You are on page 1of 36

PC235 Winter 2013 Chapter 8.

Lagrangian Mechanics Slide 1 of 36


Chapter 8.
Lagrangian Mechanics
Joseph Louis Lagrange apologizes for what is about to occur
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 2 of 36
Outline
We are now prepared to discuss a formulation of classical mechanics that was
devised by Joseph Louis Lagrange in the late 1700s. While Lagrangian mechanics
initially seems overly complex and abstract, we shall see that it has two important
advantages over the Newtonian formulation. First, Lagranges equations take the
same form in any coordinate systems. Second, in treating constrained systems, the
Lagrangian approach eliminates the forces of constraint (which are generally both
unknown and not a useful part of the solution anyhow.)
Table of Contents
1
Lagranges Equation for Unconstrained Motion
2
Example of Constrained Systems
3
Constrained Systems in General
4
Proof of Lagranges Equation with Constraints
5
Examples
6
Generalized Momenta and Ignorable Coordinates
7
More about Conservation Laws
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 3 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion JRT 7.1
Consider a particle that moves unconstrained in three dimensions, subject to a
conservative net force F(r). Its kinetic energy is
T =
1
2
mv
2
=
1
2
m(

x
2
+

y
2
+

z
2
), (8.1)
and its potential energy is
U(r) = U(x, y, z). (8.2)
The Lagrangian is dened as
L = T U. (8.3)
That is, it is the KE minus the PE. It is not the same as the total energy E = T + U.
The script L is used to differentiate the Lagrangian from angular momentum L. Note
that L depends on the particles position (x, y, z) and its velocity (

x,

y,

z), i.e.
L = L(x, y, z,

x,

y,

z). Now lets consider two of the derivatives of L,
L
x
=
U
x
= F
x
,
L

x
=
T

x
= m

x = p
x
. (8.4)
In an inertial reference frame, Newtons second law reads F
x
=

p
x
. Differentiating
the second of the above equations and comparing with the rst gives
L
x
=
d
dt
L

x
. (8.5)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 4 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion cont JRT 7.1
We can of course derive corresponding equations in y and z. Thus, we have shown
that Newtons second law implies the following three Lagrange equations in
Cartesian coordinates:
L
x
=
d
dt
L

x
,
L
y
=
d
dt
L

y
,
L
z
=
d
dt
L

z
(8.6)
The path of the particle can be determined either by the Lagrange equations (and
the methods of the previous chapter) or by Newtons second law (chapter 2 of this
course.)
The equations (8.6) have the same form of the Euler-Lagrange equations (7.13 and
7.20) from the last chapter. Therefore, they imply that the integral S =

L dt is
stationary for the path followed by the particle. This is stated by Hamiltons Principle:
The actual path which a particle follows between two points 1 and 2 in a given time
interval t
1
to t
2
, is such that the action integral
S =

t
2
t
1
L dt (8.7)
is stationary when taken along the actual path.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 5 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion cont JRT 7.1
So far, we have proved that for a single particle, the following three statements are
exactly equivalent:
1
A particles path is determined by Newtons second law F = ma.
2
The path is determined by the three Lagranges equations (8.6), at least in
Cartesian coordinates.
3
The path is determined by Hamiltons principle.
Now, we would like to eliminate the condition that Cartesian coordinates r = (x, y, z)
must be used. Consider any set of generalized coordinates (q
1
, q
2
, q
3
). These might
be (r, , ) in spherical polar coordinates, or (, , z) in cylindrical polar coordinates,
for example. It only matters that each position r species a unique value of
(q
1
, q
2
, q
3
) and vice versa. That is,
q
i
= q
i
(r) for i = 1, 2, 3, (8.8)
and
r = r(q
1
, q
2
, q
3
). (8.9)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 6 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion cont JRT 7.1
We can then rewrite (x, y, z) and (

x,

y,

z) in terms of (q
1
, q
2
, q
3
) and (

q
1
,

q
2
,

q
3
), and
we can rewrite L =
1
2
m

r
2
U(r) in terms of these new variables as
L = L(q
1
, q
2
, q
3
,

q
1
,

q
2
,

q
3
) (8.10)
and the action integral as
S =

t
2
t
1
L(q
1
, q
2
, q
3
,

q
1
,

q
2
,

q
3
) dt . (8.11)
The change of variables does not alter the value of S. Therefore, S must still be
stationary for the correct path in these new coordinates, which means that the
correct path must satisfy the Euler-Lagrange equations
L
q
1
=
d
dt
L

q
1
,
L
q
2
=
d
dt
L

q
2
,
L
q
3
=
d
dt
L

q
3
(8.12)
We can now drop the qualication in Cartesian coordinates from statement #2 on
the previous page. The usefulness of the Lagrangian formulation is evident in the
above equations - they are valid for any choice of coordinates.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 7 of 36
Lagranges Equation for Unconstrained Motion
Example #1 - One Particle in Two Dimensions JRT Ex. 7.1
Problem: Write down Lagranges equations in Cartesian coordinates for a particle moving in a conservative force eld
in two dimensions, and show that they imply Newtons second law.
Solution: The Lagrangian for a single particle in 2D is
L(x, y, x, y) = T U =
1
2
m( x
2
+ y
2
) U(x, y). (8.13)
To write down the Lagrange equations we need the derivatives
L
x
=
U
x
= F
x
,
L
x
=
T
x
= m x, (8.14)
with corresponding expressions for the y derivatives. Thus the two Lagrange equations can be rewritten as:
L
x
=
d
dt
L
x
F
x
= m x
L
y
=
d
dt
L
y
F
y
= m y

F = ma. (8.15)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 8 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion cont JRT 7.1
Notice how in eq. (8.14) the derivative L/x is the x component of the force and
L/

x is the x component of the momentum, and similarly for the y components.


When we use the generalized coordinates q
1
, q
2
, we shall nd that L/q
i
,
although not necessarily a force component, plays a role very similar to a force.
Likewise, L/

q
i
plays a role similar to a momentum. They will be called generalized
force and generalized momentum respectively:
L
q
i
= i
th
component of generalized force, (8.16)
L

q
i
= i
th
component of generalized momentum. (8.17)
and thus, the generalized force equals the rate of change of generalized momentum.
This concept is illustrated in the following example.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 9 of 36
Lagranges Equation for Unconstrained Motion
Example #2 - 1 Particle in 2 Dimensions (Polar Coords.) JRT Ex. 7.2
Problem: Write down Lagranges equations for the same system as in the previous example, but using polar
coordinates.
Solution: As always, we start by writing down the Lagrangian, T U. In polar coordinates (see Fig. 1,) we know that the
velocity components are v
r
= r and v

= r

, and thus the KE is
T =
1
2
mv
2
=
1
2
mv v =
1
2
m( r, r

) ( r, r

) =
1
2
m( r
2
+ r
2

2
). The Lagrangian is
L = L(r, , r,

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

2
) U(r, ). (8.18)
Next, we examine individually the Lagrange equations for coordinates r and .
The r equation
The Lagrange equation for the coordinate r is
L
r
=
d
dt
L
r
. (8.19)
Carrying out the appropriate derivatives of L with respect to r and r, and substituting them in the above Lagrange equation
gives
mr

U
r
=
d
dt
(m r) = mr. (8.20)
We know, however, that U/r is just F
r
, and thus we can have the familiar result - a centripetal force - from chapter 1:
F
r
= m(r r

2
) = ma
r
. (8.21)
Fig. 1: The velocity of a particle expressed in 2D polar coordinates - see JRT Fig. 7.1
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 10 of 36
Lagranges Equation for Unconstrained Motion
Example #2 - 1 Particle in 2 Dimensions (Pol. Coords.) cont
JRT Ex. 7.2
The equation
The Lagrange equation for the coordinate is
L

=
d
dt
L

(8.22)
or, for this particular L,

=
d
dt
(mr
2

) (8.23)
To calculate the left-hand side of this equation, we need to know the components of U in polar coordinates. This can be
determined by the methods discussed in chapter 1:
U =
U
r
r +
1
r
U

(8.24)
This means that
F

=
1
r
U

. (8.25)
The left side of eq. (8.23) is just rF

, which is equal to the torque on the particle about the origin. Meanwhile, the quantity
mr
2

= mr
2
is the angular momentum L about the origin. Therefore, the equation has reproduced the familiar relation
=
dL
dt
. (8.26)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 11 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion -
Several Unconstrained Particles JRT 7.1
The extension of these concepts to a system of N unconstrained particles is fairly
straightforward. For N = 2, the Lagrangian becomes
L(r
1
, r
2
,

r
1
,

r
2
) =
1
2
m
1

r
2
1
+
1
2
m
2

r
2
2
U(r
1
, r
2
). (8.27)
The forces on the two particles are F
1
=
1
U and F
2
=
2
U. Newtons second
law can be applied to each particle, yielding the six equations
F
1x
=

p
1x
, F
1y
=

p
1y
, , F
2z
=

p
2z
. (8.28)
As before, each of these six equations is equivalent to a corresponding Lagrange
equation
L
x
1
=
d
dt
L

x
1
,
L
y
1
=
d
dt
L

y
1
, ,
L
z
2
=
d
dt
L

z
2
. (8.29)
These six equations imply that the integral S =

t
2
t
1
L dt is stationary. Finally, we can
change to any other suitable set of six coordinates q
1
, q
2
, , q
6
. Lagranges
equations must be true with respect to the new coordinates:
L
q
1
=
d
dt
L

q
1
,
L
q
2
=
d
dt
L

q
2
, ,
L
q
6
=
d
dt
L

q
6
. (8.30)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 12 of 36
Lagranges Equation for Unconstrained Motion
Lagranges Equation for Unconstrained Motion -
Several Unconstrained Particles JRT 7.1
One example of a set of six generalized coordinates will be used repeatedly in an
upcoming chapter, which deals with two-body central-force problems. In place of the
six coordinates of r
1
and r
2
, we shall use the three components of the center of
mass position R = (m
1
r
1
+ m
2
r
2
)/(m
1
+ m
2
) and the three coordinates of the
relative position r = r
1
r
2
.
In the case of N unconstrained particles, there are 3N Lagrange equations:
L
q
i
=
d
dt
L

q
i
, i = 1, 2, , 3N. (8.31)
You might be asked to prove this in an upcoming assignment.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 13 of 36
Example of Constrained Systems
Example of Constrained Systems JRT 7.2
Consider the simple pendulum shown in Fig. 2. A bob of mass m is xed to a
massless rod, which pivots about O and is free to swing without friction in the
xy-plane. The constraint is that

x
2
+ y
2
remains constant. Thus, we can
eliminate one of the coordinates x or y; if we know one, we know the other. A
simpler method is to use the single parameter , the angle between the pendulum
and its equilibrium position (hanging vertically down.)
Fig. 2: A simple pendulum. The position of the mass m can be specied by the single coordinate - see JRT Fig. 7.2
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 14 of 36
Example of Constrained Systems
Example of Constrained Systems cont JRT 7.2
We can express all of the quantities of interest in terms of . Since v =

for
circular motion, KE =
1
2
mv
2
=
1
2
m
2

2
. The potential energy is
U = mgh = mg(1 cos ). Thus, the Lagrangian is
L = L(,

) = T U =
1
2
m
2

2
mg(1 cos ). (8.32)
Now, for a single particle we have (for a generalized coordinate q and our specic
coordinate )
L
q
=
d
dt
L

q
,
L

=
d
dt
L

. (8.33)
For our Lagrangian, eq. (8.32), we have
mg sin =
d
dt
(m
2

) = m
2

. (8.34)
Referring back to the gure, we see that the left side of this equation is just the
torque exerted by gravity on the pendulum, while the term m
2
is the pendulums
moment of inertia about the pivot (point mass m, a distance from the pivot.) Since

is the angular acceleration , we have shown that = I. You will recall from
PC131 that this is the rotational form of Newtons second law.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 15 of 36
Constrained Systems in General
Constrained Systems in General JRT 7.3
Consider an arbitrary system of N particles, labelled = 1, , N, with positions r

.
We say that the parameters q
1
, , q
n
are a set of generalized coordinates for the
system if each position r

can be expressed as a function of q


1
, , q
n
, and possibly
the time t :
r

= r

(q
1
, , q
n
, t ) [ = 1, , N] , (8.35)
and conversely, each q
i
can be expressed in terms of the r

and possibly t :
q
i
= q
i
(r
1
, , r
N
, t ) [i = 1, , n] . (8.36)
In addition, we require that the number of generalized coordinates n is the smallest
number that allows the system to be parametrized in this way. In a three-dimensional
world, this number is no more than 3N, but for a constrained system, it is often much
less. For example, a rigid body may have N 10
23
, whereas n = 6 (three
coordinates to specify the CM, 3 coordinates to specify the orientation; all of the
other coordinates specify inter-particle distances, which dont change in a rigid body,
and thus dont impact the Lagrangian.)
In the case of the pendulum mentioned previously, we have
r = (x, y) = ( sin , cos ) (8.37)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 16 of 36
Constrained Systems in General
Constrained Systems in General cont JRT 7.3
A slightly more complex system is the double pendulum, shown in Fig. 3. This
system has two bobs, both conned to a plane, so it has four Cartesian coordinates,
which can be represented in terms of two generalized coordinates
1
and
2
. If we
put our origin at the suspension point of the top pendulum, then we have
r
1
= (
1
sin
1
,
1
cos
1
) = r
1
(
1
)
r
2
= (
1
sin
1
+
2
sin
2
,
1
cos
1
+
2
cos
2
) = r
2
(
1
,
2
).
(8.38)
Notice that the components of r
2
depend on both of the generalized coordinates
1
and
2
. If the relation between the Cartesian coordinates r

and the generalized


coordinates q
i
does not involve the time t - as with the double pendulum - then the
generalized coordinates are said to be natural; later on, we will see that natural
coordinates allow certain simplifying approximations.
Fig. 3: The position of both masses in a double pendulum
are uniquely specied by the two generalized coordinates
1
and
2
- see JRT Fig. 7.3
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 17 of 36
Constrained Systems in General
Constrained Systems in General - Degrees of Freedom JRT 7.3
The number of degrees of freedom (DOF) of a system is the number of coordinates
that can be independently varied in a small displacement. The simple pendulum has
one DOF and the double pendulum has two. A particle that is free to move anywhere
has three DOF, and a gas comprised of N particles has 3N DOF.
When the number of DOF of an N-particle system in k dimensions is less than kN,
we say that the system is constrained. The bead on a wire and the rigid body are
constrained systems.
In the previous examples, the number of DOF was equal to the number of
generalized coordinates needed to specify the systems conguration. A system with
this property is said to be holonomic - a holonomic system has n DOF and can be
described by n generalized coordinates, q
1
, , q
n
. Nonholonomic systems are
much more difcult to describe than holonomic systems, and will not be discussed in
this course. For an example of a nonholonomic system, see pp. 249-250 of the text.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 18 of 36
Proof of Lagranges Equation with Constraints
Proof of Lagranges Equation with Constraints JRT 7.4
We will consider the case of a single particle that is constrained to move on a
surface (an ant crawling on a basketball, for example.) Because of the constraint,
there are two degrees of freedom, and we can describe the motion of the particle by
two generalized coordinates q
1
and q
2
, which can vary independently.
There are two kinds of forces on the particle. First, there are the forces of constraint.
For atoms in a rigid body, these are the interatomic forces. In general, forces of
constraint are not conservative, but this wont matter, as we shall see. We denote
the net constraining force as F
cstr
. Next, there are all of the other nonconstraint
forces on the particle (such as gravity.) We will denote their net contribution as F.
We shall assume that the nonconstraint forces satisfy at least the second condition
for conservativism (that they are derivable from a potential energy):
F = U(r, t ). (8.39)
(Note that if all of the nonconstraint forces are conservative, then U is independent
of t , but we dont need to make this assumption right now.) The total force on the
particle is F
tot
= F
cstr
+F, and the Lagrangian, as usual, is L = T U. Since U is the
potential energy for the nonconstraint forces only, this denition of L excludes the
constraint forces.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 19 of 36
Proof of Lagranges Equation with Constraints
Proof of Lagranges Equation with Constraints -
Stationarity of Action Integral JRT 7.4
Consider any two points r
1
and r
2
, through which the particle passes at times t
1
and
t
2
. If r(t ) is the correct path that the particle will follow, then we can denote a
neighbouring incorrect path by
R(t ) = r(t ) + (t ), (8.40)
which denes (t ) as the innitesimal vector pointing from r(t ) to R(t ). It is logical to
assume that both r(t ) and R(t ) lie in the surface to which the particle is conned,
and therefore so does (t ). Furthermore, since the endpoints of the path are fully
specied, (t )=0 at t
1
and t
2
.
Now, we denote by S the action integral taken along the incorrect path:
S =

t
2
t
1
L(R,

R, t )dt (8.41)
and by S
0
the action integral taken along the correct path:
S
0
=

t
2
t
1
L(r,

r, t )dt . (8.42)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 20 of 36
Proof of Lagranges Equation with Constraints
Proof of Lagranges Equation with Constraints -
Stationarity of Action Integral cont JRT 7.4
We will now prove that S is stationary for variations of the path R(t ) when R(t ) = r(t )
(or equivalently, when (t ) = 0.) Another way to say this is that the difference in the
action integrals
S = S S
0
(8.43)
is zero to rst order in the distance between the paths.
The difference S is the integral of the difference between the Lagrangian on the two
paths,
L = L(R,

R, t ) L(r,

r, t ). (8.44)
If we substitute R(t ) = r(t ) + (t ) and
L(r,

r, t ) = T U =
1
2
m

r
2
U(r, t ), (8.45)
this becomes
L =
1
2
m

r + )
2

r
2

[U(r + , t ) U(r, t )] (8.46)


= m

r U + O(
2
),
where O(
2
) denotes 2nd- and higher-order terms in and
2
. The last step in the
above equation results from the denition of the gradient, f (r + ) f (r) f .
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 21 of 36
Proof of Lagranges Equation with Constraints
Proof of Lagranges Equation with Constraints -
Stationarity of Action Integral cont JRT 7.4
We then nd that
S =

t
2
t
1
L dt =

t
2
t
1
[m

r U] dt . (8.47)
The rst term in the integral can be integrated by parts, leading to
S =

t
2
t
1
[m

r +U] dt . (8.48)
Now, remember that the path r is the correct path, which means that it satises
Newtons second law. Therefore, m

r is just the total force on the particle,


F
tot
= F
cstr
+ F. But we also know that U = F. Cancellation of these terms leaves
S =

t
2
t
1
F
cstr
dt . (8.49)
However, the constraint force F
cstr
must be normal to the surface in which the particle
moves, while lies in the surface. Thus, their dot product is zero, and we have
proved that S = 0; that is, the action integral is stationary at the correct path.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 22 of 36
Examples
Examples JRT 7.5
Chapter 7.5 of the text presents ve examples of Lagranges equation in action.
They will not all be presented here, but you should read all of them very carefully, as
they explain in explicit detail the procedure for solving problems using principles of
Lagrangian mechanics.
The rst two examples are fairly easily analyzed using Newtons second law, which
should increase our condence in the Lagrangian approach.
Example 7.3 analyzes the two-mass Atwood machine that we encountered
previously. The Newtonian solution was simple enough that Lagrangian methods
dont noticeably shorten the solution, but it should be mentioned that the Lagrangian
approach allows us to completely neglect the tension in the string (the constraint
force.)
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 23 of 36
Examples
Example #3 - A Particle Conned to Move on a Cylinder JRT Ex. 7.4
Problem: Consider a particle of mass m constrained to move on a frictionless cylinder of radius R, specied by the
equation = R in cylindrical polar coordinates (, , z), as shown in Fig. 4. In addition to the force of constraint (the
normal force of the cylinder) the only force on the mass is a force F = kr directed to the origin. Using z and as
generalized coordinates - remember, is constant - nd the Lagrangian L, and use it to solve for the motion of the
mass.
Solution: The system has two degrees of freedom, the z and coordinates of position, which can vary independently. The
velocity in cylindrical coordinates for xed = R is (v

, v

, v
z
) = (0, R

, z). Therefore, the kinetic energy is
T =
1
2
mv
2
=
1
2
mv v =
1
2
m(R
2

2
+ z
2
). (8.50)
Fig. 4: A mass m, conned to the surface of the cylinder = R and subject to a force F = kr - see JRT Fig. 7.7
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 24 of 36
Examples
Example #3 - A Particle Conned to Move on a Cylinder cont
JRT Ex. 7.4
The potential energy for the force F = kr is U =
1
2
kr
2
(this is a Hookes law force) where r is the distance from the origin to
the particle, given by r
2
= R
2
+ z
2
. Therefore, U =
1
2
k(R
2
+z
2
) and
L = L(,

, z, z) =
1
2
m(R
2

2
+ z
2
)
1
2
k(R
2
+z
2
). (8.51)
With two DOF, this system has two equations of motion. The two Lagrange equations read
L
z
=
d
dt
L
z
, or kz = m z,
L

=
d
dt
L

, or 0 =
d
dt
mR
2

. (8.52)
The z equation tells us that the mass executes simple harmonic motion in the z direction, with the usual =

k/m. The
equation tells us that the mass moves around the cylinder with constant angular velocity

. The particles motion would be
completely specied if we had initial conditions for z, , z, and

.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 25 of 36
Examples
Examples cont JRT 7.5
These previous examples illustrated the procedure by which you can solve problems
by the Lagrangian method. Provided that the constraints are holonomic and that the
nonconstraint forces are derivable from a potential energy, the procedure is:
1
Write down the kinetic and potential energies and the Lagrangian
L = T U, using any convenient inertial reference frame.
2
Choose a convenient set of n generalized coordinates q
1
, , q
n
and nd
expressions for the original coordinates of step 1 in terms of these
generalized coordinates.
3
Rewrite L in terms of q
1
, , q
n
and

q
1
, ,

q
n
.
4
Write down the Lagrange equations for each of your n generalized
coordinates.
5
Solve each of the Lagrange equations individually.
We cant guarantee that all of the Lagrange equations are easy to solve, but having
them in this form usually facilitates understanding of the system under consideration.
Next, we will take a look at a couple of examples for which the Lagrangian approach
simplies matters considerably.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 26 of 36
Examples
Example #4 - The Yo-Yo JRT Prob. 7.14
Problem: The gure below shows a crude model of a yo-yo. A massless string is suspended vertically from a xed
point and the other end is wrapped several times around a uniform cylinder of mass m and radius R. When the
cylinder is released it moves downward, rotating as the string unwinds. Using the distance x as the generalized
coordinate, show that the downward acceleration of the cylinder is x = 2g/3. For a cylinder, the kinetic energy of
rotation is
1
2
I
2
, where is the angular velocity about the center of mass, and the moment of inertia I is
1
2
mR
2
.
Solution: First, note that the downward velocity of the cylinder must equal the velocity of a point on the circumference of the
cylinder. This tells us that x = R.
The total kinetic energy of the yo-yo has two components; that of the center-of-mass motion and that of the rotational motion.
Combined, we have
T =
1
2
m x
2
+
1
2
I
2
=
1
2
m x
2
+
1
4
m(R)
2
=
3
4
m x
2
. (8.53)
Taking +x as the downward direction, we have a potential energy U = mgx. Thus, the Lagrangian is
L =
3
4
m x
2
+mgx, (8.54)
which has just the one generalized coordinate, x. The Lagrange equation reads
L
x
=
d
dt
L
x
mg =
3
2
m x, (8.55)
which rearranges to x = 2g/3, as required.
- note that it does not depend on m or R.
Fig. 5: Geometry for this example - see JRT Fig. 7.12
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 27 of 36
Examples
Example #5 - Rolling Downhill JRT Prob. 7.16
Problem: Write down the Lagrangian for an object (mass m, radius R, moment of inertia I) that rolls without slipping
straight down an inclined plane which is at an angle from the horizontal. Use as your generalized coordinate the
objects distance x measured down the plane from its starting point. Write down the Lagrange equation and solve it
for the objects acceleration x. In a race between a solid sphere and a solid cylinder (with identical m and R), which
object would win?
Solution: Since = v/R, the rooling objects KE is T =
1
2
mv
2
+
1
2
I
2
=
1
2

m + I/R
2

x
2
. The PE is U = mgx sin. Thus,
the Lagrangian is
L =
1
2

m + I/R
2

x
2
+mgx sin (8.56)
and the Lagrange equation is
mg sin =

m +I/R
2

x. (8.57)
Therefore,
x =
mg sin
m +I/R
2
, (8.58)
which is a result that you may remember from PC131.
The downhill acceleration will be greater for the object with the smaller moment of inertia. Since
I
sph
=
2
5
mR
2
and I
cyl
=
1
2
mR
2
, (8.59)
the sphere will win the race.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 28 of 36
Examples
Example #6 - Sliding Down a Moving Plane
Problem: A block of mass m is held motionless on a frictionless plane of mass M and angle of inclination . The plane
rests on a frictionless horizontal surface. The block is released. What is the horizontal acceleration of the plane?
Solution: Let x
1
be the horizontal coordinate of the plane (with positive x
1
to the left), and let x
2
be the horizontal coordinate of
the block (with positive x
2
to the right). The relative horizontal distance between the plane and the block is x
1
+ x
2
, so the
height fallen by the block is (x
1
+x
2
) tan. The Lagrangian is therefore
L =
1
2
M x
2
1
+
1
2
m

x
2
2
+ ( x
1
+ x
2
)
2
tan
2

+mg(x
1
+x
2
) tan. (8.60)
The resulting Lagrange equations in the generalized coordinates x
1
and x
2
are
M x
1
+ m( x
1
+ x
2
) tan
2
= mg tan
m x
2
+ m( x
1
+ x
2
) tan
2
= mg tan
(8.61)
We can immediately see that the difference of these two equations yields the required conservation of momentum:
M x
1
m x
2
= 0 (d/dt )(M x
1
mx
2
) = 0. (8.62)
Fig. 6: Geometry for this example
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 29 of 36
Examples
Example #6 - Sliding Down a Moving Plane cont
Eqns. (8.61) are coupled linear equations in the two unknowns x
1
and x
2
, which we can solve to nd the acceleration of the
plane:
x
1
=
mg sincos
M + msin
2

. (8.63)
This is a good time to look at limiting cases. When = 0, x
1
= 0 (makes sense, nothing will move in this case). When m = 0,
x
1
= 0 (theres no component of normal force to push the plane). When m M, x
1
= g/ tan; this also makes sense, since m
falls essentially straight down with acceleration g, and the 1/ tan factor is the ratio of leftward motion to downward motion
caused by the plane angle.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 30 of 36
Examples
Example #7 - Spring Pendulum
Problem: Consider a pendulum in which the string is a massless, Hookes-law spring (that is, it can be stretched and
compressed, but only along its length; it does not buckle). The position of the mass m can be specied by the two
generalized coordinates x and , where x is the radial displacement of the spring from its equilibirum length . Here,
the period of oscillations is perturbed by the ever-changing value of x, while x is perturbed by the centripetal force,
which varies with . We wish to nd (but not solve) the equations of motion in x and .
Solution: The kinetic energy, as usual, is T =
1
2
mv
2
. However, in this case, the velocity of m has both radial and tangential
components. Recalling that the tangential component is v

= r, where r = ( + x) is the instantaneous radius of curvature


and =

is the angular velocity, we can write
v
2
= v v = (v
r
, v

) (v
r
, v

) = ( x, ( +x)

) ( x, ( + x)

) = x
2
+ ( +x)
2

2
, (8.64)
and therefore, T =
1
2
m( x
2
+ ( + x)
2

2
).
As for the potential energy, there are both gravitational and spring components. Using the position of the support as the
reference position for U
g
, we have
U = mg( +x) cos +
1
2
kx
2
. (8.65)
Fig. 7: Geometry for this example
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 31 of 36
Examples
Example #7 - Spring Pendulum cont
Therefore,
L = T U =
1
2
m( x
2
+ ( +x)
2

2
) +mg( +x) cos
1
2
kx
2
. (8.66)
The Lagrange equation in x reads
L
x
=
d
dt
L
x
m( +x)

2
+ mg cos kx = m x. (8.67)
We can rearrange this to the form
mg cos kx = m( x ( + x)

2
). (8.68)
I claim now that this is nothing more than Newtons second law in the radial direction, F
r
= ma
r
. The four terms in the
preceding equation represent the radial component of the gravitational force, the spring force (always radial), the pure radial
acceleration, and the centripetal acceleration (r
2
).
Finally, we examine the Lagrange equation in :
L

=
d
dt
L

mg( + x) sin = m( +x)


2

2
+ 2m( +x) x

, (8.69)
which we will rearrange to
mg sin = m

( +x)

2 x

. (8.70)
This is Newtons second law in the tangential direction, F

= ma

. The two terms on the right-hand side represent the


angular acceleration and the Coriolis acceleration, which we wont encounter until later in the course - it occurs when an
object moves in a rotating frame, which is exactly the case here.
We cant proceed any further with this problem, since our Lagrange equations are coupled and nonlinear. In principle, they
could be solved using perturbation approaches, but one would normally use numerical techniques.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 32 of 36
Generalized Momenta and Ignorable Coordinates
Generalized Momenta and Ignorable Coordinates JRT 7.6
Previously, it was mentioned that for any system with n generalized coordinates q
i
,
we refer to the n quantities L/q
i
= F
i
as generalized forces and L/

q
i
= p
i
as
generalized momenta. With this relation, the Lagrange equation
L
q
i
=
d
dt
L

q
i
(8.71)
can be rewritten as F
i
=
d
dt
p
i
. That is, the generalized force equals the rate of
change of generalized momentum. In particular, if the Lagrangian is independent of
a particular coordinate q
i
, then F
i
=
L
q
i
= 0, and the corresponding generalized
momentum p
i
is constant.
Consider a single projectile moving in the vertical direction, subject only to gravity.
The potential energy (measuring z vertically up) is U = mgz, and the Lagrangian is
L = L(x, y, z,

x,

y,

z) =
1
2
m(

x
2
+

y
2
+

z
2
) mgz. (8.72)
In this case, the generalized force is just the usual force (L/x = U/x = F
x
,
etc.) and the generalized momentum is just the usual momentum
(L/

x = m

x = p
x
, etc.) Because L is independent of x and y, it follows from
Lagranges equations that the components p
x
and p
y
are constant, as we already
knew.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 33 of 36
Generalized Momenta and Ignorable Coordinates
Generalized Momenta and Ignorable Coordinates cont JRT 7.6
In general, the generalized forces and momenta are not the same as the usual
forces and momenta. For instance, we saw in a previous example that in 2D polar
coordinates, the component of the generalized force is the torque, and that of the
generalized momentum is actually the angular momentum. In any case, when the
Lagrangian is independent of a coordinate q
i
, the corresponding generalized
momentum is conserved. When the Lagrangian is independent of a coordinate q
i
,
that coordinate is said to be ignorable or cyclic. It is always a good idea to see if you
can choose a coordinate system in which as many coordinates as possible are
ignorable.
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 34 of 36
More about Conservation Laws
Here, we will see how the laws of conservation of momentum and energy t into the
Lagrangian formulation of mechanics. We already know that these laws hold, since
we derived the Lagrangian formulation from the Newtonian. However, much insight
can be gained from examining these laws strictly from a Lagrangian viewpoint.
Conservation of Total Momentum JRT 7.8
We know from Newtonian mechanics that the total momentum of an isolated system
of N particles is conserved. One of the most prominent features of an isolated
system is that it is translationally invariant. That is, if we transport all N particles
through the same displacement , nothing physically signicant about the system
should change (Fig. 8). The effect of the translation is to replace every position r

by
r

+ .
Fig. 8: An isolated system of N particles is translationally
invariant, which means that when every particle is
transported through the same displacement , nothing
physically signicant changes - see JRT Fig. 7.10
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 35 of 36
More about Conservation Laws
Conservation of Total Momentum cont JRT 7.8
In particular, the potential energy must not be affected by this displacement:
U(r
1
+ , , r
N
+ , t ) = U(r
1
, , r
N
, t ), (8.73)
that is, U = 0. Clearly, the velocities are unchanged by the translation. Therefore,
T = 0 and hence L = 0 under the translation. This is true for any . If we choose
to be an innitesimal displacement in the x direction, then all of the x coordinates
x
1
, , x
N
increase by , while the y and z coordinates are unchanged. For this
translation,
L =
L
x
1
+ +
L
x
N
= 0. (8.74)
Since is arbitrary, this implies that
N

=1
L
x

= 0. (8.75)
But using Lagranges equations, we can rewrite each derivative in the sum as
L
x

=
d
dt
L

=
d
dt
p
x
(8.76)
where p
x
is the x component of the momentum of particle .
PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 36 of 36
More about Conservation Laws
Conservation of Total Momentum cont JRT 7.8
The sum then becomes
N

=1
L
x

=
N

=1
d
dt
p
x
=
d
dt
P
x
= 0 (8.77)
where P
x
is the x component of the total momentum P =

. By choosing the
small displacement successively in the y and z directions, we can prove the same
result for the y and z components, and we conclude that - provided the Lagrangian
is unchanged by the translation - the total momentum of the N-particle system is
conserved.
This connection between a conservation law and invariance of L under certain
transformations (translations, rotations, etc.) is known as Noethers theorem.

You might also like