Вы находитесь на странице: 1из 23

1

PHYS620 Fall 2018 Notes - 3

1. Central force motion


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  m1r1  r1  m2r2  r2  U  r2  r1  . (3.1.1)
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. (3.1.2)
In terms of the particle separation,
r  r2  r1 , (3.1.3)
we have
m2
r1   r, (3.1.4)
m1  m2
and
m1
r2  r. (3.1.5)
m1  m2
The Lagrangian is then
2 2
1  m2  1  m1 
L  m1   r  r  m2   r  r  U  r 
2  m1  m2  2  m1  m2 
(3.1.6)
1 m1m2
 r  r  U  r  .
2 m1  m2
The quantity
m1m2
 , (3.1.7)
m1  m2
is called the reduced mass of the system.
We have reduced the motion of two particles to an equivalent problem of a single
particle of mass μ acted on by a force directed along the line from the particle to a fixed
point.
2

The generalized momentum conjugate to the position vector r has components


L
pi  , (3.1.8)
ri
and hence
p   r. (3.1.9)
The angular momentum
L  r  p, (3.1.10)
is orthogonal to both r and p. Hence, the motion is in the plane containing the origin and
has normal parallel to L.
The motion has two degree of freedom and can be described by using polar co-
ordinates (r, ). In these co-ordinates, the Lagrangian is
1
 
L   r 2  r 2 2  U  r  .
2
(3.1.11)

Since this does not depend explicitly on , i.e.  is ignorable, there is a first integral of
motion
L
p     r 2  l , (3.1.12)

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  , (3.1.13)
dt 2 dt 2 
conservation of angular momentum leads to Kepler’s second law of orbital motion (a
result that is independent of the particular dependence of the force on r).

d
r

O

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

1  2 2 l  
2

E    r  r  2    U  r 
2    r  
(3.1.14)
2
1 2 1 l
  r   U r .
2 2 r 2
Re-arranging to give an expression for r, we find
2 E U  l2
r    2 2. (3.1.15)
  r
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  r
. (3.1.16)

Using equation (3.1.12), we can eliminate the angular velocity to get


l2 U
 
r  . (3.1.17)
r 3
r
Also from equation (3.1.12), we have
dr  dr l
r   , (3.1.18)
d d  r 2
and
d  dr l  d  dr l  l d  dr l 
r
  2 
  2 
 2  . (3.1.19)
dt  d  r  dt  d  r   r d  d  r 2 
From the form of the last term in equation (3.1.19), we see that it is now useful to make
the substitution
1
u . (3.1.20)
r
We now have
l2 d 2u
r 
 u2 , (3.1.21)
 2
d 2
so that equation (3.1.17) becomes
d 2u  U
u   2 . (3.1.22)
d 2
l u
This equation is easily solved for a gravitational or Coulomb force.

2. Planetary motion
To investigate the motion of a planet around a star (both treated as point masses),
U  ku, (3.2.1)
where for the gravitational force
4

k  Gm1m2 . (3.2.2)
Equation (3.1.22) for this particular case is
d 2u k
u   2 , (3.2.3)
d 2
l
which has solution
k
u  A cos    0  . (3.2.4)
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.
Equation (3.2.4) for the orbit can be written as
1 Gm1m2 
 A cos   , (3.2.5)
r l2
or in an alternative form as

 1   cos  , (3.2.6)
r
where
l2
 , (3.2.7)
Gm1m2 
and  is a constant.
It can be shown that equation (3.2.6) 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.
The energy of the orbit is
2
1  dr l  1 l 2 Gm1m2
E   2 
 
2  d  r  2  r 2
r
2
1  du l  1 l 2 2
    u  Gm1m2u (3.2.8)
2  d   2 
1 l2

2 
2   2  1 .

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 (3.2.6) is that of an ellipse for   1, consider the following
construction.
5

P
Latus rectum
r

O2 O1 rmin

The minimum value of r occurs when  = 0, and the maximum value occurs when  = π.
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  1  2
From the cosine rule the distance between O2 and P is
 rmax  rmin   r 2  2  rmax  rmin  r cos 
2

2 2
          
  2 2 
   2  2 2   cos 
 1     1   cos    1     1   cos  


1  2 cos     . 2

1    1   cos 
2

Hence the sum of the lengths O1P and PO2 is


 rmax  rmin   r 2  2  rmax  rmin  r cos 
2
r




1  2 cos    2

1   cos   1   2  1   cos  
  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 Kepler’s first Law.
Once we know that the orbit is an ellipse, we relate the parameters α and  to the
lengths of the semi-major and semi-minor axes. Clearly
6

 2 
2a  rmin  rmax  
.  (3.2.9)
1  1  1  2
Also by considering the right triangle AO1B in the figure below

B
b a

O2 A O1 rmin

we see that
a 2  b 2   a  rmin  ,
2
(3.2.10)
and so
2 2
r r  r r 
b 2   max min    max min 
 2   2 
 rmax rmin
(3.2.11)
2

1  2
 1   2  a 2 ,
which contains the usual definition of eccentricity.
The energy of the orbit can be written in terms of the lengths of the axes by using
equations (3.2.7), (3.2.8), and (3.2.9). We find that
1 l 2 1   
2
1 l2 1 l2 1 1 Gm1m2
E
2 2   2
 1  
2  
 
2 l 2
 a

2 a
. (3.2.12)
 
 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 . (3.2.13)
a
Using this last expression, we can write the semi-minor axis in terms of the constants of
the motion
l2
b2  . (3.2.14)
2 E
7

Since the area of the ellipse is  ab, Kepler’s second law enables us to find the orbital
period
  Gm1m2 l2 
P   ab    Gm1m2 . (3.2.15)
A l 2E 2 E 2E
3

2
By squaring and using equation (3.2.12), we find the more familiar form of Kepler’s third
law
4 2
P2  a3. (3.2.16)
G  m1  m2 
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   a 1    , (3.2.17)
1 
and

rmax   a 1    . (3.2.18)
1 
Hence
rmax  rmin  2a, (3.2.19)

rmax  rmin  2a , (3.2.20)


and
rmax  rmin
 . (3.2.21)
rmax  rmin
Finally, note that the distance of a focus from the center of the ellipse is
 r  r
AO1  max min  a . (3.2.22)
2

3. The effective potential


The total energy
1 2 1 l2
E  r  U r , (3.2.23)
2 2 r 2
is the same as that of a particle of mass μ experiencing one dimensional motion with
potential energy
1 l2
U eff  r    U  r . (3.2.24)
2 r2
8

Hence, Ueff(r) is an effective potential energy. The force corresponding to the effective
potential energy is
dU eff l2 dU
Feff  r     3 . (3.2.25)
dr r dr
The term l 2   r 3  is commonly (but incorrectly) 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, (3.2.26)
2
with equality only when U eff  r   E.
For bound orbits, the 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.
The effective potential energy is also useful for studying the effects of
perturbations to the system.

EXAMPLE.

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
oscillation for a small perturbation from equilibrium (that does not change the angular
momentum).

Solution. The effective potential energy is


1 l2 1
U eff  r    kr 2 .
2 r 2
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 the value of r at
which 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 
9

where
12
 kl 2 
U0    .
  
A plot of U eff U 0 against s looks like
10

6
Ueff/U0

0
0 1 2 3 4 5

Clearly, 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
2E 2
s4  s  1  0,
U0
which has solutions
2
E  E 
s 
2
    1.
U0  U0 
The equation of motion is
l2
 
r  kr  0,
r3
which in terms of s is
10

 1
s
 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
k
2 .

4. Stability of circular orbits


For a general potential, the equation of radial motion is
l2 dU
 
r
 . (3.3.1)
r 3
dr
For a given angular momentum, circular orbits exist only if there are real solutions of
dU l2
 . (3.3.2)
dr  r 3
Suppose there is a solution at r = r0. We can test for stability by considering a small
perturbation such that r = r0 + r. The equation of motion is
l2 dU d 2U
 
r   r0   2  r0   r  
  r0   r 
3
dr dr
(3.3.3)
 3l 2 d 2U 
   4  2  r0    r  .
  r0 dr 
Dropping terms of second order and higher, and assuming a harmonic time behavior, the
frequency of small oscillations, , satisfies
3l 2 d 2U
 2   r . (3.3.4)
 r0 4 dr 2 0
For stability  has to be real. The circular orbit is stable if
3l 2 d 2U
  r   0. (3.3.5)
 r0 4 dr 2 0
Using equation (3.3.2) to find r0 , the stability condition becomes
d  3 dU 
r   0, (3.3.6)
dr  dr  r  r0
and the frequency of small oscillations, , satisfies
11

1 d  3 dU 
2  r . (3.3.7)
 r03 dr  dr  r  r0

If we apply these results to the potential U  r   kr 2 2, we find that


d  3 dU 
r   4kr03 , (3.3.8)
dr  dr  r  r0
and
k
2  4 . (3.3.9)

This in agreement with the result found above.

5. Apsidal precession
Two point masses interacting by Newtonian gravity orbit the system CM
periodically and so the apsides occur at the same angular positions. However, 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 a term proportional to r-4 to the force. 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 Gm2 M 3GM 2
 u   2 u . (3.4.1)
d 2 l2 c
Here m is the mass of the planet, which is assumed small compared to M, the mass of the
Sun.
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 (3.4.1) is then
Gm2 M
u   . (3.4.2)
l2
In this case, the ratio of the GR term to the Newtonian term is
G 2m2 M 2
 3 . (3.4.3)
c 2l 2
This ratio provides a convenient dimensionless parameter to use in the expansion. The
equation for the orbit is
12

d 2u u2
 u     . (3.4.4)
d 2 
We look for a solution of form
u  u0    u1    . (3.4.5)
Substituting this into equation (3.4.4), we have
 u0  u1   .
2
d 2 u0  d 2u1 
 u    2  u1        (3.4.6)
d  d 
2 0

Dropping terms of second order and higher, equation (3.4.6) becomes
d 2 u0  d 2u1  u0 2
 u    2  u1     . (3.4.7)
d 2  d 
0

Equating terms of the same order in λ, we have
d 2 u0
 u0   , (3.4.8)
d 2
and
d 2u1 u0 2
 u  . (3.4.9)
d 2 
1

We have already found the solution for u0:


u0     1   cos   . (3.4.10)
Using this in equation (3.4.9), we have
d 2u1
 u1   1  2 cos    2 cos 2  
d 2
(3.4.11)
 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
the homogeneous equation it acts to increase the amplitude of the harmonic motion. The
particular integral is
 1 1 
u1, p   1   2   sin    2 cos 2  . (3.4.12)
 2 6 
To order λ, the solution for u() is
 1 1 
u     1   cos      A cos   1   2   sin    2 cos 2  . (3.4.13)
 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
13

 1  1  1 
u     1   cos     1   2  1   2  cos    2 cos 2   sin   .
 2  3  6 
(3.4.14)
The apsides occur when u is stationary, i.e., when
1 du  1  1 
  sin    1   2  sin    2 sin 2   sin    cos    0. (3.4.15)
 d  3  3 
When   0, the solutions are   n for integer n. Hence when   0, we look for
solutions of form   n   n . Substituting this into equation (3.4.15), and keeping only
terms that are first order in , we get
 n  n . (3.4.16)
Hence the position of the apsides are given by
  n 1    . (3.4.17)
Since the apsides alternate between perihelion and aphelion, the perihelion shifts by an
angle
2
 GmM  6 GM
  2  6    .
 cl  a 1   2  c 2

6. Anomalous Perihelion precession of Mercury


The bulk of the precession of the perihelion of Mercury is due to the precession of
the equinoxes that arises from tidal forces of Moon and Sun on non-spherical Earth.
Earth’s rotation axis changes direction with a period of 25,800 years. The difference
between the observed amount and that due to the precession of the equinoxes is called the
anomalous precession. The contributions to the anomalous precession are given in the
table.

Amount (arcsec/century) Cause


532.3035 Gravitational tugs of the other planets
0.0286 Oblateness of the Sun
42.9779 General relativity1
575.31 Total predicted
574.10 ± 0.65 Observed

1 Einstein, Albert (1916). "The Foundation of the General Theory of Relativity". Annalen
der Physik 49: 769–822.
14

EXAMPLE PROBLEMS
1. A particle moves in a circular orbit in a force field given by
F (r )  k / r 2 .
Show that, if k suddenly decreases to half its original value, the particle’s orbit becomes
parabolic.

Solution: Let the radius of the circular orbit be r0. The initial energy is
1𝑘 𝑘 1𝑘
𝐸 𝑇 𝑈 .
2𝑟 𝑟 2𝑟
Assuming that the kinetic energy does not change when k is reduced to k/2, the final
energy is
1𝑘 𝑘
𝐸 𝑇 𝑈 0.
2𝑟 2𝑟
Hence, the orbit becomes parabolic.

2. Two gravitating masses m1 and m2 (M = m1 + m2) are separated by a distance r0 and


released from rest. Show that when the separation is r, the speeds are

𝑣 𝑚 , 𝑣 𝑚 .

Solution: Since the particles are released from rest, the angular momentum is zero and the
masses move along a straight line. In the CM frame the energy is
1 𝐺𝑚 𝑚 𝐺𝑚 𝑚
𝐸 𝜇𝑟 .
2 𝑟 𝑟
Here µ is the reduced mass. Hence
1 𝐺𝑚 𝑚 𝐺𝑚 𝑚 1 1
𝜇𝑟 𝐺𝑚 𝑚 .
2 𝑟 𝑟 𝑟 𝑟
The speeds of the particles are related to the magnitude of the relative velocity by
𝑚
𝑣 |𝑟|,
𝑀
𝑚
𝑣 |𝑟|.
𝑀
Hence
𝑚 2𝐺𝑚 𝑚 1 1 2𝐺 1 1
𝑣 𝑚 ,
𝑀 𝜇 𝑟 𝑟 𝑀 𝑟 𝑟

𝑚 2𝐺𝑚 𝑚 1 1 2𝐺 1 1
𝑣 𝑚 .
𝑀 𝜇 𝑟 𝑟 𝑀 𝑟 𝑟
15

3. A communications satellite is in a circular orbit around Earth at radius R and velocity


v. A rocket accidentally fires quite suddenly, giving the rocket an outward radial velocity
v in addition to its original velocity. (a) Calculate the ratio of the new energy and angular
momentum to the old. (b) Describe the subsequent motion of the satellite and plot T(r),
V(r), U(r), and E(r) after the rocket fires.

Solution: Let the masses of the satellite and Earth be m and M. Since the Earth is much
more massive than the satellite, we ignore the small difference between m and the
reduced mass.

(a) Since the impulse is directed radially, the angular momentum l = mRv is unchanged.
The initial energy is
1 𝐺𝑀𝑚
𝐸 .
2 𝑅
The impulse increases the kinetic energy by an amount equal to the initial kinetic energy.
Hence the energy after the impulse is
1 𝐺𝑀𝑚 1 𝐺𝑀𝑚
𝐸 𝐸 𝑇 0.
2 𝑅 2 𝑅
Hence the ratios are E1/E0 = 0, and l1/l0 = 1.

(b) The satellite follows a parabolic orbit after the impulse. The energy of the satellite
after the impulse is
1 1 𝑅 𝐺𝑀𝑚
𝐸 𝑟 𝑚𝑟 𝑚 𝑣 0.
2 2 𝑟 𝑟
Hence the kinetic, gravitational potential and effective potential energies are
𝐺𝑀𝑚 𝐺𝑀𝑚 𝑅
𝑇 𝑟 ,
𝑟 𝑅 𝑟
𝐺𝑀𝑚 𝐺𝑀𝑚 𝑅
𝑉 𝑟 ,
𝑟 𝑅 𝑟
and
1 𝑅 𝐺𝑀𝑚
𝑈 𝑟 𝑚 𝑣 .
2 𝑟 𝑟
Since the initial orbit is circular
𝐺𝑀
𝑣 ,
𝑅
and so
1 𝑅 𝐺𝑀𝑚 𝐺𝑀𝑚 𝐺𝑀𝑚 1 𝑅 𝑅
𝑈 𝑟 .
2𝑟 𝑅 𝑟 𝑅 2𝑟 𝑟
These energies are plotted below against r/R.
16

1.0

0.8 T
V
U
0.6 E

0.4
Energy in units of GMm/R

0.2

-0.0
0 1 2 r/R 3 4 5 6 7 8 9 10
-0.2

-0.4

-0.6

-0.8

-1.0

4. Consider a comet moving in a parabolic orbit in the plane of the Earth’s orbit. If the
distance of closest approach of the comet to the Sun is βrE, where rE is the radius of
Earth’s (assumed) circular orbit, show that the time the comet spends within the orbit of
the Earth is given by
2 1 𝛽 ∙ 1 2𝛽 3𝜋 1 year.
If the comet approaches the Sun to the distance of the perihelion of Mercury, how many
days is it within Earth’s orbit.

Solution: Since the comet’s orbit is a parabola, it has eccentricity equal to 1. The equation
of the orbit is
2𝛽𝑟
1 cos 𝜃.
𝑟
Hence, the comet is inside the Earth’s orbit if |𝜃| 𝜃 ≡ cos 2𝛽 1 . The area swept
out by the position vector of the comet while it is inside the Earth’s orbit is
1 1 1
𝐴 𝑟 𝑑𝜃 4𝛽 𝑟 𝑑𝜃 4𝛽 𝑟 𝑑𝜃
2 1 cos 𝜃 2 cos 𝜃⁄2

𝛽 𝑟 sec 𝜃⁄2 𝑑𝜃 .

On making the substitution 𝑥 tan 𝜃⁄2, this is



𝑥
𝐴 2𝛽 𝑟 1 𝑥 𝑑𝑥 2𝛽 𝑟 𝑥 ,
3
where 𝑥 tan 𝜃 ⁄2 .
17

Since
𝜃
cos 𝜃 2 cos 1 2𝛽 1,
2
we find
1 𝛽
tan 𝜃 ⁄2 .
𝛽
Hence
2 1 2𝛽 1 𝛽
𝐴 𝛽 𝑟 .
3 𝛽 𝛽
We can find the time derivative of A by considering the energy and angular momentum
of the orbit,
𝐸 1 1 𝐺𝑀
0 𝑟 𝑟 𝜃 .
𝑚 2 2 𝑟
Hence, at perihelion
1 𝐺𝑀
0 𝛽 𝑟 𝜃 ,
2 𝛽𝑟
and so the angular momentum per unit mass if the orbit is
2𝐺𝑀
𝑙 𝑟 𝜃 𝛽 𝑟 2𝐺𝑀𝛽𝑟 .
𝛽 𝑟
The rate at which the area is swept out is l/2. Hence the time inside Earth’s orbit is
1 2𝛽 1 𝛽
4𝛽 𝑟 𝛽 𝛽 4 𝑟
𝜏 1 2𝛽 1 𝛽.
3 2𝐺𝑀𝛽𝑟 3 2𝐺𝑀𝑟
The year is the orbital period of the Earth. A similar analysis shows that
2𝜋𝑟
1 year .
𝐺𝑀𝑟
Hence
𝜏 4 1 1
1 2𝛽 1 𝛽 1 2𝛽 2 1 𝛽 .
1 year 3 2𝜋√2 3𝜋

If the comet’s closest approach to the Sun is at same distance as the perihelion of
Mercury, then β = 0.3871, and so τ = 0.2084 years =76.12 days.

5. A particle of unit mass moves from infinity along a straight line that, if continued,
would allow it to pass a distance 𝑏√2 from a point P. If the particle is attracted towards P
with a force varying as 𝑘⁄𝑟 , and if the angular momentum about P is √𝑘 ⁄𝑏, show that
the trajectory is given by
18

𝑟 𝑏coth 𝜃⁄√2 .

Solution: Since the force is attractive, the potential energy is 𝑈 . The equation for
the trajectory is
𝑑 𝑢 𝑘
𝑢 𝑢 𝑏 𝑢 ,
𝑑𝜃 𝑙
where l is the angular momentum of the particle about P. Multiplying by 𝑑𝑢⁄𝑑𝜃, we get
𝑑𝑢 𝑑 𝑢 1 𝑑 𝑑𝑢 𝑑𝑢
𝑏 𝑢 𝑢 ,
𝑑𝜃 𝑑𝜃 2 𝑑𝜃 𝑑𝜃 𝑑𝜃
which on integrating gives
1 𝑑𝑢 𝑢 𝑢
𝑏 𝐶,
2 𝑑𝜃 4 2
where C is a constant of integration. We can use conservation of energy and conservation
of angular momentum to find C. Let the speed at infinity be v0. The angular momentum is
𝑙 𝑏√2𝑣 √𝑘 ⁄𝑏.
Hence
𝑘 1
𝑣 .
2𝑏
Conservation of energy gives
1 𝑘 1 1𝑙 𝑘
𝑣 𝑟 .
2 4𝑏 2 2𝑟 4𝑟
Hence, the distance of closest approach to P, rp, is given by
𝑘 1 𝑘 𝑘
.
4𝑏 2𝑏 𝑟 4𝑟
Solving for rp, we get
𝑟 𝑏.
At closest approach 𝑑𝑢⁄𝑑𝜃 0. Hence
1 𝑏 1
𝐶 .
2𝑏 4𝑏 4𝑏
The equation for the orbit is then
𝑑𝑢 1 1
𝑏 𝑢 2𝑏 𝑢 1 𝑏 𝑢 1 .
𝑑𝜃 2𝑏 2𝑏
Hence
𝑑𝑢 𝑏 𝑢 1
.
𝑑𝜃 𝑏√2
Make the substitution 𝑏𝑢 tanh 𝜙 to get
1 1 𝑑𝜙 tanh 𝜙 1
,
𝑏 cosh 𝜙 𝑑𝜃 𝑏√2
19

which gives
𝑑𝜙 1
,
𝑑𝜃 √2
so that
1
𝑏𝑢 tanh 𝜃 𝜙 ,
√2
where ϕ0 is a constant of integration. If we take θ = 0 to correspond to when the particle
is at infinity, then ϕ0 = 0, and so
1
𝑟 𝑏 coth 𝜃 .
√2
The trajectory is shown below. The orbit asymptotically becomes circular with radius b.
y

1.0

0.5

0.0
2 -1 0 1 2 x 3

-0.5

-1.0

6. A particle moves under the influence of a central force given by F  r    k r n . If the


particle’s orbit is circular and passes through the force center, show that n = 5.

Solution: Let the orbit have radius a. The situation is as shown below:
20


O a
C

Here C is the center of the circle and O is the force center. Since the triangle is a right
triangle, the equation for the orbit is
r  2a cos  .
The energy of the orbit is
1 k
2

E  m r 2  r 2 2  
 n  1 r n1
,

and the angular momentum is


l  mr 2.
Using the equation for the orbit,
1 l2 k
E m  4a 2 sin 2   4a 2 cos 2   2 
m  2a cos    n  1 2a cos  
4 n 1
2
l2 k
  .
8a m cos   n  1 2a n 1 cos n 1 
2 4

The energy is a constant of the motion, and hence must be independent of θ. This is
possible only if n = 5 and E = 0. The latter requirement leads to
km
l2  2 .
8a

7. A particle moves in an elliptical orbit in an inverse-square law central force field. If the
ratio of the maximum angular velocity to the minimum angular velocity of the particle in
its orbit is n2, then show that the eccentricity of the orbit is
n 1
 .
n 1
21

Solution: Let the angular velocity be . The angular momentum of the orbit is conserved.
Hence, the maxima and minima of the angular velocity occur at periapsis and apoapsis,
respectively, and are related by
rmax 2min  rmin 2max .
From this, we get
rmax
n .
rmin
The eccentricity is
rmax
1
rmax  rmin rmin n 1
   .
rmax  rmin rmax n  1
1
rmin

8. Discuss the motion of a particle moving in an attractive central-force field described by


F  r    k r 3 . Sketch some of the orbits for different values of the total energy. Can
circular orbits be stable in such a force field?

Solution: Assume that the particle has unit mass. The energy of the orbit is
1 1 k
E  r 2  r 2 2  2 .
2 2 2r
Using conservation of angular momentum
r 2  l ,
to eliminate  and replace the time derivative by a derivative with respect to  , we get
2 2
1  dr l  1 l 2 k 1  du  1 2
E    2  l2    l  k  u ,
2
2 
2  d r  2 r 2
2r 2  d  2
where u  1 r .
We see that if l 2  k  0, then the orbit is unbound. By differentiating the energy
expression with respect to  , we get
d 2u l 2  k
 2 u  0,
d 2 l
which has solutions
 u0 cos     0  , if l 2  k ,

u   u0  A    0  , if l 2  k ,

u0 cosh     0  , if l  k .
2

In these expressions, u0, A and  0 are constants, and


22

l2  k
2  .
l2
The total energy in each case is
1 2
 2  l  k  u0 , if l  k ,
2 2


 1 2
E kA , if l 2  k ,
 2
1 2
 2  l  k  u0 , if l  k .
2 2


The plot below shows three orbits for cases with l 2  k . The orbit with the lowest energy
is shown by the black line and the orbit with the highest energy is shown by the blue line.

The plot below shows an orbit for a case with l 2  k . The orbit is a spiral.
23

The plots below shows three orbits for cases with l 2  k . These orbits are bound.

Circular orbits are only possible if l 2  k . In this case the effective potential is identically
zero. Hence, there is no restoring force to stabilize the orbit. We conclude that the
circular orbits are unstable. A radial perturbation will give a spiral trajectory as in the
second of the three plots above.

9. An Earth satellite has a perigee of 300 km and an apogee of 3,500 km above Earth’s
surface. How far above the Earth is the satellite when (a) it has rotated 90 around Earth
from perigee and (b) when it has moved halfway from perigee to apogee?

(b)
Solution: The radius of the
Earth is 6371 km. Hence the (a)
perigee and apogee
distances are rmin = 6671 and
rmax = 9871 km respectively. Perigee Apogee

The semi-major axis is the


average of these two values.
Hence a = 8271 km. The
eccentricity of the orbit is
r r
  max min  0.1934.
rmax  rmin

(a) Using the equation for the orbit


a 1   2 
 1   cos  ,
r
we find that when   90 , r  a 1   2   7961 km. Hence the satellite is 1590 km above
the Earth’ surface.

(b) When the satellite is halfway from perigee to apogee, its distance from the center of
the Earth is a. Hence the satellite is 1900 km above the Earth’ surface.

Вам также может понравиться