx dt
(
)
d 1
2
mx + U (x)
(6.45)
=
dt 2

6.5.2

Central force in two dimensions

Let us now consider a particle moving in two dimensions subject toa potential U (), which depends only on the distance from the origin = x2 + y 2
In cylindrical coordinates, the Lagrangian is:
)
1 (
L = m 2 + 2 2 U ()
(6.46)
2
The Euler-Lagrange equations give the equations of motion:
( )
d L
L
=

m
= m 2 U ()
(6.47)
dt p

( )
L
d ( 2 )
d L
=

m = 0
(6.48)
dt

dt
Note that the angular momentum is conserved:
p = m2
(6.49)
This relation allows the elimination of from the rst Euler-Lagrange equation, leaving:
p2
m
=
U ()
(6.50)
m3
We can expect the quantity H to be conserved, due to time independence
of the Lagrangian:
H = p + p L(, )
(6.51)
L
L
+
L(, )


)
1 (
= m 2 + m2 2 m 2 + 2 2 + U ()
2
=T +U
=

(6.52)
(6.53)
(6.54)

That is, the total energy H is conserved. Note that the total energy can
also be written:
)
1 (
H = m 2 + 2 2 + U ()
(6.55)
2
p2
1
= m 2 +
+ U ()
(6.56)
2
2m2
60

Dierentiating with respect to time gives:


(
)
p2
dH

= m

+ U () = 0
dt
m3

6.5.3

(6.57)

A sliding mass on a sliding wedge

Consider now a triangular shaped wedge, characterized by angle and mass


M , that slides along a horizontal surface without friction. A point object,
characterized by mass m slides on top of the wedge, also without friction.
What is the consequent motion? Let us choose as generalized coordinates

m
M

0
0

X
x

Figure 6.1: A sliding mass on a sliding wedge.


the horizontal position X of the left corner of the wedge and the horizontal
distance x of the point object from the left corner of the wedge. The vertical coordinate of the point object is then y = x tan , where we take the
horizontal surface as lying at y = 0.
The kinetic energy is:
)2 1
1
1 (
T = M X 2 + m X + x + my 2
(6.58)
2
2
2
)
1 (
1
(6.59)
T = (M + m) X 2 + mX x + m 1 + tan2 x 2
2
2
The potential energy is:
U = mgy = mgx tan

(6.60)

)
1
1 (
(M + m) X 2 + mX x + m 1 + tan2 x 2 mgx tan
2
2

(6.61)

The Lagrangian is then:


L=

61

Using the Euler-Lagrange equations, the equations of motion are:


(
)
d L
L
+ m
=

(M + m) X
x=0
(6.62)
dt X
X
( )
(
)
d L
L
+ m 1 + tan2 x
=

MX
= mg tan (6.63)
dt x
x
The rst equation gives:

m
x

M +m
Substituting into the second equation gives:
=
X

(6.64)

(M + m) g sin cos
M + m sin2

(6.65)

x
=
and

= mg sin cos
X
(6.66)
M + m sin2
That is, both objects travel with constant acceleration, which can be expected since the gravitational force that drives the system is constant.

6.5.4

The double pendulum

As promised, Lagrangian dynamics also oers an ecient approach to the


description of double pendulums. In a double pendulum, a rst pendulum
hangs from a xed point and is characterized by a length l1 and a bob with
mass m1 . A second pendulum hangs from the mass m1 and is characterized
by a length l2 and a bob with mass m2 . As is usually the case with pendulums, it will be convenient to choose angular coordinates, 1 and 2 , which
describe the angles that each pendulum makes with the vertical direction.
Let us rst write the kinetic and potential energies in Cartesian coordinates:
(
) 1
(
)
1
T = m1 x 21 + y 12 + m2 x 22 + y 22
2
2
U = m1 gy1 + m2 gy2

(6.67)

To convert to angular coordinates, we use:


x1 = l1 sin 1
y1 = l1 cos 1

x2 = l1 sin 1 + l2 sin 2

(6.68)

y2 = l1 cos 1 l2 cos 2

(6.69)

The corresponding velocities are obtained by dierentiating with respect to


time:
x 1 = l1 1 cos 1
y 1 = l1 1 sin 1

x 2 = l1 1 cos 1 + l2 2 cos 2
y 2 = l1 1 sin 1 + l2 2 sin 2
62

(6.70)
(6.71)

L1
1

m1

L2

m2
Figure 6.2: The double pendulum.
We can now re-write the kinetic and potential energies in the chosen generalized coordinates:
(
)
1
1
T = m1 l12 12 + m2 l12 12 + 2l1 l2 (cos 1 cos 2 + sin 1 sin 2 ) 1 2 + l22 22
2
2
U = m1 gl1 cos 1 m2 gl1 cos 1 m2 gl2 cos 2
(6.72)
Noting the trigonometric identity cos 1 cos 2 + sin 1 sin 2 = cos (1 2 ),
the Lagrangian can be written:
L=T U =

1
1
(m1 + m2 ) l12 12 + m2 l1 l2 cos (1 2 ) 1 2 + m2 l22 22
2
2
+ (m1 + m2 ) gl1 cos 1 + m2 gl2 cos 2
(6.73)

The generalized momenta are:


L
= (m1 + m2 ) l12 1 + m2 l1 l2 cos (1 2 ) 2

1
L
p2 =
= m2 l1 l2 cos (1 2 ) 1 + m2 l22 2
2

p1 =

63

(6.74)
(6.75)

and the equations of motion are:


p1 = (m1 + m2 ) l12 1 + m2 l1 l2 cos (1 2 ) 2
(
)
m2 l1 l2 sin (1 2 ) 1 2 2
=
and

L
= m2 l1 l2 sin (1 2 ) 1 2 (m1 + m2 ) gl1 sin 1
1

(6.76)

)
(
p2 = m2 l1 l2 cos (1 2 ) 1 m2 l1 l2 sin (1 2 ) 1 2 1 + m2 l22 2

L
= m2 l1 l2 sin (1 2 ) 1 2 m2 gl2 sin 2
2

(6.77)

Re-writing:
l1 1 +

m2 l 2
m2 l 2
cos (1 2 ) 2 +
sin (1 2 ) 22 + g sin 1 = 0
(m1 + m2 )
(m1 + m2 )
(6.78)
l1 cos (1 2 ) 1 l1 sin (1 2 ) 2 + l2 2 + g sin 2 = 0
1

(6.79)
An analytical solution to the equations can be obtained in the limit of small
oscillations, where: cos 1,2 1, sin 1,2 1,2 and we neglect terms of
higher than rst order in 1 and 2
1 + 2 + 02 1 = 0
1 + 2 + 02 2 = 0

(6.80)
(6.81)

where we have dened:


=

m2
m1 + m2

l2
l1

02 =

g
l1

Now, let us search for an oscillating trial solution of the form:


(
) ( )
1
a
=
eit
2
b

(6.82)

(6.83)

where a and b are constant amplitudes. The equations of motion can then
be written in matrix form:
)( )
(
a
2 + 02
2
=0
(6.84)
b
2
2 + 02
For oscillations of non-zero amplitude, the determinant of the above matrix
vanishes. This gives a quadratic equation in 2 :
( 2
)(
)
02 2 02 4 = 0
(6.85)
64

The solution is:


2

02 (1 + ) 02

(1 )2 + 4

2 (1 )

(6.86)

That is, there are two normal mode solutions, corresponding to two possible
oscillation frequencies. Let us consider the case of equal pendulums with
l1 = l2 and m1 = m2 . In this case = 1/2 and = 1. The mode energies
are given by:
(
)
2 = 02 2 2
(6.87)
Substitution back into the matrix equation gives:
)
) )( )
(
( (
a
2( 2 +
1 ( 12 2
)2
2
)
0
=0
b
2 2
2 2 +1

(6.88)

From this equation, we can obtain:

1 2
a

=
b
2 2

(6.89)

with the + sign corresponding to the higher frequency mode. Thus the
two modes correspond to cases where the pendulums oscillate in and out of
phase. The in-phase case corresponds to the lowest frequency.

6.6

Constraints

We have seen that in the formulation of problems in Lagrangian mechanics


we can choose any independent set of generalized coordinates, {q1 , q2 , . . . , qn }.
However, in many cases the coordinates are not independent.
For example, consider the case where a cylinder of radius a rolls over a
half-cylinder of radius R. We can parameterize the motion using the angles
1 and 2 , where 1 denes the angle between the cylinder centers (with
1 = 0 corresponding to when the rolling cylinder is directly above the half
cylinder). 2 then denes the angle that a point on the rolling cylinder
surface makes with the downward vertical direction. If there is no slipping
between the cylinders, then the angles 1 and 2 are not independent, and
rather satisfy the constraint:
R1 = a (2 1 )

(6.90)

In this( particular
example we could solve the above equation and substitute
)
R
2 = 1 + a 1 . However, in other cases it may not be straightforward to
solve the equation representing the constraint. In the rest of this section,
well consider how constraints can be handled in the Lagrangian formalism.
65

a( 2 - 1)
2

R 1

Figure 6.3: One cylinder rolling o another. When there is no slipping, the
angles 1 and 2 are constrained by Eq. 6.90.

6.6.1

Holonomic Constraints

Holonomic Constraints are relationships between the coordinates that can


be written in the form:
f (q, t) = 0
(6.91)
with an index of the constraint (in some cases there may be multiple constraints to handle).
To handle this in the Lagrangian formalism we introduce new variables
called Lagrange multipliers and dene a new Lagrangian:
L (q, q,
t, ) = L(q, q,
t) + f (q, t)

(6.92)

We treat like new coordinates and consequently we can apply the EulerLagrange equations:
(
)
d L
L

=0
(6.93)
dt

Since L does not depend on , the Euler-Lagrange equations give:


L
= f (q, t) = 0

(6.94)

That is, the use of Lagrange multipliers acting as eective coordinates encodes the constraints into the problem. Meanwhile, the Euler-Lagrange
equations for q are:
(
)
d L
L

=0
(6.95)
dt q
q
( )
L
f
d L

=
(6.96)
dt q
q
q
The left-hand side is the equation of motion for the unconstrained problem.
The right-hand side contains the eect of the constraint forces in the system.
66

6.6.2

One cylinder rolling o another

Let us return to the example of one cylinder rolling of another. There are two
constraints in the system. First, we have the constraint that the cylinders
are in contact:
f1 = r R a = 0
(6.97)
Second, there is the no slipping constraint:
f2 = R1 a (2 1 ) = 0
The Lagrangian can be written:
) 1
1 (
L = T U = M r 2 + r2 12 + I 22 M gr cos 1
2
2

(6.98)

(6.99)

where M is the mass of the rolling cylinder. The kinetic energy contains
a contribution from both the center-of-mass movement of the cylinder and
the rotation about the center-of-mass. I is a constant representing the rotational inertia.
To account for the constraints, let us introduce Lagrange multiplies 1
and 2 . The shifted Lagrangian is:
L = L + 1 f1 + 2 f2

(6.100)

The Euler-Lagrange equations in 1 and 2 will give us back the constraint


equations. The other Euler-Lagrange equations are:
(
)
L
d L

= M r M r12 + M g cos 1 1 = 0
dt r
r
(
)
d L
L

= M r2 1 + 2M rr 1 M gr sin 1 2 (R + a) = 0
dt 1
1
(
)
d L
L

= I 2 + 2 a = 0
(6.101)
dt 2
2
Let us now implement the constraints, which give r = (R + a) a constant,

that is, r = 0, and 2 = R+a


a 1 . Substitution into the above equations gives:
M (R + a) 12 + M g cos 1 = 1
M (R + a)2 1 M g (R + a) sin 1 = 2 (R + a)
)
(
R+a
1 = 2 a
I
a

(6.102)
(6.103)
(6.104)

From Eq. 6.104 and 6.101 we have:


R+a
I
2 = 2 = 2 I 1
a
a
67

(6.105)

Substituting into Eq. 6.103 and multiplying by 1 gives:


(
)
I
M + 2 (R + a)2 1 1 M g (R + a) sin 1 1 = 0
a

(6.106)

The integral of this equation is:


)
(
1
I
(0)
(R + a)2 12 + M g (R + a) cos 1 = M g (R + a) cos 1
M 1+
2
M a2
(6.107)
(0)
where 1 is a constant of integration. Here we have assumed that 1 = 0
(0)
(0)
when 1 = 1 , that is, the rolling cylinder is released from rest at 1 = 1 .
Re-arranging, we obtain:
)
2M g (
(0)
cos 1 cos 1
M (R + a) 12 =
1+

(6.108)

where we have dened = I/(M a2 ). Note that is a dimensionless parameter and 0 1. Now, substitution into Eq. 6.102 yields:
)
Mg (
(0)
(3 + ) cos 1 2 cos 1
= 1
1+

(6.109)

There is a point at which = 0 and the constraint no longer applied. This


corresponds to the point where the cylinders lose contact and the rolling
(f )
cylinder ies o. This occurs at the angle 1 = 1 , where:
)
(
(0)
2
cos

(f )
1
1 = cos1
(6.110)
3+
(f )

The detachment angle 1 is an increasing function of , meaning that a


larger I delays detachment. This makes sense since the gain in kinetic energy
as the rolling cylinder moves is split between translational and rotational
motion.

68

You might also like