Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

1

Central force motion


1. Introduction
The motion of two particles interacting by a force that has direction along the line joining
the particles and strength that depends only on the separation of the two particles is
equivalent to a central force motion. In the center-of-mass reference frame, the system is
completely determined by the relative position of the two particles. Furthermore, because
the Lagrangian is spherically symmetric about the center-of-mass, the angular momentum
about the center-of-mass is conserved. Hence the relative motion lies in a plane that has
its normal in the direction of the angular momentum vector. The system can then be
described by two generalized co-ordinates.
Let the two particles have masses, m1 and m2, and position vectors r1 and r2 in the
center-of-mass frame. The Lagrangian is
1
1
L = m1r1 r1 + m2r2 r2 U ( r2 r1 ) .
(1.1)
2
2
However, because the center-of-mass is at the co-ordinate origin, r1 and r2 are not
independent but are related by
m1r1 + m2r2 = 0.
(1.2)
In terms of the particle separation,
r = r2 r1 ,

(1.3)

we have

m2
r,
m1 + m2

(1.4)

m1
r.
m1 + m2

(1.5)

r1 =
and
r2 =
The Lagrangian is then
2

1 m2
1 m1
L = m1
r r + m2
r r U ( r )
2 m1 + m2
2 m1 + m2

(1.6)

1 m1m2
=
r r U ( r ) .
2 m1 + m2
The quantity

=
is called the reduced mass of the system.

m1m2
,
m1 + m2

(1.7)

2
We have reduced the motion of two particles to an equivalent problem of a single
particle of mass acted on by a force directed along the line from the particle to a fixed
point.
The generalized momentum conjugate to the position vector r has components
L
pi =
,
(1.8)
ri
and hence
p = r.

(1.9)

L = r p,

(1.10)

The angular momentum


is orthogonal to both r and p. Hence the motion is in the plane through the origin with
normal parallel to L.
The motion has two degree of freedom and can be described by using polar coordinates (r, ). In these co-ordinates, the Lagrangian is
1
L = r 2 + r 2 2 U ( r ) .
(1.11)
2
Since this does not depend explicitly on , (i.e. is ignorable) there is a first integral of
motion
L
p =
= r 2 = l ,
(1.12)

where l is a constant. In other words, the angular momentum about an axis through the
origin normal to the plane of motion is conserved.
Since the rate at which the position vector sweeps out area is
dA 1 2 d
l
= r
=
,
(1.13)
dt 2 dt 2
conservation of angular momentum leads to Keplers second law of orbital motion.

Also because the Lagrangian does not explicitly depend on time, the Hamiltonian is
conserved. Furthermore, the two conditions for the Hamiltonian to be equal to the energy
are satisfied. Hence

3
2
1 2 2 l
E = r + r 2 + U ( r )
2
r

(1.14)

1 2 1 l
r +
+ U ( r ).
2
2 r 2

Re-arranging to give an expression for r, we find


r =

2 ( E U )

l2
2 2.
r

(1.15)

In general this is not an easy equation to solve for r(t).


Often we are interested in the trajectory taken by the particle. The Lagrange
equation of motion is
U

r r 2 =
.
(1.16)
r
To find an equation for the trajectory, this must be converted to an equation with as the
independent variable. This is done by using the conservation of angular momentum
equation to eliminate the time derivatives. Using equation (1.12), we can eliminate the
angular velocity from equation (1.16) to get

l2
U
=
.
3
r
r

(1.17)

Also from equation (1.12), we have


r =

dr dr l
=
,
d
d r 2

(1.18)

and
d dr l d d dr l
l d dr l
= 2
(1.19)
2=

.
2
dt d
r dt d d r r d d r 2
From the form of the last term in equation (1.19), we see that it is now useful to make the
substitution
1
u= ,
(1.20)
r
to get

r=

r =

l2

u2
2

d 2u
,
d 2

(1.21)

so that equation (1.17) becomes


d 2u
U
+u = 2
.
(1.22)
2
d
l u
This is the equation for the trajectory. It is a solvable linear differential equation if the
right hand side is a linear function of u. This includes the case of an inverse square law
force.

4
Lets take
U = ku ,

(1.23)

k = Gm1m2 .

(1.24)

where for the gravitational force


Equation (1.22) for this particular case is
d 2u
k
+u = 2 ,
2
d
l

(1.25)

which has solution


k
.
(1.26)
l2
Here A and 0 are constants. We see that the solution is formally periodic, but not
necessarily closed (e.g. consider force-free motion for which k = 0). We can take 0 to be
zero by measuring from a position in which u takes a stationary value.
u = A cos ( 0 )

2. Planetary motion
For planetary motion, k = Gm1m2 , and the orbit is given by
1
Gm1m2
= A cos +
.
r
l2

(2.1)

This can be re-written as

= 1 + cos ,

(2.2)

l2
,
Gm1m2

(2.3)

where

and is a constant.
It can be shown that equation (2.2) describes a conic section with one focus at the
origin. is called the eccentricity of the orbit, and 2 is called the latus rectum, which is
a length. (Latus rectum is from the Latin words for side and straight.)
The energy of the orbit is
2

Gm1m2
1 dr l 1 l 2
E =
+

2
2
2 d r 2 r
r
2

1 du l 1 l 2 2
=
u Gm1m2u
+
2 d 2
1 l2
=
( 2 1) .
2 2

(2.4)

5
Hence the energy is negative if < 1, and is minimum for = 0. Negative energy means
that the orbit is bound, i.e. the particle never reaches infinity. In this case the orbits are
ellipses. The minimum energy occurs for a circular orbit. If > 1, the orbit is unbound
and hyperbolic. If = 1, the orbit is a parabola.
To show that equation (2.2) is that of an ellipse for < 1, consider the following
construction.
P

latus rectum
r

O2

O1

rmin

The minimum value of r occurs when = 0, and the maximum value occurs when = .
Draw a line joining these two points (this will be the major axis of the ellipse). Mark off a
distance rmin from each end of this line (these are the positions of the two foci of the
ellipse, O1 and O2). Form the triangle O1PO2. If the sum of the lengths O1P and PO2 is
independent of then the orbit is an ellipse.
The distance between O1 and O2 is
rmax rmin =

= 2

1 1+
From the cosine rule the distance between O2 and P is

( rmax rmin )

.
1 2

+ r 2 + 2 ( rmax rmin ) r cos


2

= 2
+
+ 2 2
cos
2
2
1 1 + cos
1 1 + cos

(1 + 2 cos + ) .
=
(1 ) (1 + cos )
2

Hence the sum of the lengths O1P and PO2 is

r+

( rmax rmin )

+ r 2 + 2 ( rmax rmin ) r cos

(1 + 2 cos + )

=
+
(1 + cos ) (1 2 ) (1 + cos )
2

(1 + 2 cos + 2 )

1 +

=
(1 + cos )
(1 2 )

2
.
(1 2 )

This is independent of and so the orbit is an ellipse. This proves Keplers first Law.
Once we know that the orbit is an ellipse, we can relate the parameters and to
the lengths of the semi-major and semi-minor axes. Clearly

2
2a = rmin + rmax =
+
=
.
(2.5)
1 1+ 1 2
Also by considering the right triangle AO1B in the figure below

B
a

b
O2

O1 rmin

we see that
2

a 2 = b 2 + ( a rmin ) ,

(2.6)

and so
2

r +r r r
b 2 = max min max min
2
2

= rmax rmin

2
=
1 2
= (1 2 ) a 2 ,

(2.7)

which gives the usual definition of eccentricity.


The energy of the orbit can be written in terms of the lengths of the axes by using
equations (2.3), (2.4) and (2.5). We find that

7
2
1 l2
1 l 2 (1 )
1
1
1 Gm1m2
l2
2
E=
=
=
.
(2.8)
1) =
(
2
2
2
2
2 l
2 a
a

Gm1m2
Note that the energy depends on the length of the semi-major axis but not the semi-minor
axis. The angular momentum is related to the axis lengths by
Gm1m2
l 2 = Gm1m2 = a (1 2 ) Gm1m2 = b 2
.
(2.9)
a
Using this last expression we can write the semi-minor axis in terms of the constants of
the motion

l2

b =

2 E

(2.10)

Since the area of the ellipse is ab, Keplers second law [equation (1.13)] enables us to
find the orbital period

Gm1m2

l2

(2.11)
= Gm1m2
.
3
l 2E
2 E
2E
2
By squaring and using equation (2.8), we find the more familiar form of Keplers third
law
P=

ab =

P2 =

4 2
a3.
G ( m1 + m2 )

(2.12)

Again, we note that this expression involves only the semi-major axis.
The point on its orbit at which a planet has its closest approach to the Sun is
called the perihelion. The point at which the planet is furthest from the Sun is the
aphelion. Similarly, for satellites orbiting the Earth, the terms perigee and apogee are
used. In terms of the semi-major axis and the eccentricity,
rmin =

1+

= a (1 ) ,

(2.13)

= a (1 + ) .

(2.14)

and
rmax =

Hence
rmax + rmin = 2a,

(2.15)

rmax rmin = 2a ,

(2.16)

and

rmax rmin
.
rmax + rmin

(2.17)

8
Finally, note that the distance of a focus from the center of the ellipse is
 r r
AO1 = max min = a .
2

(2.18)

3. The effective potential

The total energy

1 2 1 l2
r +
+U (r),
(3.1)
2
2 r2
is the same as that of a particle of mass experiencing linear motion with potential
energy
E=

1 l2
+ U ( r ).
(3.2)
2 r 2
Hence Ueff(r) is an effective potential energy. The force corresponding to the effective
potential energy is
dU eff
l2
dU
Feff ( r ) =
= 3
.
(3.3)
dr
r
dr
U eff ( r ) =

The term l 2 ( r 3 ) is commonly called the centrifugal force. This force and the
corresponding centrifugal potential energy l 2 ( 2 r 2 ) arise from conservation of the
angular momentum of the motion.
Part of the usefulness of the effective potential concept is that it allows graphical
representation of the locations of equilibrium points and turning points. Since the kinetic
energy cannot be negative, we have
1
E U eff ( r ) = r 2 0,
(3.4)
2
with equality only when U eff ( r ) = E.
For bound orbits, the two turning points (where r = 0 ) are called apsides. Note that the
motion is not necessarily periodic and hence the apsides do not have to occur at the same
angular position.
Also the effective potential energy is useful for studying the effects of
perturbations to the system.
EXAMPLE 1
Suppose the interaction potential is U ( r ) = kr 2 2. Graph the effective potential energy.
Identify the equilibrium position, and representative turning points. Find the frequency of

9
oscillation for a small perturbation from equilibrium (that does not change the angular
momentum).
The effective potential energy is

1 l2
1
U eff ( r ) =
+ kr 2 .
2
2 r
2
Clearly this involves a number of parameters, whose values are unspecified. To simplify
the analysis, it is useful to introduce a length scale determined from where the two
contributions to the potential are equal. This occurs at
14

l2
r = R
.
k
Making the substitution r = sR,
1 1

U eff ( s ) = U 0 2 + s 2 ,
2 s

where
12

kl 2
U0 =
.

A plot of U eff U 0 against s looks like


10
9
8
7

Ueff/U0

6
5
4
3
2
1
0
0

Clear the effective potential energy has a minimum at s = 1, which corresponds to a point
of stable equilibrium. Also we see that for E > U0, the particle is trapped between two
turning points. These turning points are found by solving
1 1

E = U0 2 + s2
2 s

for s. This equation can be re-arranged to

10

s4

2E 2
s + 1 = 0,
U0

which has solutions


2

E
E
s =
1.
U0
U0
2

The equation of motion is

l2
+ kr = 0,
r3

which in terms of s is

1
+ s = 0.
k
s3
The equilibrium is at s = 1. Expanding about the equilibrium by setting s = 1 + , and
linearizing the resulting equation we get
k
+ 4 = 0.

Hence the frequency of oscillation is

=2

4. Stability of circular orbits

For a general potential, the equation of radial motion is

l2
dU

.
(4.1)
3
r
dr
For a given angular momentum, circular orbits exist only if there are real solutions of


r=

dU
l2
= 3.
dr r

(4.2)

Suppose there is a solution at r = . We can test for stability by considering a small


perturbation such that r = + r. The equation of motion is


r=

l2

( + r)

dU
d 2U

( ) 2 ( ) r +
dr
dr

(4.3)
3l 2 d 2U

= 4 + 2 ( ) r + .
dr

Dropping terms of second order and higher, and assuming a harmonic time behavior, the
frequency of small oscillations, , satisfies

11

3l 2

d 2U
= 4 + 2 ( ) .
dr

Hence the circular orbit is stable if


2

3l 2

d 2U
( ) > 0.
dr 2

(4.4)

(4.5)

Using equation (4.2) to find , the stability condition becomes

d 3 dU
> 0,
r

dr dr r =

(4.6)

and the frequency of small oscillations, , satisfies

2 =

d 3 dU
r
.
dr dr r =
1

(4.7)

If we apply these results to the potential U ( r ) = kr 2 2, we find that

d 3 dU
= 4k 3 ,
r

dr dr r =

(4.8)

and
k

2 = 4 .

(4.9)

This in agreement with the result found above.


5. Apsidal precession

Two point masses interacting by Newtonian gravity orbit the system CM periodically.
The apsides occur at the same angular positions. There are a number of reasons why
planetary orbits are not exactly planar ellipses. The solar mass distribution has a small
deviation from isotropy due to solar rotation. Each planet also feels the gravitational
attraction of the others. Newtonian gravity is not an exact theory and there are small
deviations from the inverse square law due to effects of general relativity. The easiest of
these effects to take into account is the small correction due to general relativity. This
adds to the force a term proportional to r-4. The force remains a central force and angular
momentum conservation holds. The equation for the orbit when the small general
relativistic correction is included is
d 2u
Gm 2 M 3GM 2
+u =
+ 2 u .
(5.1)
d 2
l2
c
Here m is the mass of the planet, which is assumed small compared to M, the mass of the
Sun.

12
Because the general relativistic correction is small, this equation can be solved by a
perturbation method involving expansion in a dimensionless small parameter. To find a
convenient small parameter, consider a circular orbit in the absence of the GR term. The
solution to equation (5.1) is then
Gm 2 M
u =
.
l2
In this case, the ratio of the GR term to the Newtonian term is

(5.2)

G 2m2 M 2
.
(5.3)
c 2l 2
This ratio provides a convenient dimensionless parameter to use in the expansion. The
equation for the orbit is

=3

d 2u
u2
+
u
=

d 2

(5.4)

u = u0 ( ) + u1 ( ) +.

(5.5)

We look for a solution of form


Substituting this into equation (5.4), we have
2

d 2u1

d 2 u0
( u0 + u1 + ) .

u
u

+
+
+
+
=
+

0
1
2

d 2
d

(5.6)

Dropping terms of second order and higher, equation (5.6) becomes

d 2u1

d 2 u0
u0 2

.
+
u
+
+
u
=
+
2
0
1

d 2
d

(5.7)

Equating terms of the same order in , we have


d 2 u0
+ u0 = ,
d 2

(5.8)

u0 2
d 2u1
+
u
=
.
1
d 2

(5.9)

and

We have already found the solution for u0:

u0 ( ) = (1 + cos ) .

(5.10)

Using this in equation (5.9), we have

d 2u1
+ u1 = (1 + 2 cos + 2 cos 2 )
2
d
(5.11)
1 2
1 2

= 1 + + 2 cos + cos 2 .
2
2

This equation is similar to that for simple harmonic motion with a periodic forcing term.
Because the cos term on the right hand side has the same frequency as the solution of

13
the homogeneous equation it acts to increase the amplitude of the harmonic motion. The
particular integral is
1
1

(5.12)
u1, p = 1 + 2 + sin 2 cos 2 .
6
2

To order , the solution for u() is


1
1

(5.13)
u ( ) = (1 + cos ) + A cos + 1 + 2 + sin 2 cos 2 .
2
6

Once again we have applied the condition that = 0 is an apside. The constant A is
determined by the value of u at = 0. We can take this to be (1 + ) as before, so that
1
1

1
u ( ) = (1 + cos ) + 1 + 2 1 + 2 cos 2 cos 2 + sin . (5.14)
6
3
2

The apsides occur when u is stationary, i.e., when

1 du
1
1

= sin + 1 + 2 sin + 2 sin 2 + sin + cos = 0.


d
3
3

(5.15)

When = 0, the solutions are = n for integer n. Hence when 0, we look for
solutions of form = n + n . Substituting this into equation (5.15), and keeping only
terms that are first order in , we get

n = n .

(5.16)

Hence the position of the apsides are given by

= n (1 + ) .

(5.17)

Since the apsides alternate between perihelion and aphelion, the perihelion shifts by an
angle
2

6 GM
GmM
= 2 = 6
.
=
a (1 2 ) c 2
cl

Perihelion precession of Mercury

Amount (arcsec/century)
5025.6
531.4
0.0254
42.980.04
5600.0
5599.7

Cause
Coordinate (precession of the equinoxes)
Gravitational tugs of the other planets
Oblateness of the Sun
General relativity1
Total
Observed

14
Table taken from Wikipedia. Original source not cited by Wikipedia.

The precession of the equinoxes arises from tidal forces of Moon and Sun on nonspherical Earth. Earths rotation axis changes direction with a period of 25,800 years.
1 Einstein, Albert (1916). "The Foundation of the General Theory of Relativity". Annalen
der Physik 49: 769822.

You might also like