You are on page 1of 7

Momentum

Nirmaan Shanker
October 2015

Introduction

We started our discussion on conservation laws in the last set of notes with work and energy. Work and energy gave
us another means to analyze various situations. In this set of notes, we will discuss momentum and its conservation.
As with the work and energy set of notes, this set of notes will be on the longer side as I would like to present a
thorough description of momentum with plenty of examples.

Momentum

2.1

A Confession

Remember Newtons second law? We know that it is F~ = m~a. But what I havent told you is that Isaac Newton did
not actually state his most famous law as F~ = m~a, but instead as,
d~
p
F~net =
,
dt
where p~ is the momentum. Clearly by the nature of the relationship, p~ = m~v .

2.2

Dynamics of Several-Particle Systems

So the above expression for Newtons second law helps us understand that the vector sum of the all the forces on
a particle is equal to the instantaneous rate of change of the particles momentum. Lets briefly extend this to a
several-particle system.
Lets consider a system of N particles. We are free to choose the boundaries of this system, but the moment we
pick the boundaries, we cannot change them in the rest of the analysis of the problem. For every particle of the
system, we know that:
d~
pi
f~i =
dt
We can split the force term into two terms as follows,
d~
pi
f~iext + f~iint =
,
dt
where f~iext is the sum of forces exerted on particle i from outside the system and f~iint is the sum of forces exerted
on particle i from within the system.
We can sum up the above equation for every single particle, and we obtain:
X

f~iext +

f~iint =

X d~
pi
dt

By Newtons third law, the sum of the internal forces must be zero, as every force has an equal and opposite force
within the system. Thus we see that the the instantaneous rate of change of the total momentum within the system
is equal to the sum of the external forces acting on the system. Mathematically, this is expressed as:
dP~
F~ext =
dt

Center of Mass

Interestingly enough, the final equation in the previous section looks remarkably the same as the equation for a single
point-particle. Lets stretch this similarity a little further. We can rewrite the above expression as:
X
~
~ = F~ext = dP =
MA
mi ai
dt
~ be the acceleration of? Clearly not every particle is
M represents the total mass of the system. But what would A
~ as the center of mass. By
accelerating at the same acceleration. So lets define the point which is travelling at A
the above equation, we immediately notice that the position of the center of mass (denoted as Xcm ) must be equal
to:
Xcm =

1 X
mi xi
M

At the center of mass, the system behaves as if all the mass is at that point and all the external forces of the system
acts at this point. Just as a reminder, remember that we are dealing with vector quantities and thus we must specify
an origin and a coordinate system on which the vector is defined.
The center of mass also helps justify us casually treating the masses in the notes on forces as point particles.
Since the particles were rigid objects, the center of mass helps provide a relatively good sense of the objects motion.
But we must also accept that we are oversimplifying a variety things with such a model. This model only helps us
understand the translation of the object in space but gives no care for the orientation in space (well deal with that
when we start considering rotation in future notes).
Anyways, the above definition for the center of mass works well when I have a system of particles, it will not
work that well for continuous masses. However, we can still attempt to use it. Lets break up our continuous mass
into several pieces, with each piece having a corresponding mass, mj . Therefore, the center of mass will be:

Xcm =

N
1 X
mj xj
M j=1

This is still a crude approximation as these mass elements are not really particles. However, if we let this pieces
become really small (infinitesimally small), our approximation becomes exact:
N
1 X
mj xj
N M
j=1

Xcm = lim

Amazingly, this looks like an integral. Thus, the center of mass for a continuous mass is simply:
Z
1
Xcm =
x dm
M
Now that we know the theory on how to find the center of mass, lets look at an example (specifically at a continuous
mass).

3.1

Nonuniform Rod

x
,
Problem Statement: A rod of length L has a nonuniform density. The linear mass density, , varies as = 0 L
where 0 is constant and x marks the distance from the end marked 0. Find the center of mass of the rod.

Solution: In order to calculate the center of mass, we must first find the total mass of the rod. However, this is
easily done as we know the density of the rod. The mass of a certain length, dx, is dm = dx. Then we simply
integrate along the length of the rod as shown below:

Z
M=

dx =

dm =

0 x
1
dx = 0 L
L
2

Now that we can have found the total mass of the rod, we can now find the center of mass. The integration is shown
below:

Xcm

1
=
M

x dm
Z L
L
2
2
2
0 x2
=
dx = 0 L
x dx =
0 L 0
0 L 0
L
3
Z

Xcm
And we are done.

Impulse

At the end of section 2, we showed that:


d~
p
F~ext =
dt
However, this is just a differential equation. We separate the differentials and integrate both sides to get:
Z

t2

~
p=

Fext dt
t1

The quantity, ~
p, is known as the impulse. Although providing no new information than the original expression,
this expression helps us understand that a change in momentum of an object results from a force acting on it for a
specified time.
We can also use impulse to help find the average force exerted on an object. Since the integral is nothing more
than a representation of an area, we can find a rectangle with the same area with base t and height Fa ve. Thus,
~
p = F~ave t
Lets now consider an example.

4.1

Calculating the Average Force on a Ball that Bounces on a Floor

Problem: A rubber ball of mass 1 kg falls down to the floor from a height of 4 m. The ball bounces and rebounds
with approximately the same speed as it had landed. Thanks to our amazing fast reflexes with a stopwatch, we know
that the ball was in contact with the floor for approximately 2 103 seconds. What was the average force exerted
on the ball by the flow:

Solution: We know from kinematics and energy that the velocity the rubber ball
gained from the drop is 2gh.
Therefore, the momentum of the rubber ball at the bottom of the drop is m 2gh. We can now compute the
impulse:
3

p
p
p
~
p = m 2gh (m 2gh) = 2m 2gh
Thus, the average force exerted on the rubber ball is:

Fave =
Fave

p
t

2m 2gh
=
= 8858.9 N = 8.86 kN
t

So the average force exerted by the floor on the ball is 8.86 kN, which is a sizable force.

Conservation of Momentum and Collisions

So at the end of section 2, we showed that F~ext =


forces is 0. That implies that

~
dP
dt

. Now lets consider the case where the sum of the external

dP~
= 0,
dt
and thus P~ must be conserved. This fact is very useful to us as we can now assert that the momentum must be the
same at any two times, as long as there are no external forces acting on the system.

5.1

Types of Collisions

The most common examples that display conservation of momentum are collisions. Since the force of interaction
in a collision of two objects is an internal force, momentum is conserved. There are several types of collisions, lets
consider a few:
5.1.1

Elastic Collision

An elastic collision occurs where the kinetic energy of the system is conserved. The nature of the collision is as
follows. When one particle collides with another particle, both particles begin to deform until they both reach
the same velocity (this is like when the spring reaches its maximum compression). Some of their kinetic energy is
converted to elastic potential energy. Now at this stage, internal forces begin developing inside the ball and and the
balls deform to their original states (the spring decompresses). Since all the forces were conservative, the kinetic
energy is conserved after the collision.
We can attempt to find the final velocities of two particles undergoing an elastic collision in one-dimension.
Assuming a particle with mass m1 and velocity v1 collides with a particle of mass m2 and velocity v2 . We have two
relationships that we know that the system must satisfy:

m1 v1 + m2 v2 = m1 v1f + m2 v2f
1
1
1
1
2
2
m1 v12 + m2 v22 = m1 v1f
+ m2 v2f
2
2
2
2
We have two unknowns and two equations. Therefore, we can solve the system of equations. The solving of this
system is not the cleanest and is not presented here (you are recommended to solve it on your own). The solutions
are presented below:
1 There is actually a flaw in this analysis. The average force we calculated was the total average force. But there are two forces at play:
the force exerted by the floor and gravity. But as it turns out, the impulse caused by the gravitational force is negligible (a thousandth
of the impulse generated by the floor).

(m1 m2 )v1 + 2m2 v2


m1 + m2
2m2 v1 + (m2 m1 )v2
=
m1 + m2

v1f =
v2f

Not the cleanest answer, but it gets the job done.


5.1.2

Partially Elastic Collisions

So in elastic collisions, the two particles completely deform back to their original shapes. Now you can imagine that
the likelihood of a complete return to the original state is quite low. Thus in this type of collision, we only know
that the momentum is conserved. In one dimension, this situation can be analyzed reasonably well. But when we
expand to two or three dimensions, there is very little that we can do.
5.1.3

Inelastic Collisions

In an inelastic collision, there are no internal forces that develop when both particles reach the same velocity. Thus,
the two particles essentially stick together after the collision. One easy example in real life is when you throw a piece
of gum and it sticks to the surface it is thrown at.

5.2

Example 1

5.3

Example 2

6
6.1

Center of Mass in Conservation of Momentum


Motion of the Center of Mass when Momentum is Conserved

We can also ask ourselves how the center of mass moves when momentum is conserved. We know that:
dP~
F~ext =
= M acm
dt
Therefore, when F~ext is 0, the velocity of the center of mass must be constant. This approach can be used to solve
several problems. Consider the example below:

6.2

Example: Object Explosion

Problem: A box with a bomb is thrown up in the air with velocity of magnitude 10 m
s at 45 degrees with respect
to the horizontal. At the top of the throw, the bomb explodes. One half of the box falls vertically down. How far
does the other half of the box go from the initial point of release of the box?
Solution: The key thing to note is that the explosion is an internal force. Thus, momentum is conserved in
the horizontal direction. It isnt conserved in the vertical direction, since gravity is still acting on the box. Thus,
the velocity of the center of mass will remain constant. All that means is that the center of mass would travel in a
parabolic path as if the box never exploded at all. From kinematics we know that the range of projectile thrown
is:
R=

v02 sin 2
g

Clearly, the first half of the box which fell straight down, fell at x = 12 R. Since the center of mass will land at x = R,
the other piece must fall x =

3R
2

3v02 sin 2
2g

away from the release point.


5

6.3

Center of Mass Frame

While our method for solving these momentum problems works okay in one dimension, it can be quite cumbersome
to apply in two and three dimensions. This is extremely true in analyzing elastic collisions (if the solving was so
ugly in one dimension, imagine the horror in two and three dimensions). In order to simplify the work, it is often
convenient to work in the center of mass reference frame. In this frame, the velocity of the center of mass is
zero. Recall the velocity of the center of mass was defined as:
V =

m1 v1 + m2 v2
m1 + m2

We can find the new velocities in the COM frame by simply subtracting V from the original velocity:
vCOM = vlab V
But what makes this reference frame so special? The key feature about the center of mass reference frame is that
the total momentum of the system is zero. We can quickly prove this for two particles (extending to N particles is
not difficult and is in the problem set). The momentum of the each particle in the center of mass frame is:
m1 v1 + m2 v2
m2 v1 m2 v2
) = m1
m1 + m2
m1 + m2
m1 v1 + m2 v2
m1 v2 m1 v1
p2 = m2 (v2
) = m2
m1 + m2
m1 + m2

p1 = m1 (v1

As you can see, there are indeed additive inverses of each other. This fact places an important restraint on collisions in
the center of mass reference frame. Since the total momentum must be 0 in the center of mass frame, the momentum
vectors before and after the collision must be parallel to each other (while this may sound trivial in one-dimension,
this facts gains more use in higher dimensions).

6.4

Elastic Collisions in One-Dimension, Round 2

We will now analyze the one dimensional elastic collision once more, but now in the center of mass frame. The key
thing to note is that after the collision, the two particles must flip directions and proceed at the same speed. Now,
the conservation of energy statement does not lead to the conclusion that the velocity must be at the same speed,
but rather conservation of energy requires it to be so. So for a particle in an elastic collision the following is true:
vi,COM = vf,COM
Converting back to the lab frame we have:

vi,lab V = (vf,lab V )
vf,lab = 2V vi,lab
This is the same result presented as we presented earlier in a different form. But clearly this solution is much cleaner
and much more approachable. This result works for the second object as well. 2

Conservation of Momentum with Changing Mass

Throughout our various discussions in physics so far, we have assumed that the object of a mass remains constant.
But that does not always have to be through. If the mass does indeed change, then we can express Newtons second
law as:
2 As

a side note, there will hopefully be extra set of notes, where we briefly explore elastic collisions in higher dimensions.

Fext =

d~
p
d~v
dm
=m
+ ~v
dt
dt
dt

There is one classic example where the mass changes: rockets. The way a rocket increases its velocity is by expelling
part of itself in the back. Lets try to model this and find how much does the velocity actually changes. We will
first assume that there are now external forces acting on the rocket. We will assume that at some time t, we have
a rocket travelling with mass m and velocity v0 . After some time t, it kicks out a mass m out of the back with
velocity ve with respect to itself which increases its velocity by some v. Therefore the conversion of momentum
statement for this situation:
mv0 = (m m)(v0 + v) + m(v0 ve )
We can simplify this equation to the following,
mve = mv
where in the last step we have ignored the double differential term as that term is pretty much zero. Now we can let
these quantities get arbitrarily small and replace them with differential elements. We will also use dm = m as
ejecting a mass decreases the mass of the rocket. Separating variables at the same time, we have:
dv = ve

1
dm
m

We can now integrate and we get the following result:

vf

mf

dv = ve
v0

dv
m0

v = ve ln

m0
mf

The above relationship is also known as the Tsiolkovsky (tsyawl-kawf-skee) rocket equation. One thing that this
result shows is that it is very difficult to increase the speed of a rocket by ejecting mass from the back as it increases
logarithmically, which is rather slow.

You might also like