Вы находитесь на странице: 1из 72

Physics I Honors Monday 25 August Fall 2008

First class! Name and course on board, read roll, introduce TAs.
Distribute outline, lecture, and lab schedules. Discuss them.
New material for today: Start thinking like a physicist. Two new ideas, namely
(1) Solving problems with dimensional analysis, and
(2) F = ma is a dierential equation
Todays Topic #1: Dimensional Analysis
Given a problem, ask What is the important physics?. Then list the quantities that have
something to do with that physics. In many cases, you can combine these things to get the
answer, at least to within a factor of 2 or or something like that.
Convenient notation:
[distance] = L
[time] = T
[mass] = M
[temperature] = K
Example: What is the radius of a black hole of mass M? This is a combination of gravity
and relativity, so use M, G, and c. Newtons constant G has units m
3
/kg-sec
2
, so [G] =
L
3
M
1
T
2
. (Easy to see this from F = GmM/r
2
= ma.)
Let the radius be R = M
x
G
y
c
z
where we need to nd x, y, and z. This gives
[R] = M
x
L
3y
M
y
T
2y
L
z
T
z
= M
xy
L
3y+z
T
2yz
= L
1
(1)
so x = y, 3y + z = 1, and 2y + z = 0. The second and third of these say that y = 1, and so
x = 1 and z = 2. Therefore we get
R = GM/c
2
(2)
The correct answer is
1
2
GM/c
2
, called the Schwarzchild Radius.
(Hint: Useful website for units and constants is
http://pdg.lbl.gov/2008/reviews/consrpp.pdf)
1
Practice Exercise
Youve heard of String Theory. The length of strings is supposedly the Planck Length,
since gravity is unied with quantum mechanics and relativity at this distance. Gravity is
governed by Newtons gravitational constant G, quantum mechanics by Plancks constant h,
and relativity by the speed of light c. Combine these to nd the Planck Length. (Note that
the dimensions of h are distancemomentum, or L
2
M/T.) Look up values of G, h, and c
and compare the Planck Length to the size of a proton (10
15
m).
2
Physics I Honors Monday 25 August Fall 2008
Todays Topic #2: Newtons Second Law as an Equation of Motion
The most important of Newtons Laws is the second. For motion in one dimension, we write
F = ma. (Well get to

F = ma later this week.) Physicists refer to this as the equation of
motion for a particle moving through time in one dimension of space.
Let x(t) be the position x of a particle at time t. Then the velocity is v(t) = dx/dt x.
(Have you all seen this notation before??) Acceleration is a(t) = dv/dt = v = x. So
Newtons Second law can be written as
x =
1
m
F(x, t) (3)
where we make it clear that the force F may itself depend on position and/or time.
Equation 3 is called a dierential equation. Unlike an algebraic equation, whose solution
is a value for a quantity like, say, x, the solution of a dierential equation is a function, say
x(t). Since x(t) describes the motion of a particle with mass m acted upon by a force F(x, t),
we call Eq. 3 the equation of motion.
Simple example. Let F(x, t) = F
0
, a constant independent of both x and t. Then
x = v =
dv
dt
=
F
0
m
(4)
v(t) =
dx
dt
=
F
0
m
t + v
0
(5)
x(t) =
1
2
F
0
m
t
2
+ v
0
t + x
0
(6)
where the last line is the solution to this equation of motion. This might look more familiar
to you from your high school physics course, if we write a = F
0
/m. You probably called
this motion under constant acceleration, but of course a constant acceleration implies a
constant force, so its the same thing.
Each step in the integration above involves a constant of integration. I called these con-
stants v
0
for the rst step, and x
0
for the second step. These are standard notations, and
in fact good ones. The constant v
0
is the velocity at time t = 0, also called the initial
velocity. Similarly, x
0
is the initial position. In order to solve any equation of motion, we
will need to specify these initial conditions.
3
Practice Exercise
Let y(t) describe the one dimensional motion of a particle with mass m that moves vertically.
Assume that the particle is acted on by a (constant) force F = mg. (Isnt that peculiar, a
force that is exactly proportional to the quantity m that appears in the equation of motion?)
Assume that the particle starts out at y = 0 with an initial (upward) velocity +V . Find the
maximum height to which the particle climbs, in terms of m, g, and V . Make a sketch of
y(t) as a function of t. What is the shape of the curve? (Mathematicians have a name for
it, and Im sure youve heard it before.)
4
Physics I Honors Thursday 26 August Fall 2008
This class will concentrate mainly on mathematics. First we will review vectors, and and
apply them to three dimensional motion. Then well talk about Taylor series, how they can
be used in approximations, and then apply them to complex numbers.
Todays Topic #1: Vectors and Motion in Three Dimensions
Introduction: The vector equation F = ma is a shorthand for three equations, namely
F
x
= ma
x
and F
y
= ma
y
and F
z
= ma
z
Notation: A in book,

A on board. Unit vectors are n in book, n on board.
The length of a vector is [A[ = A, so [ n[ = 1.
If you add two vectors, the result is still a vector: C = A+B
A number times a vector is still a vector: C = bA This means that A = A

A.
The dot product between two vectors is a scalar: A B = ABcos where is the angle
between the two vectors. This is the projection of A onto B or vice versa.
The cross product between two vectors is a vector: A B = C (Watch out! Some day
youll learn about polar and axial vectors!) The magnitude of C is ABsin . The direction of
C is given, by convention, by the right hand rule. This means that AB = BA.
Vectors can be written in terms of coordinates in any given coordinate system or reference
frame. For an (x, y, z) coordinate system, dene base vectors

i,

j, and

k, all orthogonal (so
their mutual dot products are zero) and with

i

j =

k, etc. Then
A = A
x

i + A
y

j + A
z

k (7)
A B = A
x
B
x
+ A
y
B
y
+ A
z
B
z
(8)
AB = (A
y
B
z
A
z
B
y
)

i + (A
z
B
x
A
x
B
z
)

j + (A
x
B
y
A
y
B
x
)

k (9)
(1) General Motion in Three Dimensions. A vector r(t) = x(t)

i + y(t)

j + z(t)

k gives
the motion of a particle in three dimensions. The velocity vector is v(t) = r(t) and the
acceleration vector is a(t) = r(t). The (vector) equation of motion is r(t) = F(x, t)/m.
(2) Examples. Work W = Fd is a dot product. Torque = rF is a cross product.
(3) Uniform circular motion. The motion r(t) traces out a circle at a constant rate.
Let be the number of radians per second traced out,
i.e. /2 revolutions per second. (Review radians?)
r(t) = r(

i cos t +

j sin t) (10a)
v(t) = r(

i sin t +

j cos t) (10b)
Note : v r = 0
a(t) =
2
r(

i cos t +

j sin t) =
2
r(t) (10c)
Note : a =
2
r = v
2
/r (10d)
The force that is responsible for this acceleration must
point back to the center of the circle!
5
Practice Exercise Find the period P of the conical pendulum pictured below, in terms
of the hanging angle , the string length L, and the acceleration g due to gravity.
The mass m executes uniform circular
orbits with radius R. If the speed is v,
then the period is just the time it takes
to execute one orbit, namely P = 2R/v.
So, you need to gure out v in terms of R,
or perhaps in terms of the string length L
and the hanging angle , since R = Lsin .
Since we know that F = ma, we just need
to gure out what the force F is. Clearly,
this force is provided by the tension T, but
only by part of it.
6
Physics I Honors Thursday 26 August Fall 2008
Topic #2: Taylor Series, Approximation Techniques, and Complex Numbers
It makes sense that we can approximate any function by some high order polynomial:
0 1 2 3 4
0
2
4
6
8
10


Zeroth order
First order
Second order
Full function
The function is expanded about the point
x
0
= 1. Shown are the zeroth order (i.e.
a constant) approximation, the rst order,
second order, and the full function. The
zeroth order is just the value of the func-
tion at x
0
= 1, and the rst order is the
tangent line at that point. The second or-
der approximation comes close to the full
function in the neighborhood of x = x
0
.
f(x) = a
0
+ a
1
(x x
0
) + a
2
(x x
0
)
2
+ a
3
(x x
0
)
3
+ (11)
f(x
0
) = a
0
f

(x
0
) = a
1
f

(x
0
) = 2 a
2
f

(x
0
) = 3 2 a
3
(12)
f(x) = f(x
0
) + f

(x
0
)(x x
0
) +
1
2!
f

(x
0
)(x x
0
)
2
+ (13)
Examples : (1 + x)

= 1 + x +
1
2!
( 1)x
2
+ (14)
e
x
= 1 + x +
1
2!
x
2
+
1
3!
x
3
+
1
4!
x
4
+
1
5!
x
5
+ (15)
sin(x) = x
1
3!
x
3
+
1
5!
x
5
+ (16)
cos(x) = 1
1
2!
x
2
+
1
4!
x
4
+ (17)
Obviously, if [x[ 1, i.e. x is small, then only a few terms (maybe just one) is needed in
order to get an accurate approximation.
Complex numbers. There is a number i such that i
2
= 1. Numbers formed by a real
number times i, like 2i, i

3, or xi where x is a real number, are called imaginary


numbers. Dont let the words fool you, though. All of these are perfectly valid numbers. The
imaginary numbers are a group outside integers, rational, and irrational (real) numbers.
Numbers like 3 + 2i and z = x +iy are called complex numbers. They have a real part
and an imaginary part. We write, for z = x + iy where x and y are real, x = Re(z) and
y = Im(z). The modulus of z is [z[ =
_
x
2
+ y
2
.
Now, heres a neat thing:
e
ix
= 1 + ix +
1
2!
(ix)
2
+
1
3!
(ix)
3
+
1
4!
(ix)
4
+
1
5!
(ix)
5
+
=
_
1
1
2!
x
2
+
1
4!
x
4
+
_
+ i
_
x
1
3!
x
3
+
1
5!
x
5
+
_
= cos(x) + i sin(x)
This is called Eulers Formula. It is very useful in physics and engineering!
7
Practice Exercise
First, a math rule. The derivative of an exponential is an exponential. That is
f(x) = e
ax
= f

(x) = ae
ax
Use this to nd the solution to the equation of motion for a particle of mass m subject to
a force F(x) = kx where k is a positive constant. Write down the equation of motion as
x(t) = F/m. Then show that x(t) = Ce
it
is a solution to the equation of motion, for any
value of C, so long as has one of two possible values. What are those values?
Next, write the general solution x(t) as the sum of two functions Ce
it
, one for each of
the two possible values of and with two dierent values of C. (Call them C
1
and C
2
.) Is it
obvious to you that this sum of solutions is also a solution to the equation of motion?
Finally, determine the two constants C
1
and C
2
from the initial conditions x(0) = A and
x(0) = 0. Simplify the result as much as possible. (Eulers formula will be handy.) What
kind of motion does this describe.
8
Physics I Honors Thursday 4 September Fall 2008
Lets try to stick with the 20-20-20 plan, that is new material for 20 minutes, 20 minutes
to work on a practice exercise, and then 20 minutes to talk about the solution.
Todays Topic #1: Applications of Newtons Laws
Discuss Free Body Diagrams. Statics rst. Use example of one block A sitting atop another
block B (page 68 in K&K) and take care to note the weight of each block, the normal force
on B from the lower surface, and the force F
1
(F
2
) which block B (A) exerts on block A
(B). Hence F
1
= M
A
g and N = F
2
+ M
B
g. Third law says F
1
= F
2
so N = (M
A
+ M
B
)g,
which is just what you expect, of course.
Pretty much the same thing for two masses hanging, connected by a wire under tension. The
top wire hangs both masses! Tell the story of the Hyatt Regency disaster, July 17, 1981 in
Kansas City, Missouri: http://en.wikipedia.org/wiki/Hyatt Regency walkway collapse
Dynamics is very similar, just dont set a = 0. See practice problem and also homework.
Lets do the conical pendulum. (See Example
2.8 in your textbook.) Find the period P in
terms of the angle , the string length L, and
the acceleration g due to gravity. The speed
of m is v = R, so P = 2R/v. To nd v, use
F = ma with a =
2
R = v
2
/R. (Recall Equa-
tions 10.) Find F = T sin with T cos = mg.
Putting it all together, we have
P
2
= (2)
2
R
2
v
2
= (2)
2
R
a
= (2)
2
Rm
T sin
= (2)
2
Rmcos
mg sin
P = 2

Rcos
g sin
= 2

Lcos
g
Practice Exercise
!
"
!
#
Find the acceleration a of the coupled masses on the left, in
terms of m
1
, m
2
, and g. The mass m
1
slides on a horizontal,
frictionless surface, m
2
hangs freely, and the string that
connects them is massless. The pulley is also massless, and
gives no frictional resistance to motion.
Start by drawing free body diagrams of m
1
and m
2
. Identify
all forces on each of the two masses, and then work with the
components that are in the direction of the acceleration.
9
Topic #2: The Everyday Forces of Physics
Start by making a list. Use vector notation!
Tension, the pull of a massless string: T
Weight, i.e. gravity near the Earths surface: W = mg

k
Springs, i.e. Hookes Law: F = kx
Frictional forces
Kinetic friction: f
k
=
k
N opposing the direction of motion
Static friction: f
s

s
N as big as it needs to be to keep the object from moving
Note that
s
>
k
for the same two surfaces. Demo with book on desk!
Viscous forces: F
v
= Cv
Inverse square central forces (Draw some pictures here.)
Newtonian gravity: F
ab
= (GM
a
M
b
/r
2
)r
ab
Electrostatic attraction: F
ab
= +(kQ
a
Q
b
/r
2
)r
ab
Note: General central force is just F
ab
= f(r)r
ab
Lets study the motion of an object under the inuence of a drag force. (See Example 2.16.)
Assume a mass m is moving through a viscous medium. The equation of motion is
Cv = ma = m
dv
dt
or
dv
dt
=
C
m
v (18)
Now integrate. The math trick you need is d(ln x)/dx = 1/x. We have
_
v
v
0
dv
v
=
C
m
_
t
0
dt or ln
_
v
v
0
_
=
C
m
t or v(t) = v
0
e
Ct/m
(19)
Note use of initial condition notation. The behavior is clear. The particle moves with
velocity v
0
at t = 0, but then slows down exponentially.
Note: Study the shell theorem in Note 2.1. Basically, a spherical object attracts an external
mass gravitationally, where the equivalent object mass is all that is contained within a sphere
or radius intersecting the external mass.
Practice Exercise
Express the gravitational acceleration near the Earths surface, g, in terms of G, and the
mass M
E
and radius R
E
of the Earth. Start by putting an external mass m on the surface
of the Earth. Then calculate the gravitational attraction of the Earth on this mass, using
the formula for Newtonian gravity.
Cavendish devised an sensitive experiment that measured G in 1798. People say he weighed
the Earth. What do they mean? (Note that Eratosthenes measured the Earths radius
somewhere around 200 B.C.)
10
Physics I Honors Monday 8 September Fall 2008
Todays Topic #1: Momentum and Systems of Many Particles
Newton actually wrote the Second Law as

F = dp/dt where momentum p = mv. If mass


m doesnt change with time then

F = mdv/dt = ma, so we recover the old version.


Now imagine an object colliding with a brick wall. The collision will change the momentum
of the object. The force of the collision varies over the collision time. (Draw a gure; See
Example 3.9 in K&K.) We can use Newtons second law to write an expression involving the
integral of the force and the change in momentum:
p =
_
t
2
t
1
F(t)dt J
We call J the impulse. The average force F
avg
is J divided by t = t
2
t
1
.
Now imagine two objects colliding with each other. Each exerts an equal and opposite
momentum on each other. That is J
1
= J
2
, or p
1
+ p
2
= (p
1
+ p
2
) = 0. In other
words, the total momentum P p
1
+ p
2
is conserved in collisions. Note that this doesnt
work if there is some external force acting on the masses.
Now look at two particles again, but not just collisions. Allow some external force F
ext
to
act as well as internal forces. Then
dP
dt
=
dp
1
dt
+
dp
2
dt
= m
1
a
1
+ m
2
a
2
=

F
ext
The last equality follows because the internal forces cancel in the sum. It is convenient to
write P = Mv
cm
where M = m
1
+ m
2
and v
cm
= dr
cm
/dt with
r
cm
=
m
1
r
1
+ m
2
r
2
m
1
+ m
2
=
1
M
(m
1
r
1
+ m
2
r
2
) (20)
in which case Newtons Second Law reads

F
ext
= Mr
cm
= Ma
cm
. We call r
cm
the center
of mass. (Is center of momentum better?) For more than two particles
r
cm
=
m
1
r
1
+ m
2
r
2
+ + m
N
r
N
m
1
+ m
2
+ + m
N
=
1
M
N

n=1
m
n
r
n
(21)
Dont forget that this is really two (or three) equations for x
cm
, y
cm
, and (if three) z
cm
. The
gures below show how this all works out:
11
Practice Exercise
An elastic collision is one in which kinetic energy is conserved. (The kinetic energy of an
object with mass m and speed v is K =
1
2
mv
2
. The total kinetic energy is the sum of the
individual objects kinetic energies. We will learn more about this later.) Let two objects,
one with mass m and the other with mass 2m collide elastically in one dimension. They are
initially heading towards each other with speed v. Find their speeds and directions after the
collision, in terms of m and v.
Topic #2: Calculating the Center of Mass of Solid Objects
If we are working on a solid object, we could sum over all the atoms that make it up, but
that would be nuts. Instead, we use calculus, and assume that the object is continuous.
Then the sum becomes an integral. That is
M =
_
dm and r
cm
=
1
M
_
dm r
We usually write the innitesimal mass dm as a density (usually called or ) times a
volume element dV (or some other measure), i.e. dm = (r)dV . This allows the density to
vary as a function of position inside the solid object. The rest is in the details.
Here we can slice up an eggplant, making the point that
_
dV is the sum of small pieces.
Here is a simple example. What is the volume of a right circular cone of height h and base
radius R? Let z measure the height, and slice the cone up into thin disks, each of thickness
dz. The radius of each disk is R(h z)/h = R(1 z/h), so the volume of each disk is
R
2
(1 z/h)
2
dz. Therefore, the volume of the cone is
_
dV =
_
h
0
R
2
(1 z/h)
2
dz = R
2
_
h
0
_
1
2z
h
+
z
2
h
2
_
= R
2
h
3
or one-third of the base area times the height. Of course this is volume, not mass, but thats
basically the same thing for constant density. If the density is not a constant, youd need to
include the functional form of (r) and that would be a harder problem
Practice Exercise
Find the position of the center of mass of a long thin rod of mass M and length L. The
density of the material making up the rod changes linearly from zero at one end. Start by
realizing that the mass of a short segment of length dx is dm = (x)dx where (x) is the
linear mass density. Write a formula for (x), and determine any parameters by xing the
mass of the rod to be M. Then, nd the position of the center of mass.
12
Physics I Honors Thursday 11 September Fall 2008
Todays Topic #1: Work and Kinetic Energy in One Dimension
Lets use some calculus to manipulate F = mdv/dt, in one dimension:
_
b
a
F(x)dx = m
_
b
a
dv
dt
dx = m
_
b
a
dv
dt
dx
dt
dt = m
_
b
a
vdv =
1
2
mv
2
b

1
2
mv
2
a
(22)
We write F = F(x) but it could be a constant, for example F = mg:
mg
_
y
b
ya
dy = mg(h) =
1
2
mv
2

1
2
m0
2
=v =
_
2gh (23)
for an object falling through a distance h = y
a
y
b
starting from rest. We already knew this
from solving the equation of motion (Eq. 6) but here weve done this a dierent way.
A harder, but still easy, example is a spring force F = kx. Equation 22 becomes

1
2
kx
2
b
+
1
2
kx
2
a
=
1
2
mv
2
b

1
2
mv
2
a
(24)
You will play around with this on the practice exercise.
These are both examples of something called the Work-Energy Theorem. Dene the
work (in one dimension) done by a force F(x) on a mass m through a path x
a
x
b
as
W
ba
=
_
b
a
F(x)dx. Dene the kinetic energy of that object as K =
1
2
mv
2
, for velocity v.
Then Eq. 22 says that
W
ba
= K
b
K
a
K (25)
That is, the work done on the object equals the change in kinetic energy. Of course, the
work done for a constant force is just the product of the force times the displacement.
Lets use this to calculate escape velocity. A rocket
starts at r
a
= R
e
with (escape) velocity v. It ends up
at very large distance, r
b
, with v = 0. (If the
initial velocity were larger, it would have a nite speed at
innite distance. If smaller, it would fall back to Earth.)
_

Re
_

GM
e
m
r
2
_
dr =
GM
e
m
r

Re
=
GM
e
m
R
e
= 0
1
2
mv
2
=v =
_
2GM
e
R
e
(26)
We will consider all these ideas from a purely energy point of view on Monday.
Practice Exercise An object with mass m is acted on by a spring force F = kx. It has
zero velocity when x = A. Use Eq. 24 to nd the velocity v = V when x = 0, in terms of k,
m, and A. Explain why this is consistent with x(t) = Acos t, which solves the equation of
motion so long as
2
= k/m.
13
Topic #2: Work and Kinetic Energy in Several Dimensions
How do we generalize Eq. 22 to recognize that force is a vector, i.e. F = mdv/dt? Assuming
we want to end up at the work energy theorem, we can try working backwards:
d
dt
(v
2
) =
d
dt
(v v) = v
dv
dt
+
dv
dt
v = 2v
dv
dt
= 2
dr
dt

dv
dt
(27)
K =
_
b
a
d
_
1
2
mv
2
_
=
_
b
a
1
2
m
d
dt
(v
2
)dt =
_
b
a
m
dr
dt

dv
dt
dt =
_
b
a
_
m
dv
dt
_
dr (28)
or K =
_
b
a
F(r) dr (29)
Thats all well and good formally, but what does the last integral mean? The motion r(t)
traces out some path in space, and the integral is along the path. (See the gures below.)
Indeed, we call it a path integral or line integral. (The latter terminology is more
common. The former has a larger connotation when studying quantum mechanics.)
A couple of simple things to notice:
For a constant force F acting along a straight line path with displacement d = r, the
work done by the force is W = F d
For an object moving in a circular path, the centripetal force keeping it moving this
way does no work on the object, because F(r) dr = 0 everywhere along the path.
If the path is closed, like a circle or some other closed loop, then the line integral
done all the way around is written as
_
F dr
Practice Exercise
Find the work W =
_
F dr done around the
closed rectangular loop shown, for two forces (a)
F(x, y) = y

i, and then for (b) F(x, y) = x

i. Treat
the closed loop as four, straight line paths I, II, III,
and IV, and calculate the work done on each path.
The total work done is just the sum of the work
done over each of the four straight line segments.
The force in (b) is called conservative and the
force in (a) is called non-conservative. We will
be discussing all this a lot more.
14
Physics I Honors Monday 15 September Fall 2008
Todays Topic #1: Potential Energy; Conservation of Energy
Consider a mass m on a spring, i.e. F(x) = kx, and take the time derivative of
E =
1
2
mv
2
+
1
2
kx
2
(30)
dE
dt
= mv v + kx x = v(m v + kx) = v(F + kx) = 0 (31)
We say that E is a constant of the motion. It is conserved quantity. Note that
d(kx
2
/2)/dx = kx. In general, let F(x) = dU/dx for some function U(x). So we have
E
1
2
mv
2
+ U(x) =
dE
dt
= mv v +
dU
dt
= mv v +
dU
dx
dx
dt
= v
_
F +
dU
dx
_
= 0 (32)
We call U(x) the potential energy and E the total energy. Relating this to work,
W
ba
=
_
b
a
F(x)dx =
_
b
a
dU
dx
dx =
_
b
a
dU = [U
b
U
a
] (33)
That is, the work done from a to b is the negative of the change in potential energy.
Now generalize to three dimensions. This is simple, in principle. We just write
U
b
U
a
= U(r
b
) U(r
a
) = W
ba
=
_
b
a
F(r) dr (34)
but the line integral can (in principle) depends on the path! If it does, then Eq. 34 makes
no sense. We can only dene a potential energy for forces with path independent work, also
known as conservative forces. These let us write down an energy that is conserved.
Non-conservative forces lead to a loss of (mechanical) energy. For F = F
c
+F
nc
, we have
K
b
K
a
= W
total
ba
=
_
b
a
F dr = [U
b
U
a
] +
_
b
a
F
nc
dr =E
b
E
a
= W
nc
ba
(35)
The energy lost in going from a to b is just the work done by the non-conservative forces.
Finally, lets dene power. Power is just the rate at which work is done. That is
P
dW
dt
=
F dr
dt
= F
dr
dt
= F v (36)
Practice Exercise
An object with mass m falls at a constant (that is, terminal) velocity v through the
atmosphere. If the drag force is f = bv, nd b in terms of m, g, and v. Then, calculate (a)
the change in total mechanical energy E when the object falls through a distance h, and
(b) the work W
nc
done by the non-conservative forces. Finally, show that E = W
nc
.
15
Topic #2: Energy Diagrams
We gain insight about motion by looking more literally at energy conservation, i.e.
E =
1
2
m x
2
+ U(x) so x = v =
_
2
m
[E U(x)] (37)
Velocity can be to the left or right, but speed
( [v[) is determined. The square root leads
to turning points to the motion, which limit
the position based on the energy. See the g-
ure, and discuss motions for dierent values of
E. Relate force plot to the energy plot. Dis-
cuss behaviors of particles with various E
i
, and
mention oscillations. Draw the analogy with
up and down a hill. Note that, in princi-
ple, one can get the motion x(t) by integrating
Eq. 37. Well do this in a practice problem.
Example: The Pendulum. A simple pendulum is a mass m attached to a massless string
of length which swings in a plane through an angle . (See page 255 in your textbook.)
Analyzing this in terms of F = ma is a little tricky, but we can do it. It is two dimensional,
but the two axes change with time. One is radial (in which there is no motion) and the
other is angular along the path, a variable well call s = . We have
F
s
= mg sin = m s so

=
g

sin (38)
Nobody knows how to solve this problem analytically. However, if the pendulum never swings
through large angles, then sin and we recover the harmonic oscillator equation of
motion, with
2
= g/.
We can also look at this in terms of energy. The potential energy in terms of is
U() = mgh = mg( cos ) = mg(1 cos ) (39)
Plot this as a function of and oscillate about the minimum at = 0. (Also point out the
equilibrium point at = .) This leads into a discussion of small oscillations, i.e.
U() = mg
_
1
_
1

2
2!
+

4
4!
+
__
mg

2
2
= m
g

s
2
2
(40)
which is a harmonic oscillator potential with k replaced by mg/, i.e. k/m g/.
Practice Exercise
Use Eq. 37 with U(x) = mgx and E = mgh to nd x(t). Assume that x = h when t = 0.
(This describes the motion of something falling from rest at a height h, right?) Split dx/dt
so that you get
_
dx on the left and
_
dt on the right. It should be obvious after integrating
that the constant of integration is zero for this initial condition. Youll need the following
obscure calculus rule:
_
dx

a + bx
=
2

a + bx
b
(41)
where a and b are constants.
16
Physics I Honors Thursday 18 September Fall 2008
Todays Topic: Conservation Laws and Particle Collisions
Textbook (Sec.14.4): No new physical principles are involved; the discussion is intended to
illustrate ideas we have already discussed.
P
initial
= m
1
v
1
+ m
2
v
2
P
nal
= m
1
v

1
+ m
2
v

2
(42)
P
initial
= P
nal
(43)
m
1
v
1
+ m
2
v
2
= m
1
v

1
+ m
2
v

2
(44)
Elastic and inelastic collisions. If the (internal) forces that cause the collision are conser-
vative, then mechanical energy wil be conserved. Far away from the colliding point, there are
no forces, so no (internal) potential energies, and conservation of energy means conservation
of kinetic energy. If non-conservative forces act, then energy can be lost. We write
K
i
= K
f
+ Q i.e. Q = K
i
K
f
0 (45)
If Q = 0 then the collision was elastic. Using the same notation as above
1
2
m
1
v
2
1
+
1
2
m
2
v
2
2
=
1
2
m
1
v
2
1
+
1
2
m
2
v
2
2
+ Q (46)
Example 4.18: Elastic Collision of Two Balls. (Recall Sept.8 practice exercise.) Put
m
1
= m and m
2
= 3m, v
1
= v and v
2
= v. Conserving momentum and energy,
mv 3mv = mv

1
+ 3mv

2
(47)
1
2
mv
2
+
1
2
3mv
2
=
1
2
mv
2
1
+
1
2
3mv
2
2
(48)
Equation 47 gives
v

1
= 2v 3v

2
(49)
which we can then plug into Eq. 48 to nd
2v
2
=
1
2
(2v + 3v

2
)
2
+
1
2
3mv
2
2
(50)
= 2v
2
+ 6vv

2
+ 6v
2
2
(51)
or 0 = vv

2
+ v
2
2
(52)
17
This equation has two solutions for v

2
, each of which gives v

1
by Eq. 49, namely
v

2
= v = v

1
= +v (53)
or v

2
= 0 = v

1
= 2v (54)
The rst of these is just the initial conditions. If nothing happens when the two balls pass
near each other, then nothing happens. If they do in fact collide, however, then the second
ball stops, and the rst one res backward at twice the velocity that it started with.
Collisions and the Center of Mass. The velocity of the center of mass is
V =
m
1
v
1
+ m
2
v
2
m
1
+ m
2
(55)
Lets call this the laboratory frame. We can change our frame of reference to the center
of mass frame by imagining that we are observing the collision by moving along with the
center of mass. In this frame, the velocities of the two particles are observed to be
v
1C
= v
1
V =
m
2
m
1
+ m
2
(v
1
v
2
) (56)
v
2C
= v
2
V =
m
1
m
1
+ m
2
(v
2
v
1
) (57)
so that the momenta in the center of mass frame are given by
p
1C
= m
1
v
1C
=
m
1
m
2
m
1
+ m
2
(v
1
v
2
) = v (58)
p
2C
= m
2
v
2C
=
m
1
m
2
m
1
+ m
2
(v
2
v
1
) = v (59)
where we have made the denitions
v v
1
v
2
and
m
1
m
2
m
1
+ m
2
(60)
The total momentum p
1C
+ p
2C
= 0 in the center of mass frame, as we expect. It is easy
to show that for elastic collisions in the center of mass, v

1C
= v
1C
and v

2C
= v
2C
. Your
textbook has the details.
We call v the relative velocity and the reduced mass. You will come back to these
quantities over and over again as you study the physics of collisions, both classically and
quantum mechanically.
Practice Exercise
Suppose that the two balls in Example 4.18 stick together after the collision, i.e. v

1
= v

2
.
Calculate the Q value for the collision.
18
Physics I Honors Monday 22 September Fall 2008
First exam is this Thursday, in class, at 10am. Today, we will do some math.
A Review of Some Calculus Rules
f(x) f

(x)
_
f(x)dx f(x) f

(x)
_
f(x)dx
x
a
ax
a1 1
a+1
x
a+1 1
x
ln x
e
x
e
x
e
x
ln x
1
x

cos x sin x sin x sin x cos x cos x
Chain rule :
d
dx
f(u) =
df
du
du
dx
Product rule :
d
dx
uv = v(x)
du
dx
+ u(x)
dv
dx
(61)
Partial Derivatives, Gradient, and Force & Potential Energy in 3D
The derivative (with respect to x) of a function f(x) is just the rate at which f(x) is changing,
that is f/x when the changes are very small. We write f

(x) = df/dx, where the right


side of the equation is a much more descriptive notation.
Now suppose that f is a function of two (or more) variables x and y, that is f(x, y). We
can ask how f changes with respect to x if y is held constant, or vice versa. We write these
derivatives as f/x and f/y, respectively, and call them partial derivatives. There are
other notations, but they are all bad. In many cases df/dx and f/x can be quite dierent,
but we will not get into this kind of confusing situation in this course!
Suppose x changes by an amount dx, and y changes by an amount dy. How much does
f(x, y) change? The answer is pretty obvious, namely
df =
f
x
dx +
f
y
dy (62)
So now, consider the work dW done by a force F(r) over a distance dr. The change in
potential energy U(r) = U(x, y, x) (assuming that one can be dened) is
dU = dW = F(r) dr (63)
Write the left side using Eq. 62 and write out the right side in terms of components, so
U
x
dx +
U
y
dy +
U
z
dz = F
x
dx F
y
dy F
z
dz (64)
Clearly the x component of F(r) is just U/x and so forth. It is nice to write this in a
simple notation using vectors, by dening the gradient operator as follows:

i

x
+

j

y
+

k

z
(65)
In this case, Eq. 64 is really just a statement that
F(r) = U(r) (66)
We frequently refer to F(r) as a vector eld and U(r) as a scalar eld.
19
A Useful Theorem about the Curl of a Vector Field
!"#$%
&"
&$
Imagine a tiny, closed rectangular path in the (x, y) plane, with side
lengths dx and dy, and with the lower left corner at the point (x, y).
It is very simple to calculate the work done by a force eld F(x, y)
while going around this path, since the force is approximately con-
stant along each of the straight legs. We have
_
F dr = F(x, y)dx + F(x + dx, y)dy
F(x + dx, y + dy)dx F(x, y + dy)dy (67)
Collecting some terms, and ignoring a change in y, say, when moving in the x direction,
_
F dr = [F(x + dx, y) F(x, y + dy)] dy [F(x + dx, y + dy) F(x, y)] dx
=
_
F(x + dx, y) F(x, y + dy)
dx

F(x + dx, y + dy) F(x, y)
dy
_
dxdy
=
_
F
x

F
y
_
dxdy = [F]
z
dxdy = [F] dA (68)
We call F the curl of the eld F, and have dA dxdy and the vector dA pointed
perpendicular to the plane bounded by the tiny loop.
Now imagine an arbitrary surface A bounded by some curve C. You can tile A into a
bunch of tiny rectangles, and the loop integrals around any two adjacent rectangles cancels
on the border. In other words, the tiny loop integrals (68) end up giving the loop integral
around C. Mathematically, this says that
_
C
F(r) dr =
_
A
[F] dA (69)
This is called Stokes Theorem. Obviously, a force eld F(r) is conservative if F = 0.
This is automatically satised if F(r) = U(r).
A Dierent Useful Theorem about the Divergence of a Vector Field
Now consider a vector eld E(r) and a tiny box in three dimensional space, with sides
dx, dy, and dz. The ux out of the box near the point (x + dx, y, z) is pretty clearly
E
x
(x +dx, y, z)dydz. Play a game similar to what we did with Stokes theorem, building an
arbitrary volume V out of a bunch of tiny boxes. For two adjacent boxes, the ux out of one
cancels the ux into the other, so you end up measuring the ux through the entire (closed)
surface A that contains the volume. It isnt hard to go from here to show that
_
A
E(r) dA =
_
V
[ E] dV (70)
Physicists call this Gauss Theorem; mathematicians call it the Divergence Theorem. (The
quantity E is called the divergence of the eld E.) We will nd some important uses
of Gauss theorem this semester when we study uid behavior. Next semester, you will use
Gauss theorem in your study of electromagnetism.
20
Physics I Honors Monday 29 September Fall 2008
Hand back Exam #1. Average was 74.1. Questions? Tell me. Grades 60 see me.
Angular Momentum of a Particle
Consider an object moving in two dimen-
sions under the action of a central force
F = F
x

i + F
y

j. The gure shows that


F
y
/F
x
= y/x or F
y
yF
x
= 0
Now consider a quantity dened as
m(x y y x) (71)
This quantity is a constant of the motion:
d
dt
= m( x y + x y y x y x)
= m(x y y x) = xF
y
yF
x
= 0
If we dene L r (mv) = r p, then we see that = L
z
. We call L the angular
momentum of a particle, and it is conserved for motion in a central force eld.
It is convenient to think of L in terms of parallel and perpen-
dicular components of r. If we write r = r

+r

then
L = r p =
_
r

+r

_
p
= r

p +r

p = r

p (72)
so that = L
z
= [r

[ [p[ (= [r[ [p

[) (73)
We call [r

[ the moment arm of the angular momentum.


In 1609 Kepler told the world that planets sweep out equal
areas in equal times. Newtonian gravity is a central force,
and Keplers statement is a consequence of angular momentum
conservation. Following the gure at the right
dA
1
2
(r + dr)(rd) so
dA
dt
=
1
2
r
2
d
dt
=
1
2
r
ds
dt
(74)
But ds/dt = v

= p

/m so dA/dt = rp

/2m = L
z
/2m.
Practice Exercise
Angular momentum depends on the choice of coordinate system. Show this explicitly for
something falling straight down, starting from rest at y = 0. What is L for something falling
along the y-axis? Do it again for a line parallel to the y-axis, along the line x = A?
21
Torque
If angular momentum is constant for an object acted on by a central force, then how do we
get the angular momentum to change? By brute force, we see that
dL
dt
=
d
dt
(r p) =
dr
dt
p +r
dp
dt
= r F (75)
where we used F = dp/dt and v p = 0 with v = dr/dt and p = v. We write r F,
where is the torque. It is convenient to think of torque as force times a moment arm.
Imagine a bunch of masses hanging from a string. The torque from gravity is (y is up)

n
r
n
F
n
=

n
r
n
(m
n
g

j) =
_

n
m
n
r
n
_

_
g

j
_
= r
cm

_
g

j
_
(76)
If we hang from the center of mass, then all moment arms are measured from here and
r
cm
= 0. So, there is no torque and the object is balanced. Sometimes we call the center of
mass the center of gravity.
Practice Exercise
Calculate the torque for each of the two cases in the previous practice exercise, due to a
mass m falling under the inuence of gravity, and show that in each case, = d/dt.
Angular Momentum and Fixed Axis Rotation
So what is the motion resulting from a torque? Consider one single particle, inside the solid
object, located at r and acted on by F. It can only move in a circle of radius r and the force
in that direction is F sin . If the path length it travels is called s, then Newtons Second
Law says F sin = m s = mr

. Multiply both sides by r, and we have = (mr


2
), where
=

is called the angular acceleration. The quantity mr
2
has replaced mass when we
describe the motion in angular variables.
Since is the same for all parts of the object, and since is either from one force at one
point, or due to a sum of such things, Newtons Second Law for a solid object becomes

= I where I =

n
m
n
r
2
n
(77)
is called the rotational inertia or moment of inertia. Notice that I is a property of the solid
object and the specied axis. (Do you get the feeling that I is not really a scalar but not
a vector either? If so, you are right. It is actually something called a tensor.)
You can prove something called the Parallel Axis Theorem which states that the rotational
inertia of any body about an arbitrary axis equals the rotational inertia about a parallel axis
through the center of mass plus the total mass times the squared distance between the two
axes. Mathematically, we write I = I
cm
+ Mh
2
. See Example 6.9 in your textbook.
Rotational Inertia of Solid Objects
Of course, for solid objects, we dont do the sum over all the masses in Eq. 77 to get the
rotational inertia. Instead, we do the integral I =
_
r
2
dm. This is just like calculating the
center of mass, that is
_
r dm, but with another factor of r. See Example 6.8 in your textbook.
We will be doing more of these sorts of things, include homework problem 6.8.
22
Physics I Honors Thursday 2 October Fall 2008
Solid objects both rotate and translate. There is a lot of physics associated with these mo-
tions. See Chap.6. Rather than derive and review it all, here are the rules (Table 6.1):
Pure rotation (about a xed axis) Rotation plus translation (0 means CM)
L = I L
z
= I
0
+ (RMV)
z
= I
z
=
0
+ (RF)
z
with
0
= I
0

K =
1
2
I
2
K =
1
2
I
0

2
+
1
2
MV
2
Good examples to review include the Atwoods Machine (Example 6.10), Katers Pendulum
(6.12), and the Disk on Ice (6.15). Well do two applications in class today.
Objects Rolling Down a Plane. (Refer to Example 6.16)
Consider some object with radius b, mass M, and I
0
= Mb
2
, rolling down a ramp. How
does the shape (i.e. ) aect the time it takes for the object to get to the bottom of the
ramp?
First analyze it using just

F = ma. This says that


W sin f = Mg sin f = Ma
Friction makes it roll, with torque bf, so
bf = I
0
= (Mb
2
)(a/b) = Mab
Solving, a = g sin /(1 + ) (78)
A dierent approach uses torque and angular momentum.

z
=
0
+ (RF)
z
= bf + b(f W sin ) = bMg sin
(The normal force cancels W cos so no net torque.)
L
z
= I
0
+(RMV)
z
= Mb
2
bMV = (1+)Mb
2

Then put =

L
z
= (1+)Mb
2
= (1+)Mba and get the same answer as before.
So, a larger of gives a smaller acceleration and takes longer to get down the ramp.
A third way to look at the problem is in terms of energy. The potential energy at the start
is U
i
= Wh = Mgh and there is no kinetic energy. At the bottom of the ramp, the CM has
speed V so K
f
=
1
2
I
0

2
+
1
2
MV
2
=
1
2
(Mb
2
)(V/b)
2
+
1
2
MV
2
=
1
2
(1 + )MV
2
. Then,
U
i
= K
f
= Mgh =
1
2
(1 + )MV
2
= V =

2gh
1 +
(79)
and, again, large travels more slowly at the bottom. (Friction is irrelevant here. Why?)
Practice Exercise
Use one dimensional kinematic relations, for something moving through a distance d with
acceleration a, to show that Eq. 79 is the same as Eq. 78.
23
Kinematics of an Ultraelastic Rough Ball
See Richard L. Garwin, American Journal of Physics 37(1969)88. It is posted on our website.
(Discuss how to look up papers in e-Journals using http://library.rpi.edu/)
Play with the ball a little, as in Figure One and then as in Figure Three.
The analysis is based on two important assumptions about the ball and its collisions:
(a) The collision is elastic, conserving the sum rotational+translational kinetic energy
(b) There is no slip in contact between the ball and a surface, i.e. the ball is rough
The gure shows the coordinate system and nomenclature
used in the paper. We use b for before and a for
after the collision. The abbreviation C = R is used for
brevity, where R is the radius of the ball. The moment of
inertia (about the CM) of the ball is I
0
= MR
2
with =
2/5. (Garwin uses instead of , but I switched so that there
is no confusion with our notation for angular acceleration.)
The kinetic energy of the ball is just
K =
1
2
I
0

2
+
1
2
MV
2
=
1
2
MR
2

2
+
1
2
MV
2
=
1
2
M(V
2
+ C
2
) (80)
so K
b
= K
a
. The angular momentum about the contact point (see the gure) is just
L = I
0
R(MV ) = MR
2
MV R = MR(C V ) (81)
and since all of the forces in the collision pass through the contact point, L
b
= L
a
. This im-
mediately gives a simple result, since Eqs. 80 and 81 imply, with conservation applied,
(C
2
b
C
2
a
) = V
2
a
V
2
b
and (C
b
C
a
) = (V
a
V
b
) (82)
These equations are divided to yield
C
b
+ C
a
= (V
a
+ V
b
) or C
a
+ V
a
= (C
b
+ V
b
) (83)
Garwin calls V + C the parallel velocity, the velocity tangent to the balls surface at the
contact point. We have shown that this velocity is conserved in magnitude, but reverses in
direction after a collision with the wall or the oor. Subracting Eq. 83 from 82 yields
C
a
=
1
+ 1
C
b

2
+ 1
V
b
and then V
a
=
2
+ 1
C
b

1
+ 1
V
b
(84)
Lets apply this to a super ball bouncing o the oor. Pause for a moment to realize that
the velocity V

perpendicular to the oor is conserved in the collision. (Why?) If a ball falls


vertically downward, then V
b
= 0 and V
a
= 2/( + 1)C
b
= (4/7)C
b
for = 2/5, i.e. a
solid sphere. That is, it bounces to the side with a transverse velocity that depends on how
fast it is spinning.
If on the other hand, a ball that is not spinning (C
b
= 0) bounces at a 45

angle so that
V
b
= V

, then V
a
= (1 )/(1 + )V
b
= (3/7)V

. That is, it bounces up at an angle (with


respect to the vertical)
a
= tan
1
(V
a
/V

) = tan
1
(3/7) = 23

< 45

. Try it and see!


See Garwins paper for other calculations, including the trajectory for bouncing o the
underside of a table.
24
Physics I Honors Monday 6 October Fall 2008
Today we recognize angular velocity , angular momentum L, and torque as full-edged
vector objects. The physical implications are interesting.
What exactly is a vector? Not angular position!
I have told you that vectors are just bookkeeping but thats not really true. To be useful
in physics, they need to have some obvious (?) mathematical properties. In fact, if I try to
specify the angular position of an object with the construct
?
=
x

i +
y

j +
z

k where
x
,

y
, and
z
are angles with respect to xed axes, then I will get into trouble quickly.
Show the trick with the textbook. Rotations do not commute but vector addition must!
However, innitesimal rotations do commute. (See Note 7.1 in K&K.) The trick is that the
part that that does not commute enters in second order in the angles. So I can write
=
d
x
dt

i +
d
y
dt

j +
d
z
dt

k =
x

i +
y

j +
z

k (85)
for angular velocity. For its direction, see the following gures of a rotating rigid body:
Pick some particle in the object. Its posi-
tion is r and its velocity is v = dr/dt. The
magnitude of v is (r sin )d/dt. In other
words, we write
dr
dt
= v = n r
d
dt
= r (86)
where we dene (d/dt) n. In particu-
lar, points in the direciton of the rotation
axis, specied by the unit vector n.
Example: Angular Momentum of a Rotating Skew Rod
Find the angular momentum L of the two masses around the
rod midpoint. By denition, L =

(r
i
p
i
). For each mass
r p and r p is perpendicular to the rod! Since r p we
have [r p[ = lp = lmv = lml cos so L = 2ml
2
cos .
For this axis I =

mr
2
= 2m(l cos )
2
so L ,= I! The
equality is violated in both magnitude and direction, although
L
z
= Lcos = I. This is because the object is asymmetric
and our non-tensor formulation of rotational inertia is naive.
Practice Exercise
Suppose a particle rotates counter clockwise in the xy plane. Let be the polar angle,
relative to the x-axis. Then, its angular velocity would be =

k where =

. Show that
equation 86 gives the correct vector velocity v = (x

j y

i) = r

. Draw a picture with the


particle at a couple of points, and sketch r and v.
25
Example: Torque on the Rotating Skew Rod
The (direction of the) angular momentum of the skew rod changes with time. What torque
causes this? To see this, lets nd dL/dt and interpret the result. We can do
L(t) = Lsin cos t

i + Lsin sin t

j + Lcos

k (87)
=
dL
dt
= Lsin
_

i sin t +

j cos t
_
so = Lsin (88)
Alternatively, we can use a geometric approach. Looking down
the axis from the top, on the horizontal (h) plane, we see that
[L
h
[ = L
h
. This is the only component of L that changes.
Therefore
=
dL
dt
=
dL
h
dt
= L
h
d
dt
= (Lsin )() (89)
so we get the same result as before. The torque is zero if = 0, and also if = 90

(L = 0).
The torque is from forces which hold the vertical bearing onto the rotation axis. Can you see
why the torque zero in both cases? (Imagine youre holding the bearing in your hand.)
The Gyroscope
A simple gyroscope is a spinning wheel on a horizontal axis, supported on one end only:
Contrary to our intuition, it does not fall down. Instead, gravity exerts a torque which causes
it to precess about a vertical axis which passes through the support point.
Formally, this is easy to see. The angular momentum vector L is along the spinning axis. The
torque about the support point is only the weight W, since the normal force Nhas no moment
arm about this point. The torque has magnitude = lW and direction perpendicular to
both W and L. So L changes in direction, not magnitude, tracing the circle in the horizontal
plane, shown on the left. Calculating the precession angular frequency
= lW =
dL
dt
= L so =
lW
L
=
lW
I
0

s
(90)
A simple model makes it easier to see, physically, why the gyroscope
precesses. Consider two masses on an arm, rotating in the horizontal
plane. Instantaneously apply a torque with opposing forces F up and
down on the two masses, so there is no net force and the center of mass
doesnt move. If the force acts for time t, then p = mv = Ft
for each mass. One mass moves up and the other moves down. Thus

v
v
=
Ft
mv
=
2lFt
2lmv
=

L
t so
d
dt
=

L
(91)
which is the same result that we got in Equation 90.
26
Physics I Honors Thursday 9 October Fall 2008
Dierent format today. We will spend most of the period doing worksheets which you will
turn in. These will be counted together as one homework assignment. The point is to help
you get up to speed on the material weve been covering, especially the use of vectors to
describe motion, both translation and rotation, in two and three spatial dimensions.
Worksheet #1: Angular velocity for motion in a plane.
Collect the worksheets, and work the problem out on the board.
Rigid Bodies and Conservation of Angular Momentum
For a symmetric rigid body, we now know that the angular momen-
tum vector is parallel to the angular velocity vector. We write
L = I where I =

m
i
r
2
i
is the rotational moment of inertia about the rotation axis. We say
that L is conserved if there are no external torques but nobody can
prove this is so, starting from Newtons laws! See the diagrams on
the right. The torques
ij
= r
i
f
ij
cancel only if f
ij
works along the
line r
i
r
j
. That is, f
ij
= f
ji
(Newtons Third Law) is not enough.
There is some higher reason for angular momentum conservation.
Worksheet #2: Conservation of Angular Momentum
Collect the worksheets, and work the problem out on the board.
General Angular Momentum of Rigid Bodies: The Rotational Inertia Tensor
If all position vectors r
i
, for the particles of mass m
i
within a rigid body, are measured with
respect to the center of mass, then the angular momentum vector is
L =

i
r p =

i
r m
i
r
i
=

i
m
i
r
i
( r
i
)
You can reduce this expression with some tedium. (See K&K.) Youll nd, for example,
L
z
=
_

i
m
i
(x
2
i
+ y
2
i
)
_

i
m
i
x
i
z
i
_

i
m
i
y
i
z
i
_

y
or L = I (92)
where I is the rotational inertia tensor, a 3 3 matrix of numbers. This is a general formu-
lation of the angular momentum of a rigid body. Note that if =

k then L
z
= I
z
where
I =

mr
2
just as we had before. Also, as we have seen, if the o diagonal elements of I
are nonzero, then there are other terms in L. (How much do you guys know about matrices?
Can I tell you that the axes which diagonalize I are called the principle axes?)
Worksheet #3: The Rotational Inertia Tensor for the Skew Rod
Collect the worksheets, and work the problem out on the board, if there is time.
27
Name:
Physics I Honors Thursday 9 October Fall 2008
Worksheet #1: Angular velocity for motion in a plane
Consider a particle which moves in the (x, y) plane. Answer the following questions:
A. In one sentence, why does r(t) = x(t)

i + y(t)

j tell us the position of the particle?


B. The particle moves in a circle with angular velocity . It starts at (x, y) = (r, 0). Find
the functions x(t) and y(t) in terms of r, , and t.
C. Find the velocity vector v(t) = r(t) in terms of r, , and t, and unit vectors

i and

j.
D. The angular velocity is =

k. Take the cross product r(t) and show that it is the


same as v(t). Refer to Sec. 1.4 in your book for the cross product of the unit vectors.
28
Name:
Physics I Honors Thursday 9 October Fall 2008
Worksheet #2: Conservation of Angular Momentum for Rigid Bodies
A student sits in a swiveling chair that allows her to rotate about a vertical axis. Her
rotational inertia is I
0
. She is holding a wheel on an axle. The wheel has rotational inertia
I
W
. The axle is oriented vertically. Initially, the student is not moving, but the wheel is
spinning with an angular velocity , counter clockwise as viewed from the top.
A. What is the magnitude and direction of the initial angular momentum vector L
i
?
B. The student turns the wheel and axle over by 180

so that it again is oriented vertically,


but now the wheel rotates clockwise as seen from the top. If the students angular velocity
is now , write the nal angular momentum L
f
in terms of I
0
, I
W
, , and . Be careful of
signs. You can ignore the mass of the wheel as compared to the mass of the student.
C. Find in terms of I
0
, I
W
, and .
29
Name:
Physics I Honors Thursday 9 October Fall 2008
Worksheet #3: The Rotational Inertia Tensor for the Skew Rod
Use Eq. 92 to calculate the angular momentum of the skew rod
which we studied last class. The angular velocity is =

k.
A. Using as dened at the left, derive L
z
in terms of m, , and .
B. By rearranging x, y, and z in Eq. 92, show that
L
x
=
_

i
m
i
x
i
z
i
_
and L
y
=
_

i
m
i
y
i
z
i
_

C. Complete the following table for the coordinates of the two masses in terms of , , and h:
Particle 1 Particle 2
x
1
= cos t x
2
=
y
1
= y
2
= sin t
z
1
= z
2
= h
D. Derive L
x
and L
y
in terms of m, , h, and . Compare to our result from last class.
30
Physics I Honors Tuesday 14 October Fall 2008
Today we nally (!?) leave rotations of rigid bodies, but we will use those topics later.
Accelerating Reference Frames
Newtons rst law of motion says something like An object at rest stays at rest unless some
force acts on it. Also, an object in motion doesnt change its direction or speed unless some
force acts on it. These statements are fundamental assumptions and cannot be proven. In
fact, they arent always true. We say that an inertial reference frame is one in which
Newtons rst law, sometimes called the law of intertia, holds.
A reference frame in which Newtons law doesnt hold is an accelerating reference frame.
The second law really embodies the rst law. The second law says that

F = mA (93)
so that if there are no forces (

F = 0) then A = 0, i.e. the rst law. If you are inside the


frame that is accelerating, then you feel something thats like a force, but isnt really there.
So, invent a ctitious force F
ct
= mA and everything is ne. Heres a simple example:
Analyze this two ways, for an observer standing on the
street, and for someone inside the car.
Outside (inertial) Inside (accelerating)
F
ct
0 mA

F
x
T sin T sin + F
ct

F
y
T cos mg T cos mg
So, the person outside the car says that T sin = mA, while the person inside the car says
that T sin + F
ct
= T sin mA = 0. These are the same equations. This is really pretty
trivial, but more complex situations are more complicated.
Practice Exercise
Heres a slightly more complex situation, from Example 8.2 in your textbook.
The plank accelerates to the right with acceleration A. The
cylinder rolls without slipping on the plank. To an observer
in an inertial frame, the cylinder accelerates to the right due
to the frictional force f. To an observer on the plank, a
ctional force F
ct
= MA acts on the center of mass and
accelerates the cylinder to the left. If the acceleration of the
cylinder is a

to the observer on the plank, then the torque


fR accelerates the cylinder by

via fR = I
0

= I
0
a

/R
where I
0
= MR
2
/2.
So, rst, write x = x

+ X to locate the cylinder in the inertial frame relative to the accel-


erating frame, where X locates the plank. Use this to explain why the acceleration in the
inertial frame is a = a

+ A. Then, use the relationships between rolling and translational


motion above, to nd the accelerations a and a

in terms of A.
31
Galilean Relativity
No lets be a little more formal about this idea of reference frames. In particular, we will
show that Newtons laws of motion are the same in dierent inertial reference frames, that
is, two frames that are moving with constant velocity with respect to each other.
For this discussion, I am using the notation and gures from Sec. 11.4 in your textbook, as
opposed to Sec. 8.2. Both are called The Galiliean Transformations, but the latter is a
more common formulation.
Two observers follow motion of the same particle. It is lo-
cated at r in the coordinate system of one of them, and at
r

in the coordinate system of the other. The origin of the


primed system is located at R in the unprimed system.
We therefore have
r

= r R where R = vt (94)
That is, the primed coordinate system moves with a (constant) velocity v relative to the
unprimed system. We let v be along the x (and x

) axis.
Obviously r

= r v and r

= r. Therefore, even though the velocity of the particle is


shifted by v in one coordinate system relative to the other, the accelerations are the same.
This means that

F = ma in both systems, so Newtons second law is intact. In fact, you


cannot tell which frame you are in by studying the particles motion. Notice also that, for
two particles a and b, r

a
r

b
= r
a
r
b
. This means that the distances and directions between
the two particles are the same in both frames, and so are two-body forces like gravity.
(This all makes some plausible, albeit ultimately wrong, assumptions. For example, it as-
sumes that lengths are measured the same by both observers, but this ends up being incorrect.
Leave this for now, but remember it when you come to study special relativity.)
Now assume that some sort of pulse is emitted at t = 0 from
the origin in the (x, y) system. Its position along the x axis
in the (x, y) system is
x = ct
where Ive assumed that its velocity is c. Its position along
the x

axis in the (x

, y

) system is therefore, from Eq. 94,


x

= x vt = ct vt = (c v)t
In other words, if you follow the pulse in the (x

, y

) system, then it appears to move in the x

direction with a velocity dx

/dt = c v. Of course, this is just what youd expect. However,


as you will learn next semester, the laws of electromagnetism (aka Maxwells Equations)
say that the speed of light is c in any inertial frame. This inconsistency is what led to the
development of special relativity, and the eventual recognition that Newtons laws are only
an approximation in a world where all velocities are small compared to c.
Practice Exercise
Show that the force of gravity is the same between two objects a and b, regardless of whether
they are viewed in a primed or unprimed system, related by the transformation Eq. 94.
32
Physics I Honors Thursday 16 October Fall 2008
Thanks to Brian for covering class today (and lab yesterday!)
Remember: Exam #2 is two weeks from today.
Note: Section 8.4 in book on Equivalence Principle is interesting, but were skipping it.
Rotating Coordinate Systems
Recall from last class: Frame accelerating with acceleration A doesnt obey Newtons laws,
but you can x things by inventing a ctitious force F
ct
= mA.
Familiar example: For a turntable rotating at constant angular velocity , there is an inward
acceleration
2
R where R is the distance from the rotation axis. So, if you are on the
turntable you feel a ctitious force m
2
R outward. This is called centrifugal force.
Our goal today is to be more formal with this concept of rotating coordinate systems.
See the gure. One set or coordinates rotates
about the other with angular velocity . The
origins coincide, that is, we are only concerned
with rotation, not translation. The (x, y, z)
system is inertial. At time t all three axes
coincide. Rotation is about the z (z

) axis.
Our job is to relate the derivatives of the vector r(t) in the inertial (unprimed) and rotating
(primed) frames. Examine the change in this vector between times t and t + t, below:
In short, the rotating observer is fooled. After the time t has elapsed, he measures a
change r

based on the correct location of the nal position, but a wrong starting position
because his axes have rotated. The gure makes it clear that
r(t) = [r

(t) + r

(t)] r(t) (95)


We know how to relate r

(t) and r(t). They are connected through the angular velocity
vector =

k according to Equation 86, that is


r

(t) r(t) = (r)t (96)


If we combine equations 95 and 96 we can write
r(t)
t
=
r

(t)
t
+r (97)
33
This is essential. If we just let t 0, we can relate the velocities in the two frames as
v
inertial
= v
rotating
+r (98)
This proof relies only on the geometric properties of r, and really relates the derivative of this
vector when viewed by two dierent coordinate systems. So, we could just as well write
_
dr
dt
_
in
=
_
dr
dt
_
rot
+r or, generally,
_
dB
dt
_
in
=
_
dB
dt
_
rot
+B (99)
for any vector B. (See also Note 8.2.) For example, the acceleration vector becomes
a
in
=
_
dv
in
dt
_
in
=
_
dv
in
dt
_
rot
+v
in
=
_
d
dt
(v
rot
+r)
_
rot
+(v
rot
+r)
= a
rot
+ 2v
rot
+(r) (100)
where we have assumed that the rotation vector does not change with time. Thus, the
ctitious force in the rotating coordinate system
F
ct
= 2mv
rot
m(r) (101)
has two terms. The second term has magnitude m
2
r for a position r in the rotation plane,
so is just our old friend the centrifugal force. The rst term is new. It depends on the
velocity v
rot
in the rotating frame, and is called the coriolis force.
Example 8.7. Bead sliding without friction on a radial rod,
rotating with angular velocity . Use the rotating coordinate
system. The vector r is in the horizontal plane, upward
in the lower gure. So, F
cent
= m ( r) is outward,
and F
cor
= 2mv
rot
is as shown for v
rot
radially outward.
Radial

F = ma so m
2
r = m r
Transverse

F = ma so N 2m r = 0
The radial equation gives r(t) = Ae
t
+ Be
t
where A and B are determined from initial
conditions. Thus N = 2m
2
(Ae
t
Be
t
) for the normal force of the rod on the bead.
Practice Exercise
Get some practice with the directions of coriolis and centrifugal forces on the Earths surface.
What is the direction of ? What is the magnitude and direction of r for a particle
at the equator? What if the particle is at the North Pole? Do the same for ( r).
Now calculate, in terms of and the radius R of the Earth, the magnitude of the centrifugal
force on a particle at the equator and at the pole.
Next, instead of being stationary, let the particle fall from some height near the Earths
surface. It therefore has a velocity v
rot
downward in the rotating frame. Go through a
similar process, and determine the magnitude and direction of the coriolis force on the
falling particle, in terms of v
rot
, , and R.
34
Physics I Honors Monday 20 October Fall 2008
Motion under Central Forces
Recall our discussion of two-body motion and center of mass. We found that we can treat a
two body problem as a one-body problem with a reduced mass. Lets review that.
First, write Newtons Second Law for each of the two bodies:
m
1
r
1
= f(r)r and m
2
r
2
= f(r)r
Add these two equations to see what happens to the center of mass:
m
1
r
1
+ m
2
r
2
= M

R = 0
where M = m
1
+m
2
and R = (m
1
r
1
+m
2
r
2
)/M is the position of the center of mass. On the
other hand, the two equations of motion can also be combined by subtracting, hence
r = r
1
r
2
=
_
1
m
1
+
1
m
2
_
f(r)r so r = f(r)r where =
m
1
m
2
M
is called the reduced mass. We now can deal with the one-body central force problem.
Note that if m
2
m
1
(like Sun & planet) then M m
2
and m
1
, so we ignore it.
Recall that angular momentum L = r r is conserved, in both magnitude and direction,
for central forces. One obvious conclusion is that the motion must be in a plane.
We can go further by using polar coordinates. The mag-
nitude of the angular momentum is = r
2

. The total
mechanical energy is
E =
1
2
v
2
+ U(r) =
1
2
( r
2
+ r
2

2
) + U(r) (102)
where U(r) is the potential energy with f(r) = dU/dr. So,
E =
1
2

_
r
2
+

2

2
r
2
_
+ U(r) =
1
2
r
2
+ U
e
(r) where U
e
(r) =

2
2r
2
+ U(r) (103)
is called the eective potential energy. It is useful for analyzing central force motion in terms
of the radial coordinate. The coordinate obeys d/dt = /r
2
, where is constant.
Eective potential for the potential energy of two gravitating
bodies. The solid curve is
U
e
(r) =

2
2r
2

Gm
1
m
2
r
(104)
For E = E
min
, the radial coordinate doesnt change. For
E < 0, the radial coordinate oscillates between min and
max values. For E 0, there is no maximum. Later!
Practice Exercise
Put m
1
= m M = m
2
in Eq. 104. Show that = mvR is the correct circular orbit.
35
Dark Matter in Galaxies
We can learn a lot about one of the most important problems in physics today, just by
combining what we know now about circular orbits, Newtonian gravity, and observations of
the Milky Way and some distant galaxies.
A circular orbit of some small mass m about some large mass M obeys Newtons second law
in the form
G
mM
r
2
= m
v
2
r
which gives v =
_
GM
r
for the orbital velocity. Note that the period of revolution is T = 2r/v so we have
T
2
=
4
2
GM
r
3
which is called Keplers Law of Periods. (Well look at this more on Thursday.) Note that this
is a way to determine the mass of the sun from the revolution periods of the planets.
One very basic result from this is that the tangential rotation speed v should be proportional
to 1/

r as soon as the mass is concentrated at distances smaller than r. Galaxies should


look like this, since there is generally a big bulge of bright stars near the center. However,
the galactic rotation curves look very dierent! (See slides.) This is one of the strong
pieces of evidence for dark matter in the universe.
For some interesting reading, see Alternatives to dark matter and dark energy. Philip D.
Mannheim (UConn). Published in Prog.Part.Nucl.Phys.56:340-445,2006. You can also visit
http://arxiv.org/ and get e-Print: astro-ph/0505266.
Practice Exercise
Use the Milky Way galactic rotation curve to estimate the mass of the galaxy contained
within a distance of 16 kpc. (Note that 1 kpc=10
3
pc=3.1 10
19
m.) The mass of the sun is
210
30
kg. How many sun-like stars do you need to get this mass? Note that the brightness
of a typical spiral galaxy (like the Milky Way?) seems to be about 10
10
times that of the
sun. (Table 21-1 in Introductory Astronomy and Astrophysics by Zeilik and Gregory.)
36
The Milky Way Galaxy
Visible Light
Infrared Light
37
1
9
8
5
A
p
J
.
.
.
2
9
5
.
.
4
2
2
C
N
o
t
e
:

T
h
e

v
e
r
t
i
c
a
l

a
x
i
s

h
a
s

a

s
u
p
p
r
e
s
s
e
d

z
e
r
o
!

K
e
p
l
e
r
i
a
n


p
r
e
d
i
c
t
i
o
n

f
o
r

1
0
1
0

S
o
l
a
r

M
a
s
s
e
s
38
39
Physics I Honors Thursday 23 October Fall 2008
Remember: Exam #2 is one weeks from today. Same ground rules as before.
Motion under Newtonian Gravity
We want to nd the motion under Newtonian gravity in three dimensions. Assume an object
(say, a planet) with mass m orbiting a (massive) body M. (Dont worry about reduced mass
distinction now.) The potential energy is (See Example 5.3 in your textbook)
U(r) = G
mM
r
=
C
r
where C GmM (105)
We use C because that makes it easy to generalize to other central force laws.
First lets review from last class. Motion is in a plane since L = mr r is conserved for
central forces. This is easy to show using Newtons second law:
dL
dt
= m r r + mr r = 0 +r
_

F
_
= r f(r)r = 0 (106)
where the sum of forces takes the form that denes a central force.
Central forces are conservative. (See Example 4.8 in your textbook.) So write f(r) = dU/dr
for some U(r). We know that this means that the energy
E =
1
2
m r
2
+ U(r) =
1
2
m
_
r
2
+ r
2

2
_
+ U(r) (107)
doesnt change with time. It is a constant of the motion. Note that weve written this in
terms of polar coordinates r and . See Section 1.9 of your textbook for a review.
The magnitude = [L[ of the angular momentum is also a constant of the motion. The
component of velocity perpendicular to the moment arm r is v

= r

by denition. So
= mr
2

(108)
Once again, for as long as there is motion, the value of does not change, even though r
and

are of course functions of time. They just always combine so that is constant.
Now we can nd the shape of the orbits. We can combine Equations 107 and 108 to
form two dierential equations describing the motion as r(t) and (t), namely
d
dt
=

mr
2
and
dr
dt
=
_
2
m
[E U
e
(r)] where U
e
(r) U(r) +

2
2mr
2
(109)
However, the shape of the orbit is some function r(). So, divide the two to nd
d
dr
=

mr
2
1
_
(2/m)[E U
e
(r)]
=

r(2mEr
2
+ 2mCr
2
)
1/2
(110)
The integral looks messy, but in fact it can be done analytically. (You can nd it in tables.)
Integrating both sides, and putting the constant of integration,
0
, on the left, we have

0
= sin
1
_
mCr
2
r

m
2
C
2
+ 2mE
2
_
(111)
40
Divide through my mC in the parentheses, take the sine of both sides, and solve for r:
r = r() =

2
/mC
1
_
1 + (2E
2
/mC
2
) sin(
0
)
(112)
Lets clean this up. Pick
0
= /2 (its arbitrary) and dene
r
0

2
/mC and
_
1 +
2E
2
mC
2
> 0 so r = r() =
r
0
1 cos
(113)
The parameter is called the eccentricity. The curves r() are called conic sections. The
forms become familiar if we work in cartesian coordinates x = r cos and y = r sin , so
_
x
2
+ y
2
x = r
0
= (1
2
)x
2
2r
0
x + y
2
= r
2
0
(114)
Case = 0. The curve is r = r
0
, a circle. Setting mv
2
/r = p
2
/mr =
2
/mr
3
= C/r
2
gives
r =
2
/mC = r
0
. This is precisely what we know we should get for a circular orbit.
Case = 1, i.e. E = 0. In cartesian form, we see that the orbit is a parabola,
x =
y
2
2r
0

r
0
2
(115)
Case < 1, i.e. 1
2
> 0. Your book works this out in polar coordinates, but lets try it
in cartesian. First nd the ends of the orbit by setting y = 0. This leads to
x
min
=
r
0
1 +
and x
max
=
r
0
1
= a
1
2
(x
max
x
min
) =
r
0
1
2
(116)
Now translate to x

= xx
0
where x
0
= a+x
min
= r
0
/(1
2
) is the halfway point between
the two ends. Substituting for x in Equation 114, and simplifying, we nd
(1
2
)x

2
+ y
2
=
r
2
0
1
2
b
2
(117)
Clearly y = b are the top and bottom of the orbit. Since a
2
= b
2
/(1
2
) we have
x

2
a
2
+
y
2
b
2
= 1 (118)
which is the familiar equation of an ellipse. The parameters a and b are called the semimajor
and semiminor axes. The position x = 0 is called the focus; it is not the center!
Case > 1, i.e.
2
1 > 0. Multiply Equation 117 through by 1, but keep the sign in
front of b
2
so that it is still positive. Equation 118 becomes
x

2
a
2

y
2
b
2
= 1 (119)
This is the equation of a hyperbola. Note that x
0
is now a negative number.
This is enough for today. You should read through your book on Keplers laws, but we wont
specically cover them in this course. They will be important, though, if you should take a
course on astrophysics.
41
Name:
Physics I Honors Thursday 23 October Fall 2008
Worksheet: The following shows a circular orbit of a planet around the Sun.
a. Draw on this graph a circular orbit with larger total energy.
b. Draw on this graph an elliptical orbit with the same total energy as the original circular
orbit. Make the eccentricity e of the orbit equal to 4/5. (Hint: Use the denitions of r
0
, ,
and a to show that the energy E only depends on C and a.) Calculations:
c. At a point where the elliptical and (original) circular orbit cross, draw the velocity vectors
for each of the two orbits at that point. Indicate which is which.
d. Explain briey why the magnitudes of the two velocity vectors must be the same.
e. Of the original circular orbit and your elliptical orbit, which has the largest angular
momentum? (Assuming the same orbiting mass in each case.) Explain your answer.
42
Physics I Honors Monday 27 October Fall 2008
Second exam is on Thursday. Everything the same as for Exam #1.
The Simple Harmonic Oscillator (Again)
First, start with a review of old stu. Pay attention to the notation, e.g.
0
.
Newtons Second Law says

F = ma = m x. For this
example,

F = F
spring
= kx. So we need to solve
kx = m x = x(t) =
2
0
x(t) where
0

_
k
m
(120)
Any function x(t) proportional to cos
0
t or sin
0
t solves this. So, general solution is
x(t) = Bsin
0
t + C cos
0
t (121)
We need more information to determine B and C. Typically, we use initial conditions.
Everybody knows that x and x(t) mean exactly the same thing, right?
Solution using complex numbers. We solved Eq. 120 by recognition. A better way is
to use an ansatz, namely x(t) = Ae
it
. Substitute into Eq. 120 and solve for :

2
Ae
it
=
2
0
Ae
it
= =
0
(122)
Once again, there are two possible solutions, and so once again the general solutions is:
x(t) = A
1
e
i
0
t
+ A
2
e
i
0
t
(123)
Initial conditions. Work with solution 121. Obviously x(0) = C. We also have
v(t) = x(t) =
0
Bcos
0
t
0
C sin
0
t = v(0) =
0
B or B = v(0)/
0
(124)
A trig identity gives us a dierent, more useful way, to write the solution. We write
x(t) = Acos(
0
t + ) = Acos cos
0
t Asin sin
0
t (125)
which we compare to 121 to see that
B = Asin and C = Acos = A =

x(0)
2
+
v(0)
2

2
0
and tan =
v(0)

0
x(0)
(126)
Energy. Have U(x) = kx
2
/2 since dU/dx = F
spring
. Also K = mv
2
/2 = m x
2
/2, so
E = K + U =
1
2
m
2
0
A
2
sin
2
(
0
t + ) +
1
2
kA
2
cos
2
(
0
t + ) =
1
2
kA
2
(127)
So, the total energy is the potential energy when the position x = A, the amplitude.
Practice Exercise
Apply the initial conditions to solution 123. Solve for A
1
and A
2
and then show that the
complete solution is the same as (121) with B and C in terms of x(0) and v(0). Recall
Eulers formula, namely
e
i
= cos + i sin
43
The Damped Harmonic Oscillator
Now

F = F
spring
+F
damp
= kxbv = kxb x. The equation of motion becomes
m x = kx b x = x + x +
2
0
x = 0 where
b
m
and
0

_
k
m
(128)
(The denition of
0
is just as before.) Solving this by recognition is too hard, so we
immediately move on to our ansatz x(t) = Ae
it
. Inserting this into Eq. 128 gives

2
+ i +
2
0
= 0 = = i

2

_

2
0


2
4
(129)
Case 1: < 2
0
. Dene
1

_

2
0

2
/4 and the solution to Eq. 128 becomes
x(t) = Ae
(/2)t
cos(
1
t + ) (130)
Some comments are in order. First, the oscillation frequency is
1
<
0
. Secondly, the
oscillation maximum is not a xed amplitude, but decays away with a 1/e time constant
(/2)
1
= 2m/b. Finally, the exponential factor in front makes the association of A and
with x(0) and v(0) kind of messy. This isnt so interesting, so we wont worry about it.
0 2 4 6 8 10
8
6
4
2
0
2
4
6
8
10
Time t
x
(
t
)
The red plot shows a damped oscillator with =
0
/4, with
the dashed line showing the exponential factor alone. It is
clear that the energy decreases with time. In fact, in the
case where
0
(a situation called lightly damped) it
is not hard to show that
E(t) = E
0
e
t
where E
0
=
1
2
kA
2
(131)
So, 1/ is the energy decay time constant.
Case 2: > 2
0
. In this case, we write the solutions to Eq. 129 as = i where
=

2


2

1
4
2
0

2
=
1
,
2
= x(t) = C
1
e

1
t
+ C
2
e

2
t
(132)
where both
1
and
2
are positive real numbers, so both terms of the solution are decaying
exponential functions. The blue curve in the gure is for = 3
0
, and the same initial
conditions as for the underdamped case. This situation is called overdamped.
Case 3: = 2
0
. This case is called critical damping. Something is weird, though. It
looks like our ansatz has broken down, because we only have one solution, = /2, when
we know there should be two. But dont worry, math is good. See your textbook, or a book
on dierential equations. The second solution in this case is proportional to te
t
, so
x(t) = D
1
e
t
+ D
2
te
t
(133)
is the general solution. This is the green line plotted in the gure.
Practice Exercise
Derive Eq. 131 for the decay time of a lightly damped harmonic oscillator. Follow the discus-
sion in your textbook, pages 416 and 417. Our Eq. 131 is the books equation 10.16.
44
Physics I Honors Monday 3 November Fall 2008
Hand back exam? If so, give grade statistics. Some Notes:
Wednesday is make-up lab day; need email to me or Scott by the end of today
Preliminary lab notebooks due in class next Monday!
Stay tuned: Testing something new, perhaps, in lab on coupled oscillations (Nov.19)
Review: Simple and Damped Harmonic Oscillations
The simple harmonic oscillator is based on

F = kx and has the equation of motion


x(t) =
2
0
x(t) (134)
where
0

_
k/m. We solve this with the ansatz x(t) = Ae
it
. Substituting,

2
=
2
0
= =
0
(135)
So, both x(t) = A
1
e
i
0
t
and x(t) = A
2
e
i
0
t
are valid solutions. In fact, any linear combina-
tion of the two is a valid solution. We nd A
1
and A
2
by applying the initial conditions.
The damped harmonic oscillator is based on

F = kx b x and has the equation of


motion
x(t) =
2
0
x(t) x(t) (136)
where b/m. We again solve this with x(t) = Ae
it
. Substituting, we nd a quadratic
characteristic equation for . The solutions to this equation are
= i

2

1
where
1

_

2
0


2
4
(137)
(This is the underdamped solution, where < 2
0
, which is the only case which yields
harmonic oscillations.) So, both x(t) = A
1
e
t/2
e
i
1
t
and x(t) = A
2
e
t/2
e
i
1
t
are valid
solutions. Again, any linear combination of the two is a valid solution, and we nd A
1
and
A
2
by applying the initial conditions. For future reference,
x(t) = A
1
e
t/2
e
i
1
t
+ A
2
e
t/2
e
i
1
t
(138)
is the general solution to the equation of motion for a mass m attached to a spring with force
constant k and being damped out with a force proportional to velocity, with coecient b; the
parameters
1
and are dened in terms of m, k, and b by the relations given above.
Note that we could rewrite Eq. 138 simply by using e
i
1
t
= cos
1
t sin
1
t and then
dening coecients B
1
and B
2
appropriately in terms of A
1
and A
2
so that
x(t) = B
1
e
t/2
cos
1
t + B
2
e
t/2
sin
1
t (139)
= Ae
t/2
cos(
1
t + ) (140)
This is exactly the same result which we got in Eq. 130 from last class.
45
The Forced Harmonic Oscillator
Now imagine that we are forcing the damped harmonic oscillator by applying a sinusoidally
varying force as a function of time. That is

F = kx b x + F
0
cos t where is now a
frequency that we can control independently. The equation of motion becomes
x + x +
2
0
x =
F
0
m
cos t (141)
We can solve this equation using something of a trick. Pick another variable y(t) so that
y + y +
2
0
y =
F
0
m
sin t (142)
and dene z(t) = x(t) + iy(t). The dierential equation that governs z(t) is therefore
z + z +
2
0
z =
F
0
m
e
it
(143)
where we just need to recall that x(t) = 'z(t) to get the motion of our oscillator.
Applying our ansatz z(t) = /e
it
yields something a little dierent than before. Instead of
nding a characteristic equation to solve for , we instead nd an equation for the (complex)
amplitude / as a function of . That is
/
_

2
+ i +
2
0

=
F
0
m
= / =
F
0
m
1

2
0

2
+ i
(144)
(Note: We can always add in a solution like Eq. 139 since it just yields zero in the dierential
equation. This is where we get the constants we need to apply the initial conditions. This
part of the solution, though, dies out with time, so we usually just ignore it; it is called a
transient.) This is easy to interpret if we write
/ = Ae
i
where A =
F
0
m
1
[(
2
0

2
)
2
+
2

2
]
1/2
and = tan
1
_

2
0
_
(145)
In other words, the motion of the forced damped oscillator
is given by x(t) = Acos(t + ) where A and are strong
functions of the driving frequency . These functions
are shown on the left, for the lightly camped case where

0
. The motion has an amplitude A that becomes very
large when the driving frequency gets close to the natural
frequency
0
. This phenomenon is called resonance.
Note the phase. The motion is in phase = 0 for frequencies
well below
0
, and precisely out of phase ( = 180

) far above

0
. Also, it passes through = ()90

as moves through
the resonance.
The resonance has a width that depends directly on . In fact, if we dene the width
as the distance between the two values for which the curve drops to half its value, then
= . The value Q
0
/ is a dimensionless number which characterizes the width,
and can be interpreted as the amount of energy lost in one oscillation. Large Q means low
damping, and is an important gure of merit for resonating systems.
46
Physics I Honors Thursday 6 November Fall 2008
Hand back exam, discuss grade statistics. (Average on test was 76.73.)
Waves on a Stretched String: The Wave Equation
These notes follow Chap.18 of Physics, by Resnick, Halliday, and Krane, 5th Ed.
What is the equation of motion for a string that is stretched under some tension F?
The motion is up and down, so analyze the forces
in the vertical direction on a small piece of string
at xed x. We assume small amplitude waves so
x can represent the length of the small piece. See
the gure. If F is the tension in the string, then

F
y
= F sin
2
F sin
1
= (m)a
y
(146)
Put m = x for linear mass density . Obviously a
y
= y =
2
y/t
2
since we are working
at xed x. Also, for small deections, 1 so sin tan = slope = y/x. So,
F(sin
2
sin
1
) F
_
y
x

y
x

1
_
= x

2
y
t
2
=
1
x
_
y
x

y
x

1
_
=

F

2
y
t
2
(147)
The left side Eq. 147 is just
2
y/x
2
as x 0. This gives

2
y
x
2
=
1
v
2

2
y
t
2
(148)
where 1/v
2
/F. Eq. 148 is called the wave equation. It
describes a shape of the string, function y of x, which changes
in time, hence y(x, t). The solution is any function like
y(x, t) = f(x

) = f(x vt) i.e. x

x vt (149)
This means that the change in time is rather peculiar. The
shape f(x) moves to the right or left with speed v. This is
the denition of a (non-dispersive) wave.
Practice Exercise
Which of the following solves the wave equation, Eq. 148, for v = 1? Take derivatives, or
show which are of the form Eq. 149. Note that the sum of two waves is also a wave, i.e. the
principle of superposition; This is because Eq. 148 is a linear dierential equation.
(A) y(x, t) = x
2
t
2
; then try y(x, t) = x
2
t
2
(B) y(x, t) = sin x
2
sin t
(C) y(x, t) = log(x
2
t
2
) log(x t)
(D) y(x, t) = e
x
sin t
47
Sine Waves
We almost always describe waves as if the shape f(x vt) is a sine function.
It is handy to describe a sine wave as follows:
y(x, t) = y
m
sin
_
2

(x vt)
_
(150)
We call the wavelength, for obvious reasons. Note that for xed x (say x = 0) we have
y(0, t) = y
m
sin [2vt/] which means that the transverse oscillations are sinusoidal with
period T = /v = 1/f where f (or, maybe, ) is the frequency.
The more common way to write the equation of a sine wave is
y(x, t) = y
m
sin(kx t ) (151)
where k = 2/, = 2f and is called the phase constant.
Standing Waves
Add two sine waves of the same speed and frequency, but moving in opposite directions:
y
1
(x, t) = y
m
sin(kx t) and y
2
(x, t) = y
m
sin(kx + t)
y(x, t) = y
1
(x, t) + y
2
(x, t) = [2y
m
cos(t)] sin(kx) (152)
where the result follows from a trig identity. This wave has zero velocity (i.e. kx = k(x0t))
but an amplitude that varies like cos t. It is called a standing wave. Watch:
Something important happens when we set up standing
waves on a stretched string that is xed at both ends, like
a guitar string. As shown in the gure, we can only have
standing waves that t into the distance L between the
two xed ends. That is
L = n

2

n
=
2L
n
f
n
= n
v
2L
(153)
That is, frequencies have to be integer multiples of a fun-
damental frequency v/2L. This is the basis of the har-
monic sound of nearly all string and wind instruments.
48
Physics I Honors Monday 10 November Fall 2008
The topic we call coupled oscillations has far reaching implications. The formalism ends
up being appropriate for many dierent applications, some of which bear only a passing
resemblance to classical oscillation phenomena. This includes the mathematics of eigenvalues
and eigenvectors, for example.
These notes describe the elementary features of coupled mechanical oscillations. We use one
specic example, and describe the method of solution and the physical implications implied
by that solution. More complicated examples, and their solutions, are easy to come by, for
example on the web.
1
The following gure describes the problem we will solve:
x
1 x
2
m m
k k
k
c
Two equal masses m slide horizontally on a frictionless surface. Each is attached to a xed
point by a spring of spring constant k. They are coupled to each other by a spring with
spring constant k
c
. The positions of the two masses, relative to their equilibrium position,
are given by x
1
and x
2
respectively.
Now realize an important point. We have two masses, each described by their own position
coordinate. That means we will have two equations of motion, one in terms of x
1
and the
other in terms x
2
. Furthermore, since the motion of one of the masses determines the extent
to which the spring k
c
is stretched, and therefore aects the motion of the other mass, these
two equations will be coupled as well. We will have to develop some new mathematics in
order to solve these coupled dierential equations.
We get the equations of motion from F = ma, so lets do that rst for the mass on the
left, i.e., the one whose position is specied by x
1
. It is acted on by two forces, the spring k
on its left and the spring k
c
on its right. The force from the spring on the left is easy. It is
just kx
1
.
The spring on the right is a little trickier. It will be proportional to (x
2
x
1
) since that is
the extent to which the spring is stretched. (In other words, if x
1
= x
2
, then the length of
spring k
c
is not changed from its equilibrium value.) It will also be multiplied by k
c
, but we
need to get the sign right. Note that if x
2
> x
1
, then the string is stretched, and the force on
the mass will be to the right, i.e. positive. On the other hand, if x
2
< x
1
, then the spring is
compressed and the force on the mass will be negative. This makes it clear that we should
write the force on the mass as +k
c
(x
2
x
1
).
1
See http://math.fullerton.edu/mathews/n2003/SpringMassMod.html.
49
The equation of motion for the rst mass is therefore
kx
1
+ k
c
(x
2
x
1
) = m x
1
(154a)
We get some extra reassurance that we got the sign right on the second force, because if
k = k
c
, then the term proportional to x
1
does not cancel out.
The equation of motion for the second mass is now easy to get. Once again, from the spring
k on the right, if x
2
is positive, then the spring pushes back so the force is kx
2
. The force
from the coupling spring is the same magnitude as for the rst mass, but in the opposite
direction, so k
c
(x
2
x
1
). This equation of motion is therefore
kx
2
k
c
(x
2
x
1
) = m x
2
(154b)
So now, we can write these two equations together, with a little bit of rearrangement:
(k + k
c
)x
1
k
c
x
2
= m x
1
(155a)
k
c
x
1
+ (k + k
c
)x
2
= m x
2
(155b)
To use some jargon, these are coupled, linear, dierential equations. To solve these equa-
tions is to nd functions x
1
(t) and x
2
(t) which simultaneous satisfy both of them. We can
do that pretty easily using the exponential form of sines and cosines, which we discussed
earlier when we did oscillations with one mass and one spring.
Following our discussions about single mass oscillating systems, we assume the same ansatz
and write
x
1
(t) = a
1
e
it
x
2
(t) = a
2
e
it
(156)
where the task is now to see if we can nd expressions for a
1
, a
2
, and which satisfy the dif-
ferential equations, including the initial conditions. Youll recall that taking the derivative
twice of functions like this, brings down a factor of i twice, so a factor of
2
. There-
fore, plugging these functions into our coupled dierential equations gives us the algebraic
equations
2
(k + k
c
)a
1
k
c
a
2
= m
2
a
1
(158a)
k
c
a
1
+ (k + k
c
)a
2
= m
2
a
2
(158b)
This is good. Algebraic equations are a lot easier to solve than dierential equations. To
make things even simpler, lets divide through by m, do a little more rearranging, and dene
two new quantities
2
0
k/m and
2
c
k
c
/m. Our equations now become
(
2
0
+
2
c

2
)a
1

2
c
a
2
= 0 (159a)

2
c
a
1
+ (
2
0
+
2
c

2
)a
2
= 0 (159b)
2
A mathematician would call collectively call these two equations an eigenvalue equation. The reason
becomes clearer when you write this using matrices. In this case, Eqs. 158 become
_
k + k
c
k
c
k
c
k + k
c
_ _
a
1
a
2
_
= m
2
_
a
1
a
2
_
(157)
50
So, what have we accomplished? We think that some kind of conditions on a
1
, a
2
, and
will solve these equations. Actually, we can see a solution right away. In mathematicians
language, these are two coupled homogeneous (i.e. = 0) equations in the two unknowns a
1
and a
2
. The solution has to be a
1
= a
2
= 0. Yes, that is a solution, but it is a very boring
one. All it means as that the two masses dont ever move.
The way out of this dilemma is to turn these two equations into one equation. In other
words, if the left hand side of one equation was a multiple of the other, then both equations
would be saying the same thing, and we could solve for a relationship between a
1
and a
2
,
but not a
1
and a
2
separately. Mathematically, this condition is just that the ratio of the
coecients of a
2
and a
1
for one of the equations is the same as for the other, namely
3

2
c

2
0
+
2
c

2
=

2
0
+
2
c

2

2
c
(160)

4
c
= (
2
0
+
2
c

2
)
2

2
c
=
2
0
+
2
c

2

2
=
2
0
+
2
c

2
c
(161)
In other words, these two equations are really one equation if

2
=
2
0

2
A
(162)
or, instead, if

2
=
2
0
+ 2
2
c

2
B
(163)
Borrowing some language from the mathematicians, the physicist refers to
2
A
and
2
B
as
eigenvalues. We will also refer to A and B as eigenmodes.
What is the physical interpretation of the eigenmodes? To answer this, we go back to
Eq. 159a or, equivalently, Eq. 159b. (Remember, they are the same equation now.) If we
substitute
2
=
2
A
into Eq. 159a we nd that
a
A
1
= a
A
2
(164)
where the superscript just marks the eigenmode that were talking about. In other words,
the two masses move together in lock step, with the same motion. This happens when
the frequency is =
0
=
_
k/m, and that is just what you expect. It is as if the two
masses are actually one, with mass 2m. The eective spring constant is 2k, as you learned
in your laboratory exercise, where you had one cart (with variable mass) and a spring on
each side. Therefore, we expect this double-mass to oscillate with
2
= (2k)/(2m) = k/m.
Good. This all makes sense.
Now consider eigemode B. Substituting
2
=
2
0
+ 2
2
c
into Eq. 159a, we nd that
a
B
1
= a
B
2
(165)
In other words, the two masses oscillate against each other, and with a somewhat higher
frequency. In this case, it is interesting to see the dierence between
c

0
(in other
words, a very weak coupling spring) and
c

0
(strong coupling spring). For the weak
3
A mathematician would say that we are setting the determinant equal to zero.
51
coupling spring, the frequency is just the same as
0
, and that makes sense. If the two
masses arent coupled to each other very strongly, then they act as two independent masses
m each with their own spring constant k. On the other hand, for a strong coupling between
the masses, the frequency is =
c

2, which gets arbitrarily large. Sometimes, this high


frequency mode can be hard to observe.
A simple experimental test of this result is to set up two identical masses with three identical
springs. In other words,
c
=
0
. In that case
2
B
= 3
2
A
and therefore the frequency for
mode B should be

3 times larger than the frequency for mode A. We should be able to


make this test as a demonstration in class.
We have been talking about general properties of the two eigenmodes. Of course, this
doesnt tell you how the system behaves given certain starting values, that is, specic initial
conditions.
For this, we need to understand that the two eigenmodes can be combined. In fact, the gen-
eral motion of each of the masses really needs to be written as a sum of the two eigenmodes.
Each of these comes with a positive and negative frequency, just as it did when we applied
all this to the motion of a single mass and spring. (Recall that the sum and dierence of
a positive and negative frequency exponential, are equivalent to a cosine and sine of that
frequency.) Incorporating what weve learned about the relative motions of modes A and B,
we can write the motions as follows:
x
1
(t) = ae
i
A
t
+ be
i
A
t
+ ce
i
B
t
+ de
i
B
t
(166a)
x
2
(t) = ae
i
A
t
+ be
i
A
t
ce
i
B
t
de
i
B
t
(166b)
where a, b, c, and d are constants which are determined from the initial conditions. Note
that these four constants appear as they do in Eqs. 166 because each frequency term must
match up between the equation for x
1
(t) and x
2
(t), with the correct relative sign, in order
to solve the coupled dierential equations.
An obvious set of initial conditions is starting mass #1 from rest at x = x
0
, and with mass
x
2
from rest at its equilibrium position. This leads to the solution
x
1
(t) = (x
0
/2) [cos(
A
t) + cos(
B
t)] (167)
x
2
(t) = (x
0
/2) [cos(
A
t) cos(
B
t)] (168)
The last laboratory experiment for this course will investigate this motion. Some number
of setups will be using new electronics that allows each mass to be followed separately and
simultaneously.
On to the plots. Two cases are shown, one with
c
= omega
0
, and the other with
c
=
0
/4.
For each case, we plot x
1
(t) with x
2
(t), and also x
1
(t) + x
2
(t) with x
1
(t) x
2
(t). The rst
case shows clearly that the motion is somewhat complicated for either mass, but clearly
splits into two eigenmodes with rather dierent frequencies for the sum and dierence. The
second case shows something much closer to a beat pattern, with the energy shifting from
one mass to the other and then back again.
52
0 5 10 15 20 25 30
20
15
10
5
0
5
10
15
20

0
=1
c
=1


x
1
x
2
0 5 10 15 20 25 30
20
15
10
5
0
5
10
15
20

0
=1
c
=1


x
1
+x
2
x
1
x
2
53
0 20 40 60 80 100 120
20
15
10
5
0
5
10
15
20

0
=1
c
=0.25


x
1
x
2
0 20 40 60 80 100 120
20
15
10
5
0
5
10
15
20

0
=1
c
=0.25


x
1
+x
2
x
1
x
2
54
Physics I Honors Thursday 13 November Fall 2008
Today we start a new general topic, namely Thermodynamics. I will introduce it to you
by discussing the related subject of Statistical Mechanics. This is a way of simplifying
complicated systems by averaging over their properties, an approach that should work if
they are complicated enough.
The Ideal Gas
Lets imagine a box lled with a very large number of particles. The particles are very small
and dont interact with each other. There are so many particles, and they are moving so
quickly, that all you feel is a steady pressure on the walls.
See the gure. Put N gas molecules each of mass m in a
cubic box of side length L. Molecule has x velocity v
x
. It
bounces o the wall and incurs momentum change 2mv
x
.
The time between bounces is 2L/v
x
so F
x
(2L/v
x
) = 2mv
x
.
The pressure on the wall at x = L is the sum of the forces
of all the molecules, divided by the area of the wall. That is
p =
1
L
2

F
x
=
1
L
2

mv
2
x
L
=
Nm
V
_
1
N

v
2
x
_
(169)
where we recognize that the volume of the cube is V = L
3
. Now the quantity in parentheses
is the average value of v
2
x
, written v
2
x
). There is nothing special about the x direction, and
the number N is very large, so we expect that v
2
x
) = v
2
y
) = v
2
z
) =
1
3
v
2
) where we call v
2
)
the mean square velocity. Since all molecules have the same mass, K) =
1
2
mv
2
) is the
average kinetic energy of the molecules. So, we can now write
pV = N
2
3
K) (170)
Now we invent something called temperature which measures the amount of energy per
particle in the gas. For historical reasons, the temperature T is measured in dierent units
called Kelvin (K). The conversion constant from temperature to energy is called k, known
as Boltzmanns constant, and has the value k = 1.38 10
23
J/K. We assign half a unit of
kT to each of the three directions in which the particle can move. In other words
K) =
3
2
kT = pV = NkT (171)
This is known as the ideal gas law. You may have already seen this in a chemistry course,
but chemists count particles in a dierent way. They prefer to count moles where a mole is
Avogadros number N
A
= 6.02 10
23
of particles. That is, N = nN
A
where n is the number
of moles. Chemists also write R N
A
k, that is pV = nRT.
Practice Exercise
A one liter cubical box is 10 cm on a side. Assume it is lled with molecules of oxygen (atomic
mass 32 amu) at room temperature (T = 300K) and under a pressure of one atmosphere
(10
5
Pa, where a pascal is one newton per square meter). Calculate the number N of oxygen
molecules in the box, the average speed, i.e v
2
)
1/2
, and the average number of collisions
with the box walls in one second.
55
Practical Aspects of Temperature
The kind of temperature weve just dened is called absolute temperature. Sometimes it is
called the Kelvin temperature scale. As far as physics is concerned, it is the only kind of
temperature that matters.
There are other common temperature scales used in everyday life, although neither is used
much by physicists. The two ordinary scales are Fahrenheit (T
F
) and Celsius (T
C
), i.e.
T
C
= T 273.15 and T
F
=
9
5
T
C
+ 32 (172)
This is no historical introduction to temperature. Initially people wanted to measure how
hot or cold something was. They found some material with a property that was sensitive
this quantity. Then they dened temperature scales by pinning two values to the behavior
of the particular material property, that is, boiling and freezing of pure water in the case of
Celsius, and body temperature and freezing point of brine in the case of Fahrenheit.
Then we came across the more physical meaning of temperature, based on internal energy,
and the absolute temperature scale was born. It was chosen to have the same size of degree
as Celsius, but put the zero of the scale at the place where there would be no internal kinetic
energy. Sometimes we call that point absolute zero.
Thermal Expansion: A Material Property that Used to Dene Temperature
Most solids expand when they get hot. The reason is subtle.
Higher temp means faster
molecular motion. Solids
are like atoms connected by
anharmonic springs, and
average position shifts to
larger separations as the en-
ergy increases. So, solids ex-
pand.
The eect is rather linear and not large for typical temperatures. For solids, we write
L = L T (173)
where is called the coecient of linear expansion and expresses the fractional change in
length (L/L) per degree of temperature change, and varies from solid to solid. Since the
fractional change is small, we expect A = 2A T for expansion of surface area, and
V = 3V T for volume. (Remember your Taylor expansions.) Note that uids (like
water!) are more complicated than solids.
Practice Exercises
(1) Determine the temperature at which the Fahrenheit and Celsius scales coincide, that is,
give the same value for the temperature. Have you ever been outside on a day or night when
this was the air temperature?
(2) Explain why our denition K)
3
2
kT for the temperature of an ideal gas, would not
make sense if we wrote it in terms of T
F
or T
C
, instead of absolute temperature T.
56
Physics I Honors Monday 17 November Fall 2008
Energy, Work, and Heat
We are going to start talking about systems like the one
shown on the right. It is a cylinder of gas with a piston
that moves without friction. Weight on the top of the
piston keeps the gas in an equilibrium state. That is,
all changes are slow ones. The gas has some internal en-
ergy that we will call U. (For an ideal gas, U = 3NkT/2,
the average kinetic energy of a molecule times the num-
ber of molecules.) Work can be done on the piston by
the gas, and work can be done on the gas by the piston.
This gure plots pressure versus volume for the gas in the
cylinder. (It is called a pV diagram.) Both quantities
can be changed in dierent ways, like adding or subtract-
ing gas molecules or changing the mass M sitting on top
of the piston. (There is also something called a thermal
reservoir but well get to that in a minute.) We write
that work W is done on the gas by the piston. We have
dW = Fdh = pAdh = pdV (174)
where A is the area of the piston and F = pA is the force acting on the gas, in the direction
opposite to that which increases h.
Now consider taking the gas from state B (p = p
f
and V = V
i
) to state D (p = p
f
and
V = V
f
) along path #1. The pressure p = p
f
, a constant, along this path. Therefore the
work done on the gas is
W
path #1
=
_
path #1
pdV = p
f
(V
f
V
i
) = p
f
V (175)
In contrast, along path #2, the volume does not change, so the work done along this path,
connecting state C to state D, is zero. From there, it is simple to see that if we go from B
to D instead of along path #1, but rather down to A, then across to C and then back up to
D, the work done on the gas is
W
path BACD
= p
i
(V
f
V
i
) = p
i
V ,= W
path #1
(176)
So, the work done in going from one state to another will depend on the path taken on
the pV diagram. We refer to U, p, V , and also the temperature T, as state variables
because they only depend on the properties of the gas at any one time. Work W is not a
state variable; it depends on the path taken to get between dierent states.
Next lets think about what happened to the gas in going from B to D along path #1. For
simplicity, assume it is an ideal gas. Also assume that the gas supply valve is turned o, so
57
no gas molecules are added or taken away. Then since pV = NkT, the temperature must
have increased, since p = p
f
has not changed but the volume has increased from V
i
to V
f
.
Since the temperature has gone up, so has the internal energy U. The work done on the
gas, Eq. 175, is negative, so that would have decreased the internal energy. Where did the
extra internal energy come from? It came from the thermal reservoir, by transferring energy
in the form of heat. Heat (Q) is a new concept. It is a dierent way of transferring energy,
other than work (W), in a thermodynamic system.
Energy is conserved. Always. In thermodynamic systems as well as mechanical ones. Our
expression of energy conservation for the gas in the cylinder can be written as
Q + W = U (177)
This is called the First Law of Thermodynamics. Note the signs. If heat is added to a
system, then Q is positive. If the gas expands against the piston, then W is negative.
Specic Heat Capacity
We pause for a moment and talk about heat in a general sense. This will be useful later.
Add heat and the internal energy goes up, so the temperature increases. For solids and
liquids, the volume doesnt change, so work is not an issue. For all materials, the amount
by which the temperature increases depends on the mass and the specic properties of the
material. This is all collected into the denition of the specic heat capacity c, namely
c =
1
m
Q
T
(178)
where m is the mass of the object, Q is the heat added or lost, and T is the change in tem-
perature. The specic heat capacity is tabulated in many places for lots of substances.
We write C = mc for heat capacity. Clearly, C depends on how much matter is present.
Specic heat capacity depends on temperature, but it is not a strong function of temperature
for most substances for common temperatures. Under these conditions
Q = m
_
T
f
T
i
cdT = mc(T
f
T
i
) (179)
For an ideal gas under constant volume, Q = U = (3/2)NkT. Therefore C = (3/2)Nk =
(3/2)(N/N
A
)N
A
k = (3/2)nR. We write C
V
= 3R/2 for the molar heat capacity at
constant volume. Real, monatomic gases show molar specic heat capacities very close to
this value. Diatomic or polyatomic gases show higher values, reecting the additional degrees
of freedom in those molecules.
For an ideal gas that expands under constant pressure, W = pV = NkT. Therefore
Q = U W = (5/2)NkT and the molar heat capacity at constant pressure is C
p
=
5R/2. Again, real monatomic gases agree with this very well. In addition, one expects that
C
p
C
V
= R = 8.31 J/mol-K. We commonly write

C
p
C
V
=
5
3
(180)
for a monatomic gas. A table is included which demonstrates these concepts.
58
Cyclic Processes: An Introduction
The world makes a lot of use of so-called cyclic processes in thermodynamics. These are
processes that operate in a cycle, ending up at the same place from which they start. Well
do more with these next class, but lets start with a simple one and see what we can learn
from it.
The gure shows a cyclic process for an ideal gas. It goes
rst fromAto B by reducing the pressure at constant vol-
ume, taking out heat Q
1
and lowering the temperature;
then to C by increasing the volume at constant pressure,
in which case heat Q
2
is added to raise the temperature
and work W
2
< 0 is done on the gas; and then back to
A by reducing the volume and increasing the pressure in
such a way that the temperature remains constant.
Since the process is cyclic, the change in internal energy around the cycle must be zero.
Therefore, by the rst law of thermodynamics, Q + W = 0 where Q and W represent the
net work around the cycle. As drawn in the gure, W
2
< 0. Recall Eq. 175. On the other
hand, W
3
> 0 since the volume decreases. Furthermore, W
3
> W
2
since the area under the
curve dened by path #3 is larger than that for path #2. The work done along path #1 is
zero.
Therefore, the net work is W = W
2
+W
3
> 0, and so the net heat Q < 0. That is, net work
is done on the gas and heat leaves it. We call such a thing a refrigerator. If instead the cycle
operated clockwise, then heat is added and the gas does work on the piston. We call this
an engine. These conclusions will always be the case for cyclic processes, regardless of the
curves that bound them.
The constant volume path #3 is called an isotherm. It follows the curve p = NkT
3
/V , where
T
3
is the temperature at both C and A. Therefore, the work done along this path is
W
3
=
_
V
A
V
C
pdV = NkT
3
_
V
C
V
A
dV
V
= NkT
3
log
_
V
C
V
A
_
> 0 (181)
as advertised. Recall that we know that Q
3
= W
3
since the path is an isotherm.
Adiabatic Processes in an Ideal Gas Left to Next Class
A process carried out in thermal isolation is called adiabatic. It is characterized by Q = 0,
so W = U = (3/2)NkT = (3/2)nRT = nC
V
T. If the temperature change is
innitesimal, then we can write dW = pdV = nC
V
dT. The ideal gas law gives d(pV ) =
pdV + V dp = nRdT or V dp = nC
V
dT + nRdT = nC
p
dT. Dividing these gives us
V dp
pdV
=
nC
p
dT
nC
V
dT
=
C
p
C
V
= =
dp
p
=
dV
V
(182)
Integrating this equation is simple. For limits called A and B, it just says that
log
_
p
B
p
A
_
= log
_
V
B
V
A
_
= log
_
V
A
V
B
_

or pV

= constant (183)
This is the equation that describes an adiabatic path on a pV diagram.
59
Physics I Honors Fall 2008
Useful Tables for the Study of Thermodynamics
Taken from Physics, by Halliday, Resnick, and Krane, Fifth Edition
Coecients of Linear Expansion
Substance (10
6
/K)
Ice 51
Lead 29
Aluminum 23
Brass 19
Copper 17
Steel 11
Ordinary glass 9
Pyrex glass 3.2
Invar alloy 0.7
Fused quartz 0.5
Heat Capacities of Some Substances
Substance Specic Heat Capacity (J/kg-K)
Lead 129
Tungsten 135
Silver 236
Copper 387
Carbon 502
Aluminum 900
Brass 380
Granite 790
Glass 840
Mercury 139
Water 4190
Molar Heat Capacities of Ideal and Real Gases
C
p
C
V
C
p
C
V
Gas (J/mol-K) (J/mol-K) (J/mol-K)
Monatomic
Ideal 20.8 12.5 8.3 1.67
Helium 20.8 12.5 8.3 1.67
Argon 20.8 12.5 8.3 1.67
Diatomic
Ideal 29.1 20.8 8.3 1.40
H
2
28.8 20.4 8.4 1.41
N
2
29.1 20.8 8.3 1.40
O
2
29.4 21.1 8.3 1.40
Polyatomic
Ideal 33.3 24.9 8.3 1.33
CO
2
37.0 28.5 8.5 1.30
NH
3
36.8 27.8 9.0 1.31
60
Physics I Honors Fall 2008
Homework Assignment Due Thursday 20 November
(1) The following diagram shows the problem in coupled oscillators that we solved in class:
x
1 x
2
m m
k k
k
c
Using the parameters
2
0
k/m and
2
c
k
c
/m, nd the motions x
1
(t) and x
2
(t) for the
following sets of initial conditions:
a) x
1
(0) = x
2
(0) = A and x
1
(0) = x
2
(0) = 0
b) x
1
(0) = x
2
(0) = 0 and x
1
(0) = x
2
(0) = V
c) x
1
(0) = x
2
(0) = x
2
(0) = 0 and x
1
(0) = V
(2) As a result of a temperature rise of 32 C

, a bar
with a crack at its center buckles upward, as shown
in the gure on the right. The distance L
0
= 3.77 m
between the bar supports is xed, and the coecient
of linear expansion of the bar is 25 10
6
/K. Find x,
the distance to which the center rises.
(3) In an experiment, 1.35 moles of oxygen (O
2
, a diatomic molecule) are heated at con-
stant pressure, starting at 11

C. How much heat must be added to the gas to double its


volume?
(4) Gas within a chamber undergoes the process shown in the pV diagram below:
Find the net heat (in Joules) added to the system during one complete cycle.
61
Physics I Honors Thursday 20 November Fall 2008
Review: Statistical Mechanics and Thermodynamics
Concentrate for now on gases. For a monatomic ideal gas, the equation of state is
pV =
2
3
K)NkT = nRT (184)
which relates the state variables p, V , and T to the number of molecules N through a
fundamental constant k. The internal energy U = N(3/2)kT = NK) where K) is the
average molecular kinetic energy. The 3 counts the number of degrees of freedom,i.e.
three because all a monatomic gas molecule can do is move in three directions. If, for
example, the molecule is diatomic and can rotate, there are ve degrees of freedom.
If the gas container has movable walls, then work W can be done on the gas. If it is in
contact with some kind of thermal bath, then heat Q can be transferred into the gas. These
change the state of the gas by an amount the depends on the path taken between the
states. The rst law of thermodynamics says that
U = W + Q (185)
where by convention, W is the work done on the gas. In eect, this equation denes Q.
Entropy: A New State Variable
Now we dene a new state variable S called entropy. Actually, rather than dening entropy
itself, well dene the change in entropy S between two states. We have
S
_
dQ
T

reversible
(186)
The notation implies that the path for the integral is reversible, that is, connected by a
series of equilibrium states. For example, S = Q/T for a reversible isothermal path.
Heres a trickier example, a class problem of free expan-
sion. The walls are perfectly insulated, so no heat comes
in to the gas. Opening the stopcock lets gas molecules move
into the evacuated chamber, until there is equal density
everywhere. What is the change in entropy?
You are tempted to say S = 0 since Q = 0. However, the
process shown is not reversible. We need to compute S =
_
dQ/T along some reversible path. Since no work is done
and no heat is transferred, U = 0. So, choose an isothermal
expansion from V
i
to V
f
. Therefore dQ = dW = pdV and
S =
_
V
f
V
i
pdV
T
= Nk
_
V
f
V
i
dV
V
= Nk ln
_
V
f
V
i
_
> 0 (187)
62
The Second Law of Thermodynamics
Note that free expansion of a gas (Eq. 187) always yields an increase in entropy. This is a
system completely isolated from its surroundings, and the expansion process is irreversible.
In a reversible process, like isothermal expansion or compression, heat is transferred through
a common thermal reservoir, so we dont expect any change in net entropy if we include both
the gas and the thermal bath with which it is in contact.
So, now we state the Second Law of Thermodynamics, namely When changes occur within
a closed system, its entropy either increases (for irreversible processes) or remains constant
(for reversible processes). It never decreases. In other words S 0.
Adiabatic Processes in an Ideal Gas
A process carried out in thermal isolation is called adiabatic. It is characterized by Q = 0,
so W = U = (3/2)NkT = (3/2)nRT = nC
V
T. If the temperature change is
innitesimal, then we can write dW = pdV = nC
V
dT. The ideal gas law gives d(pV ) =
pdV + V dp = nRdT or V dp = nC
V
dT + nRdT = nC
p
dT. Dividing these gives us
V dp
pdV
=
nC
p
dT
nC
V
dT
=
C
p
C
V
= =
dp
p
=
dV
V
(188)
Integrating this equation is simple. For limits called A and B, it just says that
log
_
p
B
p
A
_
= log
_
V
B
V
A
_
= log
_
V
A
V
B
_

or pV

= constant (189)
This is the equation that describes an adiabatic path on a pV diagram.
The Carnot Engine
The gure on the right shows a cyclic process
which is bounded by two isotherms (at two
temperatures T
L
< T
H
) and two adiabats which
connect those two temperatures. This is called a
Carnot cycle. The cycle proceeds clockwise, so
negative work is done on the gas from A to B and
heat Q
H
enters the gas. During the isothermal
contraction C to D, the work is positive and so
heat Q
L
leaves the gas. (Remember that U = 0
along an isotherm, and Q = 0 along an adiabat.)
Since U = 0 around the cycle, we must have
W
DA
+ W
H
+ Q
H
+ W
BC
+ W
L
+ Q
L
= 0 (190)
Now W
H
< 0 and W
L
> 0. Futhermore, since paths H and L are isotherms, W
H
+Q
H
= 0 =
W
L
+Q
L
. Therefore W
DA
+W
BC
= 0, and also Q
H
= W
H
> 0 (heat enters along H) and
Q
L
= W
L
< 0 (heat exits along L). Since (by areas) [W
L
[ < [W
H
[, we have [Q
L
[ < [Q
H
[,
so net heat enters the gas. Furthermore, the work done around the cycle is
W = W
DA
+ W
H
+ W
BC
+ W
L
= W
H
+ W
L
< 0 (191)
and the gas does work on the piston. Heat in, work out. This is an engine.
63
How ecient is this engine? In other words, how much work do you get out, for heat that
you need to put in? Let us dene an engine eciency just that way, namely

[W[
[Q
H
[
=
[W
H
+ W
L
[
[Q
H
[
=
[Q
H
[ [Q
L
[
[Q
H
[
= 1
[Q
L
[
[Q
H
[
(192)
We can go further by realizing that entropy change is zero around the cycle. That is
S =
Q
H
T
H
+
Q
L
T
L
=
[Q
H
[
T
H

[Q
L
[
T
L
= 0 (193)
Therefore [Q
L
[/[Q
H
[ = T
L
/T
H
and the eciency of a Carnot engine is
= 1
T
L
T
H
(Carnot engine eciency) (194)
Of course, operated counter-clockwise around the cycle, a Carnot engine becomes a Carnot
refrigerator. Instead of eciency we speak of coecient of performance. This is the ratio
of what you want, i.e. the heat Q
L
taken out of the low temperature reservoir, and what
you pay for, namely the work W you have to do on the gas. We write
K =
[Q
L
[
[W[
=
[Q
L
[
[Q
H
[ [Q
L
[
=
T
L
T
H
T
L
(Carnot refrig coe of performance) (195)
A very large coecient of performance is needed for a refrigerator that operates between two
temperatures that are close to each other.
There is no such thing as a perfect engine, i.e. one with = 1. This requires T
L
= 0
for a cold reservoir, that is one with no internal energy. There is also no such thing as a
perfect refrigerator, i.e. one with an innite coecient of performance. This would mean
that [Q
H
[ = [Q
L
[ Q and S = Q/T
H
Q/T
L
< 0, which violates the second law.
These statements can be put together to show that No real engine can have an eciency
greater than that of a Carnot engine working between the same two temperatures. That
is, Eq. 194 is the most ecient an engine can ever be. In other words, by Eq. 192, any real
engine must always discharge some heat Q
L
to its surroundings.
Dont Forget: Short homework assignment due in class on Monday.
64
Physics I Honors Fall 2008
Homework Assignment Due Monday 24 November
1) Two identical blocks, each of mass M and made from a material with specic heat capacity
c, are kept thermally isolated from each other and from their surroundings. One block is at
temperature T
1
and the other is at temperature T
2
> T
1
. The blocks are then put in thermal
contact with each other, but still thermally isolated from their surroundings.
a. Use the relationship between heat transfer and specic heat capacity to prove that the
nal equilibrium temperature of the two blocks together is T = (T
1
+ T
2
)/2.
b. Find the change in entropy for the system consisting of the two blocks. (You can
imagine a reversible process by which one block is taken slowly to its nal temperature,
and then the other block is taken to this same temperature, but contact with a slowly
changing thermal bath.)
2) The gure on the right shows a cyclic
process, plotted as a TS diagram, rather
than a pV diagram.
a. Explain why this is a Carnot cycle.
b. Calculate the heat that enters.
c. Calculate the work done on the gas.
65
Ludwig Boltzmann is one of the few people who understood en-
tropy. He came up with a way to understand it in terms of disor-
der, which he wrote as W. His result, in the form of the equation
S = k log W, is etched on his grave stone.
66
Physics I Honors Monday 24 November Fall 2008
Remember that homework and nal lab notebooks are due on Monday 1 Dec.
How to Build a Star: Hydrostatics of really big ball of gas
Start with the balance of pressure against gravitation. Then use ideal gas law to relate to
temperature. (Will assume that stars like the sun have uniform density.) Finish with the
behavior of white dwarf stars, aka degenerate Fermi gas.
More reading (in order of increasing sophistication):
Introductory Astronomy and Astrophysics, 4th Edition, Zeilik and Gregory, esp. Chap.10
Order-of-magnitude theory of stellar structure, George Greenstein, Am. J. Phys. 55,
804 (1987); Erratum Am. J. Phys. 56, 94 (1988)
Stars and statistical physics: A teaching experience, Roger Balian and Jean-Paul
Blaizot Am. J. Phys. 67, 1189 (1999)
Begin. First, set up the notation. Build our star with shells of radius r. Total mass of star
is M and radius is R. Density is = M/(4R
3
/3) assumed to be independent of r. Write

M(r) be the mass enclosed at radius r, i.e.



M(R) = (4r
3
/3) = M(r/R)
3
.
Now, balance pressure p(r) against gravity. Take a small piece of area A, thickness dr and
mass dm = Adr, at radius r. Outward force from dierence in pressure must equal the
gravitational force from mass inside radius r. (Recall shell theorems.) Write it out:
pA (p + dp)A = G

M(r)dm
r
2
= G

M(r)Adr
r
2
or,
dp
dr
= G

M(r)(r)
r
2
(196)
This equation expresses hydrostatic equilibrium for a large, spherical self-gravitating object
like a star (or a planet, or a moon). It is written here in a way that holds for any density
function (r), but we will be taking to be a constant.
Lets use this to determine the pressure p
C
at the center of a star, i.e. r = 0. Make the
replacements for constant density in Eq. 196 to get
dp
dr
= GM
_
r
R
_
3
M
4R
3
/3
1
r
2
=
3GM
2
4R
6
r = p(r) = p
C

3GM
2
8R
6
r
2
At the surface of the star, the pressure is zero by denition, i.e. p(R) = 0. Therefore,
p
c
=
3GM
2
8R
4
(197)
Next let us try to estimate the temperature at the center of a star. This requires us to know
the equation of state which relates the gas pressure and density to the temperature, for
the matter that makes up the star. Understanding the equation of state is a big deal, and
67
an area of active research interest. For now, we will use an equation of state based on an
ideal gas, that is Eq. 184, namely pV = NkT.
If we assume that the star is made of N identical particles with mass m, then = Nm/V and
p = kT/m = (3/4)(M/m)kT/R
3
. In fact, stars are hot, so hot that matter is a plasma
of free nuclei and electrons. The nuclei are from mostly hydrogen, about 25% helium, and
a few percent of heavier elements. For now, though, it is close enough to assume the star is
100% hydrogen atoms. Therefore, the temperature at the center of a star is
T
C
=
1
k
4R
3
3
m
M
p
C
=
1
k
GmM
2R
(198)
Can you see why we would not be able to get this answer from dimensional analysis?
The White Dwarf Star
We found some properties of stars assuming they consisted of an ideal gas, but the as-
sumptions used to get Eq. 184 dont always hold. Instead, we need another equation of
state.
In deriving Eq. 184 we wrote F
x
= p/t = (2p
x
)/(2L/v
x
), and that is still ne. For N
particles, this gives a pressure P = NF
x
/L
2
= Npv/V . (We temporarily switch to P to
avoid confusion with momentum. Also, we are making a rough estimate, so put p
x
and v
x
to p and v.) Our concern will be with electrons, so write an electron density n
e
= N/V in
which case P = n
e
pv.
At very high densities, the ideal gas law assumptions are violated. The electrons are so close
together, they start to violate the Pauli Exclusion Principle. The distance between them is
d 1/n
1/3
e
and the Heisenberg Uncertainty Principle says that pd h so that p hn
1/3
e
.
Putting v = p/m
e
and switching back to p for pressure, the equation of state becomes
p = n
e
_
hn
1/3
e
_ _
hn
1/3
e
/m
e
_
=
h
2
m
e
n
5/3
e

h
2
m
e
n
5/3
=
h
2
m
e
_

m
_
5/3
(199)
where n is the number density of (what would have been) atoms, and we make the (somewhat
incorrect) assumption that the star is still made up of hydrogen atoms with mass m. This
is an odd sort of equation of state. Combined with Eq 197 it leads to a star whose radius
decreases with mass like 1/M
1/3
.
As the mass increases, and the radius decreases, the electrons are more and more conned
until they are moving as fast as they can, i.e. v = c. The equation of state becomes
p = n
e
_
hn
1/3
e
_
c = hc n
4/3
e
hc n
4/3
= hc
_

m
_
4/3
(200)
Combined this with Eq 197 gives an equation where the radius drops out! One solves for a
mass that is the highest possible mass of a white dwarf star. Doing a more careful job leads
to a value of 1.44 solar masses. This is the Chandrasekhar limit after the physicist who
rst derived it. What do you suppose would happen for a star of mass larger than this?
68
Physics I Honors Fall 2008
Homework Assignment Due Monday 1 December
(1) This problem asks you to carry through some numerical calculations that lead to an
understanding of what it is like in the center of the Sun. The following web sites contains
useful data:
http://pdg.lbl.gov/2008/reviews/consrpp.pdf
http://pdg.lbl.gov/2008/reviews/astrorpp.pdf
a. Use Eq. 197 to estimate the central pressure of the Sun. Express it in units of atmo-
spheres, where one atmosphere is the air pressure at the surface of the Earth.
b. Use Eq. 198 to estimate the central temperature. Express it in Kelvin (K).
c. Recall that temperature is related to the average kinetic energy of the gas particles.
What is the average kinetic energy of the hydrogen atoms in the center of the Sun?
Express the answer in electron volts (eV). What does this tell you about the kind of
matter in the center of the Sun? (A typical binding energy of an electron in an atom
is around 10 eV.)
(2) Show that Eq. 199 leads to a star with radius R and mass M where R 1/M
1/3
.
69
Physics I Honors Monday 1 December Fall 2008
Last class! I will miss all of you, but dont be a stranger.
Hand in lab books. Third midterm exam is this Thursday. Final exam is Wednesday, 3 Dec
3-6pm, this room.
Conservation Laws and the Continuity Equation
Our last class will concentrate on how to write down a general conservation principle. Then,
well apply this to uid dynamics.
The basic idea is simple. Imagine some substance that is conserved. In other words, it
cannot appear or disappear unless it is somehow supplied. If we talk about the amount Q
of this substance in some region 1 of space, then the only way to change this amount is to
let some ow into (or out of) the region.
Make this statement mathematical. The rate of change of Q is dQ/dt. Dene some quantity
J = J(r, t) to be the ux density of the substance. That is, it is the amount of the
substance owing (perpendicularly) through some small area per unit time. So, through
some small surface dA = dA n, an amount J dA = JdAcos ows per unit time. Here, n
is a unit vector perpendicular to the surface, and is the angle between n and J.
If we let o denote the surface of 1, we then write our conservation condition as
dQ
dt
=
_
S
J dA (201)
The negative sign is there because of the convention that elements of area dA point outward
on o. Therefore, if the integral is positive, there is a net ux out of 1 so Q decreases.
Now let = (r, t) be the substance density. Then Q is the integral of over 1, and
dQ
dt
=
d
dt
_
R
(r, t)dV =
_
R

t
dV (202)
The derivative becomes a partial derivative inside the integral, because all we care about is
the explicit dependence on time. Thats all that would be left after we carry out the integral
over space.
For the right side of Eq. 201 we will make use of a result we rst mentioned in Eq. 70.
Imagine that 1 is small and rectangular, with one corner at (x, y, z) and with sides of length
dx, dy, and dz. In this case
_
S
J dA = J
x
(x + dx, y, z)dydz J
x
(x, y, z)dydz
+ J
y
(x, y + dy, z)dxdz J
y
(x, y, z)dxdz
+ J
z
(x, y, z + dz)dxdy J
z
(x, y, z)dxdy
=
_
J
x
(x + dx, y, z) J
x
(x, y, z)
dx
+
_
dxdydz
=
_
J
x
x
+
J
y
y
+
J
z
z
_
dxdydz (203)
70
Now we can divide any arbitrary region 1 into little rectangular boxes that butt up against
each other. The ux out of one box goes into the one next to it, so in the end, all that matters
are the sides of the boxes along the surface o. We also have dV = dxdydz, and
J =
J
x
x
+
J
y
y
+
J
z
z
(204)
which is called the divergence of J(r, t). Therefore, for any region 1,
_
S
J dA =
_
R
J dV (205)
which is known at the Divergence Theorem or Gauss Theorem. So, combining Eq. 202 with
Eq. 205, we turn Eq. 201 into
_
R
_

t
+ J
_
dV = 0 (206)
which holds for any region 1. In particular, if we make 1 very tiny, then there is no
appreciable variation over the volume, and the therefore the integrand must be zero. In
other words, our statement of conservation becomes

t
+ J = 0 (207)
This is called the Continuity Equation. It nds application in all areas of physics, including
the conservation of material in uid mechanics; conservation of electric charge in electro-
magnetism; conservation of energy and momentum in general relativity; and conservation of
probability in quantum mechanics.
Lets make this physical with the specic example of uid ow. In this case, the conserved
substance is the uid mass, and density (r, t) is just ordinary mass density. To identify
J(r, t), consider the tiny uid mass owing through a small element of area dA in a time
dt. Let the velocity of this tiny mass be v(r, t), and n be a unit vector perpendicular to the
surface of dA. Then the size of the volume element of mass swept out is v(r, t)dt ndA, and
the mass which passes through dA is v(r, t) ndtdA = J dAdt. In other words, J(r, t) = v
is the amount of mass owing (perpendicularly) through some small area per unit time.
The gure shows ow from P to Q in a closed
region. Specialize to = (r), independent of
time. Then Eq. 201 says that the integrated
ux over the surface is zero. Assume the areas
A
1
and A
2
on the right are small, so that the
velocities v
1
and v
2
do not vary much over
them, and are perpendicular to them. Then

1
v
1
A
1
=
2
v
2
A
2
(208)
If the uid is incompressible (liquids), that
is, the density (r, t) =constant everywhere,
then the continuity equation becomes
v
1
A
1
= v
2
A
2
(209)
I.e., water ows faster in constricted pipes.
71
Physics I Honors Fall 2008
Practice Exercises on the Continuity Equation
(1) A pipe of diameter 34.5 cm carries water moving at 2.62 m/sec. How long will it take to
discharge 1600 m
3
of water? Answer: 1h 49 min.
(2) The gure below shows the ow of an incompressible liquid through a straight pipe with
a constriction midway between the input and output:
Let z measure the distance along the axis of the pipe, with x and y perpendicular to z.
What are J
x
and J
y
at values of z near points P and Q? For a value of z near the midpoint
between P and Q (i.e., in the middle of the constriction), make a sketch of J
x
as a function
of x. Use your sketch, and Eq. 207, to explain why [v
2
[ > [v
1
[.
(3) The gure shows the conuence of two streams
to form a river. One stream has a width of 8.2 m,
depth of 3.4 m, and current speed of 2.3 m/sec. The
other stream is 6.8 m wide, 3.2 m deep, and ows
at 2.6 m/sec. The width of the river is 10.7 m and
the current speed is 2.9 m/sec. What is its depth?
Answer: 3.9 m
(4) The gure shows a familiar eect, namely the necking down
of a stream of water when it falls under gravity. (Water adheres
to itself, so that no pipe is necessary to keep the mass continuous.)
Use the continuity equation and conservation of mechanical energy
to explain this eect. If the cross sectional area A
1
is 1.2 cm
2
and
that of A
2
is 0.35 cm
2
, and their separation h = 45 mm, at what
rate R does water ow from the tap? Answer: R = 34 cm
3
/sec.
72

Вам также может понравиться