Notes CH 8

You might also like

Download as pdf or txt
Download as pdf or txt
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

Example of Constrained Systems

Constrained Systems in General

Proof of Lagranges Equation with Constraints

Examples

Generalized Momenta and Ignorable Coordinates

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=
and its potential energy is

1
1
mv 2 = m(x 2 + y 2 + z 2 ),
2
2

(8.1)

U (r) = U (x , y , z ).

(8.2)

L = T U.

(8.3)

The Lagrangian is defined as


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,

U
L
T
L
=
= Fx ,
=
= mx = px .
(8.4)
x
x
x
x
In an inertial reference frame, Newtons second law reads Fx = p x . Differentiating
the second of the above equations and comparing with the first gives
L
d L
=
.
dt x
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
d L
=
,
x
dt x

L
d L
=
,
y
dt y

L
d L
=
z
dt 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
R (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 t1 to t2 , is such that the action integral
Z t2
L dt
S=
(8.7)
t1

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.

The path is determined by the three Lagranges equations (8.6), at least in


Cartesian coordinates.

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 (q1 , q2 , q3 ). These might
be (r , , ) in spherical polar coordinates, or (, , z ) in cylindrical polar coordinates,
for example. It only matters that each position r specifies a unique value of
(q1 , q2 , q3 ) and vice versa. That is,

qi = qi (r) for i = 1, 2, 3,

(8.8)

r = r(q1 , q2 , q3 ).

(8.9)

and

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 (q1 , q2 , q3 ) and (q 1 , q 2 , q 3 ), and


2
we can rewrite L = 21 mr U (r) in terms of these new variables as
L = L(q1 , q2 , q3 , q 1 , q 2 , q 3 )

(8.10)

and the action integral as

S=

t2
t1

L(q1 , q2 , q3 , 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
d L
=
,
q1
dt q 1

L
d L
=
,
q 2
dt q 2

L
d L
=
q 3
dt q 3

(8.12)

We can now drop the qualification 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 field
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
m(x 2 + y 2 ) U (x , y ).
2

(8.13)

To write down the Lagrange equations we need the derivatives


L
U
=
= Fx ,
x
x

L
T
=
= mx ,
x
x

(8.14)

with corresponding expressions for the y derivatives. Thus the two Lagrange equations can be rewritten as:

Fx = mx

F = ma.

L
L
d

F
=
m
y

y
dt y
y

L
d L
= dt
x
x

(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 q1 , q2 , we shall find that L/qi ,
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
= i th component of generalized force,
q i

(8.16)

L
= i th component of generalized momentum.
(8.17)
q i
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
and thus the KE is
velocity components are vr = r and v = r ,
(r , r )
= 1 m(r 2 + r 2 2 ). The Lagrangian is
T = 1 mv 2 = 1 mv v = 1 m(r , r )
2

= T U = 1 m(r 2 + r 2 2 ) U (r , ).
L = L(r , , r , )
2
Next, we examine individually the Lagrange equations for coordinates r and .

The r equation
The Lagrange equation for the coordinate r is

(8.18)

d L
L
=
.
(8.19)
dt r
r
Carrying out the appropriate derivatives of L with respect to r and r , and substituting them in the above Lagrange equation
gives
d
U
mr 2
=
(mr ) = mr .
(8.20)
r
dt
We know, however, that U /r is just Fr , and thus we can have the familiar result - a centripetal force - from chapter 1:
Fr = m(r r 2 ) = mar .
(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

or, for this particular L,

d L
L
=

dt

(8.22)

d
U

(mr 2 )
(8.23)
=
dt

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 =

1 U
U
r +

r
r

(8.24)

This means that

1 U
.
(8.25)
r
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

F =

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
1
1
2
2
m1 r 1 + m2 r 2 U (r1 , r2 ).
(8.27)
2
2
The forces on the two particles are F1 = 1 U and F2 = 2 U . Newtons second
law can be applied to each particle, yielding the six equations
L(r1 , r2 , r 1 , r 2 ) =

F1x = p 1x , F1y = p 1y , , F2z = p 2z .

(8.28)

As before, each of these six equations is equivalent to a corresponding Lagrange


equation
d L L
d L
L
d L
L
,
, ,
.
(8.29)
=
=
=
dt x 1 y1
dt y 1
dt z 2
x1
z2
R t2
These six equations imply that the integral S = t L dt is stationary. Finally, we can
1
change to any other suitable set of six coordinates q1 , q2 , , q6 . Lagranges
equations must be true with respect to the new coordinates:
L
d L L
d L
d L
L
=
=
=
,
, ,
.
dt q 1 q2
dt q 2
dt q 6
q1
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 r1 and r2 , we shall use the three components of the center of
mass position R = (m1 r1 + m2 r2 )/(m1 + m2 ) and the three coordinates of the
relative position r = r1 r2 .
In the case of N unconstrained particles, there are 3N Lagrange equations:
L
d L
=
,
q i
dt q i

i = 1, 2, , 3N .

You might be asked to prove this in an upcoming assignment.

(8.31)

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 fixed to a
massless rod, which pivots aboutpO 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 specified 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 = 12 mv 2 = 12 m2 2 . The potential energy is
U = mgh = mg (1 cos ). Thus, the Lagrangian is

= T U = 1 m2 2 mg (1 cos ).
L = L(, )
(8.32)
2
Now, for a single particle we have (for a generalized coordinate q and our specific
coordinate )
d L L
d L
L
=
,
.
(8.33)
=
dt q
dt
q

For our Lagrangian, eq. (8.32), we have

d
= m2 .

(m2 )
(8.34)
dt
Referring back to the figure, we see that the left side of this equation is just the
torque exerted by gravity on the pendulum, while the term m2 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.


mg sin =

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 q1 , , qn are a set of generalized coordinates for the
system if each position r can be expressed as a function of q1 , , qn , and possibly
the time t :
(8.35)
r = r (q1 , , qn , t ) [ = 1, , N ] ,
and conversely, each qi can be expressed in terms of the r and possibly t :

qi = qi (r1 , , rN , 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 1023 , 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 confined 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
r1 = (1 sin 1 , 1 cos 1 ) = r1 (1 )
(8.38)
r2 = (1 sin 1 + 2 sin 2 , 1 cos 1 + 2 cos 2 ) = r2 (1 , 2 ).
Notice that the components of r2 depend on both of the generalized coordinates 1
and 2 . If the relation between the Cartesian coordinates r and the generalized
coordinates qi 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 specified 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 configuration. A system with
this property is said to be holonomic - a holonomic system has n DOF and can be
described by n generalized coordinates, q1 , , qn . Nonholonomic systems are
much more difficult 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 q1 and q2 , 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 Fcstr . 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 Ftot = Fcstr + F, and the Lagrangian, as usual, is L = T U . Since U is the
potential energy for the nonconstraint forces only, this definition 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 r1 and r2 , through which the particle passes at times t1 and
t2 . 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 defines (t ) as the infinitesimal 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 confined,
and therefore so does (t ). Furthermore, since the endpoints of the path are fully
specified, (t )=0 at t1 and t2 .
Now, we denote by S the action integral taken along the incorrect path:
Z t2
t )dt
S=
L(R, R,

(8.41)

t1

and by S0 the action integral taken along the correct path:


Z t2
S0 =
L(r, r , t )dt .
t1

(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 S0
(8.43)
is zero to first order in the distance between the paths.
The difference S is the integral of the difference between the Lagrangian on the two
paths,
t ) L(r, r , t ).
L = L(R, R,
(8.44)
If we substitute R(t ) = r(t ) + (t ) and
L(r, r , t ) = T U =

1 2
mr U (r, t ),
2

(8.45)

this becomes
L

=
=

i
1 h
2 r 2 [U (r + , t ) U (r, t )]
m (r + )
2
mr U + O (2 ),

(8.46)

where O (2 ) denotes 2nd- and higher-order terms in and 2 . The last step in the
above equation results from the definition 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 find that
S =

t2

L dt =

t1

t2
t1

[mr U ] dt .

The first term in the integral can be integrated by parts, leading to


Z t2
[mr + U ] dt .
S =

(8.47)

(8.48)

t1

Now, remember that the path r is the correct path, which means that it satisfies
Newtons second law. Therefore, mr is just the total force on the particle,
Ftot = Fcstr + F. But we also know that U = F. Cancellation of these terms leaves
Z t2
S =
(8.49)
Fcstr dt .
t1

However, the constraint force Fcstr 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 five 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 first two examples are fairly easily analyzed using Newtons second law, which
should increase our confidence 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 Confined 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 , specified 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 = k r directed to the origin. Using z and as
generalized coordinates - remember, is constant - find 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
z ). Therefore, the kinetic energy is
velocity in cylindrical coordinates for fixed = R is (v , v , vz ) = (0, R ,

T=

Fig. 4:

1
1
1
mv 2 = mv v = m(R 2 2 + z 2 ).
2
2
2

(8.50)

A mass m, confined to the surface of the cylinder = R and subject to a force F = k r - see JRT Fig. 7.7

PC235 Winter 2013 Chapter 8. Lagrangian Mechanics Slide 24 of 36


Examples

Example #3 - A Particle Confined to Move on a Cylinder cont


JRT Ex. 7.4
The potential energy for the force F = k r is U = 21 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 = 12 k (R 2 + z 2 ) and

z , z ) = 1 m(R 2 2 + z 2 ) 1 k (R 2 + z 2 ).
L = L(, ,
2
2

(8.51)

With two DOF, this system has two equations of motion. The two Lagrange equations read

d L
L
=
,
dt z
z

or

kz = mz,

L
d L
=
,
dt

or 0 =

mR 2 .
dt

(8.52)

The z equation tells us that the mass executes simple harmonic motion in the z direction, with the usual = k /m. The
The particles motion would be
equation tells us that the mass moves around the cylinder with constant angular velocity .

completely specified 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.

Choose a convenient set of n generalized coordinates q1 , , qn and find


expressions for the original coordinates of step 1 in terms of these
generalized coordinates.

Rewrite L in terms of q1 , , qn and q 1 , , q n .

Write down the Lagrange equations for each of your n generalized


coordinates.

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
simplifies 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 figure below shows a crude model of a yo-yo. A massless string is suspended vertically from a fixed
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 12 I 2 , where is the angular velocity about the center of mass, and the moment of inertia I is 21 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
1
1
3
1
1
T = mx 2 + I 2 = mx 2 + m(R )2 = mx 2 .
(8.53)
2
2
2
4
4
Taking +x as the downward direction, we have a potential energy U = mgx . Thus, the Lagrangian is
L=

3
mx 2 + mgx ,
4

(8.54)

which has just the one generalized coordinate, x . The Lagrange equation reads

d L
L
=
dt x
x

mg =

3
mx,
2

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

(8.55)

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 = 12 mv 2 + 12 I 2 = 21 m + I /R 2 x 2 . The PE is U = mgx sin . Thus,
the Lagrangian is

1
L=
m + I /R 2 x 2 + mgx sin
(8.56)
2
and the Lagrange equation is

Therefore,



mg sin = m + I /R 2 x.

x =

mg sin
,
m + I /R 2

(8.57)

(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

Isph =
the sphere will win the race.

2
1
mR 2 and Icyl = mR 2 ,
5
2

(8.59)

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 x1 be the horizontal coordinate of the plane (with positive x1 to the left), and let x2 be the horizontal coordinate of
the block (with positive x2 to the right). The relative horizontal distance between the plane and the block is x1 + x2 , so the
height fallen by the block is (x1 + x2 ) tan . The Lagrangian is therefore
i
1 h
1
(8.60)
L = M x 12 + m x 22 + (x 1 + x 2 )2 tan2 + mg (x1 + x2 ) tan .
2
2
The resulting Lagrange equations in the generalized coordinates x1 and x2 are

M x1 + m(x1 + x2 ) tan2 = mg tan


mx2 + m(x1 + x2 ) tan2 = mg tan

(8.61)

We can immediately see that the difference of these two equations yields the required conservation of momentum:

M x1 mx2 = 0 (d /dt )(M x 1 mx2 ) = 0.

Fig. 6:

Geometry for this example

(8.62)

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 x1 and x2 , which we can solve to find the acceleration of the
plane:
mg sin cos
x1 =
.
(8.63)
M + m sin2
This is a good time to look at limiting cases. When = 0, x1 = 0 (makes sense, nothing will move in this case). When m = 0,
x1 = 0 (theres no component of normal force to push the plane). When m M , x1 = 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 specified 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 find (but not solve) the equations of motion in x and .
Solution: The kinetic energy, as usual, is T = 21 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
(x , ( + x ))
= x 2 + ( + x )2 2 ,
v 2 = v v = (vr , v ) (vr , v ) = (x , ( + x ))
and therefore, T

(8.64)

= 21 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 Ug , we have
1
U = mg ( + x ) cos + kx 2 .
(8.65)
2

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
1
m(x 2 + ( + x )2 2 ) + mg ( + x ) cos kx 2 .
2
2

(8.66)

The Lagrange equation in x reads


L
d L
=
m( + x ) 2 + mg cos kx = mx.
x
dt x

(8.67)

mg cos kx = m(x ( + x ) 2 ).

(8.68)

We can rearrange this to the form


I claim now that this is nothing more than Newtons second law in the radial direction, Fr = mar . 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 L

=
mg ( + x ) sin = m( + x )2 2 + 2m( + x )x ,

dt

(8.69)

h
i
mg sin = m ( + x ) 2x .

(8.70)

which we will rearrange to

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 qi ,
we refer to the n quantities L/qi = Fi as generalized forces and L/q i = pi as
generalized momenta. With this relation, the Lagrange equation
L
d L
=
q i
dt q i

(8.71)

can be rewritten as Fi = dtd pi . That is, the generalized force equals the rate of
change of generalized momentum. In particular, if the Lagrangian is independent of
= 0, and the corresponding generalized
a particular coordinate qi , then Fi = L
qi
momentum pi 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
1
(8.72)
L = L(x , y , z , x , y , z ) = m(x 2 + y 2 + z 2 ) mgz .
2
In this case, the generalized force is just the usual force (L/x = U /x = Fx ,
etc.) and the generalized momentum is just the usual momentum
(L/x = mx = px , etc.) Because L is independent of x and y , it follows from
Lagranges equations that the components px and py 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 qi , the corresponding generalized
momentum is conserved. When the Lagrangian is independent of a coordinate qi ,
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 fit 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 significant 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 significant 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 (r1 + , , rN + , t ) = U (r1 , , rN , 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 infinitesimal displacement in the x direction, then all of the x coordinates
x1 , , xN increase by , while the y and z coordinates are unchanged. For this
translation,
L
L
= 0.
(8.74)
+ +
L =
x1
xN
Since is arbitrary, this implies that
N
X
L
= 0.

x
=1

(8.75)

But using Lagranges equations, we can rewrite each derivative in the sum as
L
d L
d
=
= px
x
dt x
dt
where px is the x component of the momentum of particle .

(8.76)

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
N
X
X
d
d
L
=
px = Px = 0

x
dt
dt

=1
=1

(8.77)

where Px is the x component of the total momentum P = 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