You are on page 1of 11

Orbits in a central force field: Bounded orbits

Subhankar Ray∗
Dept of Physics, Jadavpur University, Calcutta 700 032, India

J. Shamanna
Physics Department, Visva Bharati University, Santiniketan 731235, India
(Dated: August 1, 2003)
arXiv:physics/0410149v1 [physics.ed-ph] 19 Oct 2004

The nature of boundedness of orbits of a particle moving in a central force field is investigated.
General conditions for circular orbits and their stability are discussed. In a bounded central field
orbit, a particle moves clockwise or anticlockwise, depending on its angular momentum, and at the
same time oscillates between a minimum and a maximum radial distance, defining an inner and an
outer annulus. There are generic orbits suggested in popular texts displaying the general features of
a central orbit. In this work it is demonstrated that some of these orbits, seemingly possible at the
first glance, are not compatible with a central force field. For power law forces, the general nature
of boundedness and geometric shape of orbits are investigated.

I. INTRODUCTION B. Newtonian Synthesis

The central force motion is one of the oldest and widely Almost 100 years later Newton realized that the plan-
studied problems in classical mechanics. Several familiar ets go about in their nearly circular orbits around the sun
force-laws in nature, e.g., Newton’s law of gravitation, under the influence of the same force that causes an ap-
Coulomb’s law, van-der Waals force, Yukawa interaction, ple to fall to the ground, i.e., gravitation. Newton’s law
and Hooke’s law are all examples of central forces. The of gravitation gave a theoretical basis to Kepler’s laws.
central force problem gives an opportunity to test one’s Kepler’s laws can be derived from Newton’s law of grav-
understanding of the Lagrange’s equation, Hamilton’s itation; this is often referred to as the Newtonian syn-
equation, Hamilton Jacobi method, and classical pertur- thesis. Kepler’s first and third laws are valid only in the
bation. It also serves as an introduction to the concept of specific case of inverse square force. There are, however,
integrals of motion and conservation laws. We need only certain general features which are observed in all cen-
to appeal to the principles of conservation of energy and tral field problems. They include (i) certain conserved
angular momentum to describe the nature and geometry quantities (energy, and angular momentum), (ii) planer
of the possible trajectories in central force motion. Most nature of orbits, and (iii) constancy of areal velocity (Ke-
books in classical mechanics1,2,3,4,5,6 , treatise7,8 , and ad- pler’s second law). A large class of central forces allows
vanced texts9,10 discuss the central force problem. In this circular orbits (stable or unstable), bounded orbits, and
article we present some interesting features of bounded even closed and periodic orbits. Certain common charac-
orbits in a central field. teristics about the generic shapes of bounded orbits can
also be ascertained.

A. Kepler’s Laws
II. EQUATIONS OF MOTION AND THEIR
One of the most remarkable discoveries in the history of FIRST INTEGRALS
physics is that of Keplerian orbits. A tremendous wealth
of data on planetary positions was collected by Tycho A. Central field orbits: confinement in a plane
Brahe and Johannes Kepler after detailed observation
spread over several decades. After a thorough analysis
The central force motion between two bodies about
of this data Johannes Kepler formulated three empirical
their center of mass can be reduced to an equivalent one
laws that described and correlated the motion of the five
body problem in terms of their reduced mass m and their
planets then known:
relative radial distance r. Hence in this reduced system,
1. Each planet moves in an elliptical orbit, with the a body having the reduced mass moves about a fixed
sun at one of its foci. center of force.
Consider the motion of a body under a central force,
2. The radius vector from the sun to each planet F = F(r) = f (r)r̂ with the origin as its force center. The
sweeps out equal areas in equal times. potential V (r) from which this force is derived is also a
function of r alone, F = −∇V, V ≡ V (r).
3. The square of the periods (T 2 ) of the planets are On account of the central nature of the force, the me-
proportional to the cube of the lengths of the cor- chanical properties of the body do not vary under ro-
responding semimajor axes (a3 ). tation in any manner around the center of force. Let
2

the body be rotated through an infinitesimal angle δθ,


where the magnitude δθ is the angle of rotation while
the direction is that of the axis of rotation n̂. The
change in the radius vector from the origin to the body
is | δr |= rsin(θ)δθ, with δr being perpendicular to r r r dθ
and δθ. Hence δr = δθ × r. The change in velocity is
similarly given by δv = δθ × v. The Lagrangian of the
system is a function of r and ṙ. The motion is governed
by the Lagrange’s equation, Area = 1/2 (r) ( r d θ )
 
d ∂L ∂L
dt ∂ ṙ

∂r
=0 dθ
If we now require that the Lagrangian of the system re-
main invariant under this rotation we obtain r
∂L ∂L FIG. 1: Area swept by radius vector
δL = · δr + · δ ṙ = 0 .
∂r ∂ ṙ
One can define generalized momentum as,
C. First integrals and conservation laws
∂L
p=
∂ ṙ The canonical momentum corresponding to θ is called
From the Lagrange’s equation we get, the angular momentum (or rather the magnitude of the
  angular momentum that we discussed before),
. d ∂L ∂L
ṗ = =
dt ∂ ṙ ∂r ∂L
pθ = = mr2 θ̇ = l
Replacing ∂L/∂ ṙ by p and ∂L/∂r by ṗ in the equation ∂ θ̇
for δL we get, As θ is a cyclic coordinate, i.e., the Lagrangian is inde-
ṗ · δθ × r + p · δθ × ṙ = 0 . pendent of θ, this angular momentum is conserved. This
can be shown from the Lagrange’s equation for θ.
d  
δθ · (r × p) = 0 . d ∂L ∂L
dt − = 0
dt ∂ θ̇ ∂θ
As δθ is arbitrary we conclude l = r × p is a conserved
d
quantity. l is called the angular momentum of the sys- p˙θ = (mr2 θ̇) = 0.
tem. Since l is a constant and is perpendicular to r it dt
follows that the radius vector of the particle lies in a The corresponding integral of motion is,
plane perpendicular to l. This implies that the motion
of the particle in a central field is confined to a plane. mr2 θ̇ = l (4)

and this l can easily be shown to be the magnitude of the


B. Lagrangian and equations of motion angular momentum vector l = r × p.
This conservation law is essentially equivalent to Ke-
As the motion in a central force field is confined to a pler’s 2nd Law:
plane, it suffices to use plane polar coordinates. One may The elementary triangular area swept by the radius
write the Lagrangian of the particle as, vector in an infinitesimal time interval dt is,
1 2 1 1
L= mṙ − V (r) = m(ṙ2 + r2 θ̇2 ) − V (r) . (1) dA = r(rdθ)
2 2 2
We assume the center of force to be at the origin. The
hence it follows from Eq. (4) that the rate of areal sweep
coordinates of the body of mass m undergoing the central
is a constant.
field motion are given by (r, θ).
The Lagrange’s equation for the θ and r coordinates dA 1 l
are given respectively by, = r2 θ̇ = (5)
dt 2 2m
 
d ∂L ∂L There is another first integral of motion associated
− =0 (2)
dt ∂ θ̇ ∂θ with the Lagrange’s equation for the r coordinate.
 
d ∂L ∂L
− =0 (3) d ∂V
dt ∂ ṙ ∂r (mṙ) − mrθ̇2 + =0
dt ∂r
3

11 60

10
40
9
kinetic energy
velocity
8 20

7
Angular momentum
0
6
total energy
5
-20

3 -40
potential energy

2
-60
radial position
1

0 -80
0 1 2 3 4 5 6 0 1 2 3 4 5 6

FIG. 2: Constancy of angular momentum in an inverse square FIG. 3: Constancy of total energy in an inverse square poten-
potential. Total energy is 60% of the minimum of effective tial. Total energy is 60% of the minimum of effective potential
potential Ṽ (r). Ṽ (r).

The force in terms of the potential (conservative force) D. Equation for the orbit
is given by, f (r) = −∂V /∂r, whence the above equation
becomes From the equation giving energy as a first integral of
motion, we get the expression for radial velocity,
mr̈ − mrθ̇2 = f (r) s  
Using the first integral of motion, one can convert this 2 l2
ṙ = E−V − (7)
second equation into an equation for r alone. m 2mr2

1 l2 On integration we get,
 
d
mr̈ = − V + Z r
dr 2 mr2 dr
t= p . (8)
Integrating we get r0 2/m(E − V − l2 /(2mr2 ))

1 2 1 l2 This relation can be inverted to give r as a function of t,


mṙ + + V (r) = E (6) r = r(t). The other first integral gives,
2 2 mr2
where E is a constant of integration, called the energy. l
. θ̇ =
This is the law of conservation of total mechanical energy. mr2
It is interesting to note that for motion in a general Substituting the expression for r as a function of t, and
force field, l = mr2 θ̇ remains invariant even though r and integrating,
θ̇ vary with time (see Fig. 2). Similarly E = mṙ2 /2 + Z t
l2 /(2mr2 )+V (r) remains a constant even though ṙ and r l dt
θ − θ0 = . (9)
and hence mṙ2 /2 and V (r) + l2 /(2mr2 ) each varies with m 0 r(t)2
time (see Fig. 3). These two expressions for r(t) and θ(t) express the equa-
Lagrange’s equations are two second order Ordinary tion of the orbit for the particle in a central field in terms
Differential Equations (ODE) in r and θ. However they of time t as a parameter.
decouple, i.e., each equation is expressible in terms of ei- Instead of expressing the orbit parametrically in terms
ther r or θ. On integrating each equation once we get of t, one often wants to express the orbit directly as an
the first integrals of motion namely, the total mechanical equation connecting r and θ. Such an equation may be
energy and the angular momentum. A further integra- obtained by eliminating t from the above expressions for
tion will yield the complete solution to the problem. This ṙ and θ̇.
second integration introduces two more constants of in-
tegration namely, the initial radial (r0 ) and angular (θ0 ) dθ l/(mr2 )
= p
positions. dr 2/m(E − V (r) − l2 /(2mr2 ))
4

30
On integration this yields,
Z r
l/(mr2 ) l^2/(2mr^2) (a) circular orbit
θ − θ0 = p dr 20 (b) elliptic orbit
(c) elliptic orbit of higher eccentricity
r0 2/m(E − V (r) − l2 /(2mr2 ))
For understanding the qualitative nature of motion in 10
a central field one looks at the equivalent one dimensional ~
V(r)
problem.
r 0
2 l2
ṙ = (E − V (r) − )
m 2mr2
r -10 (c)
2
= (E − Ṽ (r)). (b)
m
(a)
-20
We call Ṽ the effective potential, introduced to make
the problem similar to that of a particle moving in a
one dimensional potential field. The effective radial force -30
V(r) ~ - 1/r
(f˜(r)) is connected to the effective radial potential (Ṽ (r))
by the expected relation,
-40
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
r
∂ Ṽ (r)
f˜(r) = − (10)
∂r FIG. 4: Effective potential for an inverse square force

At a point where the effective potential Ṽ (r) equals the


energy E, the radial velocity vanishes (ṙ = 0). In one For 0 > E > Ṽ (r0 ) there is a range of radial positions
dimensional motion this corresponds to a particle com-
(rmax ≥ r ≥ rmin ) for which E ≥ Ṽ (r). The particle can
ing momentarily to rest, and having zero kinetic energy.
move in this range of r with varying ṙ. The radial velocity
However in the case of central field, the motion is not
ṙ vanishes at the end points rmax and rmin where the
really one dimensional, and even for ṙ = 0, the particle
energy E equals the effective potential Ṽ and ṙ reaches a
is not at rest (v = rθ̇ θ̂), and it has a non-zero kinetic
maximum at r = r0 The points rmax and rmin are called
energy ((1/2m)r2 θ̇2 ). the turning points. The central field particle cannot move
For the inverse square force, as in the case of gravita- beyond these points, as the energy becomes less than the
tion, we have f (r) = −k/r2 and V (r) = −k/r. Effective effective potential, and the expression for radial velocity
potential (see Fig. 4) is given by, (ṙ) turns imaginary.
k l2 E < Ṽ (r0 ) is a physically impossible situation, since
Ṽ (r) = − + (11) no (radial) position is physically allowed for the particle.
r 2mr2
For E ≥ 0 we have an unbounded motion, where the
The following properties of the effective potential are particle can fly off to infinity. Thus in the case of inverse
easily noted: square force field we can have bounded (E < 0) or un-
1. Ṽ (r) = 1/r2 (l2 /2m − kr), as r → 0 the term bounded (E ≥ 0) motion depending on the energy of the
within bracket is essentially l2 /(2m) and hence particle. In particular for E = 0 it is parabolic and for
E > 0 it is hyperbolic.
limr→0 Ṽ (r) → +∞
2. limr→∞ Ṽ (r) = 0
III. EXISTENCE AND STABILITY OF
3. Ṽ (r) = −1/r(k − l2 /2mr2 ) and for large values of CIRCULAR ORBITS FOR CENTRAL FORCES
r, l2 /2mr2 is negligible compared to k, hence Ṽ (r)
has a negative value. At all positions other than where the effective potential
∗ 2 is a minimum or a maximum, we have a net effective force
4. At r = l /2mk the function Ṽ (r) intersects the r
f˜(r) = −∂ Ṽ /∂r 6= 0. When the total energy E is not
axis, i.e., Ṽ (r∗ ) = 0.
equal to the minimum of effective potential Ṽ (r), at the
5. Ṽ (r) reaches a minimum at r0 = 2 · r∗ = l2 /mk. points of instantaneous zero radial velocity the particle
With ∂ Ṽ /∂r|r0 = 0 and ∂ 2 Ṽ /∂ 2 r|r0 > 0. is pushed away. If the effective potential has a minimum
and energy is greater than that minimum then the radial
For total energy E = Ṽ (r0 ) the particle has zero radial distance has a lower and an upper bound and the particle
velocity at r = r0 , and no other radial position is phys- moves between these radial limits. On the other hand if
ically accessible, since ṙ becomes imaginary at r 6= r0 . the effective potential has a maximum, and the energy is
This corresponds to circular motion. less than that maximum, the effective force pushes the
5

2
(a) spiral in
the effective potential has a minimum, the effective forces
(b) circle cause the particle to remain in a bound orbit confined
(c) spiral out
1.5 in an annular space. For small deviations the annular
radii are nearly the same, and hence the orbit is close to
1
a circle. One can understand the stability question by
studying the forces (restoring or unsettling) that come
into play when the system is moved infinitesimally from
0.5
the position of circular orbit.
The circular orbit is stable if,
0
∂ 2 Ṽ
> 0. (16)
-0.5
∂r2
(a)
That is equivalent to,
(b)
-1
∂ 2 Ṽ 3l2
 
∂f
= − + >0
(c) ∂r2 ∂r mr4 r0
-1.5
∂f 3l2
|r0 <
-2
∂r mr0 4
-1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
Using Eq. (13),
FIG. 5: Spiralling orbits, f (r) ∼ −1/r 4
∂f 3f (r0 )
|r < − (17)
∂r 0 r0
particle away from the positions of zero radial velocity
in an inward or an outward spiral. If the total energy
IV. NATURE OF BOUNDED ORBITS
is equal to the maximum or minimum of the effective
potential, the system can stay with zero radial velocity,
and hence move in a circular orbit. General bounded motion has both lower and upper
The condition for circular orbit is, bounds. It means that the particle cannot approach
nearer than some minimum or move farther than some
maximum distance. One has to remember that the angu-

l2

∂ Ṽ ∂V
= − =0 (12) lar velocity has a constant sign, same as that of the con-
∂r ∂r r0 mr03

r0 stant angular momentum, throughout the motion. How-
ever its magnitude decreases with increase in the radial
whence we get, distance (∼ 1/r2 ). Together with the angular motion, the
radial distance changes from rmax to a rmin , then back
l2

∂V to rmax and so on. From this general nature of motion
f (r0 ) = − =− 3 (13)
∂r r0
mr0 two families of generic orbits are suggested in popular
texts11 (Fig. 6, Fig. 7). However it can be shown that
The negative sign on the right hand side clearly shows the generic types shown in the later figure (Fig. 7) are
that the force must be attractive. The particle moves in not feasible for any attractive potential.
a circular orbit, since the force of attraction due to the We consider below a general bounded orbit confined in
central field provides the necessary centripetal force. an annular region. We would like to investigate the na-
For a given attractive central force ( f (r) and V (r) ture of the orbit close to the point where it touches the
given ) it is possible to have a circular orbit of radius inner or the outer annulus. Let us choose the reference
r0 provided the angular momentum and energy of the line (polar line) of the coordinate system such that the
particle are given by, orbit touches the annular ring at θ = 0, or more specifi-
cally at r = R, θ = 0. Since at this point P (r = R, θ = 0)
l2 = −mr03 f (r0 ) (14) the orbit is at its closest (or farthest) approach from the
l2 pole, (∂r/∂t)P = 0. As r is a function of θ, using Eq. (4)
E = V (r0 ) + (15)
2mr02 we get for motion along the trajectory,

However when the effective potential has a maximum dr l dr


=
the system is in an unstable circular orbit. A small devi- dt mr2 dθ
ation from this radial position causes the orbit to be un- hence
bounded. The effective force that comes into play makes
the particle move away from the position of zero radial dr
r′ (0) = =0 (18)
velocity in an inward or outward spiral (Fig. 5). When dθ θ=0
6

2 1.5

1.5
1

0.5
0.5

0 0

-0.5
-0.5

-1

-1
-1.5

-2 -1.5
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2 -1.5 -1 -0.5 0 0.5 1 1.5
1.5 1.5

1 1

0.5 0.5

0 0

-0.5 -0.5

-1 -1

-1.5 -1.5
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5

FIG. 6: Generic orbits in central field (allowed) FIG. 7: Generic orbits in central field (not allowed)

The equation of motion for r is given by, One can expand r(θ) in a Taylor series in θ, and remem-
ber that r′ (0) = 0,
mr̈ − mrθ̇2 = f (r)

From which we get r′′ (θ) along the trajectory, θ


r(θ)traj = r(0) + r′ (0)θ + r′′ (0) + h.o.
2!
l2
 
l d l dr 4
mr(0) ˜ θ 2
m 2 − mr = f (r) = r(0) + f (r(0)) + h.o.
mr dθ mr2 dθ mr3 2
 l 2!
mr(0)4
 2
d2 r mr4 ˜ θ
= f (r) = r(0) + r(0) + f (r(0)) + h.o.
dθ2 l2 l2 2!
mr4
= r + 2 f (r) Consider a tangent to the annulus at the point P , Fig.
l
7

r( θ) r( θ)
tan tan
r( θ)
Inner traj trajectory
Outer r( θ) trajectory
annulus annulus
θ (not allowed) θ traj
(not allowed)

trajectory
r (allowed)
rannul annul

tangent tangent
trajectory trajectory
(allowed) (not allowed)

FIG. 8: Curvature of the trajectory in comparison to the FIG. 9: Curvature of the trajectory in comparison to the outer
tangent to the inner annulus annulus

8, we study the orbit near the point P where it touches the


outer annulus.
r(0)
= cos(θ) mr(0)4 ˜ (∆θ)2
r(θ) r(∆θ)traj = r(0) + f (r) + h.o.
l2 2!
which can be expanded to give r(θ)tan along the tangent, hence

θ2 r(∆θ)traj < r(0) = rmax (20)


r(θ)tan = r(0) + r(0) + h.o.
2 This confirms that the outer annulus is indeed the outer
Notice that the constant or the θ independent terms in bound of the trajectory. The nature of the orbit near the
r(θ)traj and r(θ)tan are equal, and the first leading power point of contact P at the outer annulus is shown in figure
of θ is θ2 in both the cases. The coefficient of θ2 (cur- 9.
vature) will determine the nature of the trajectory in
reference to the tangent.
We first study the curvature of the orbit in reference V. BOUNDED ORBITS FOR THE POWER LAW
CENTRAL FORCE
to the tangent. The force being attractive f (r(0)) < 0,
for small angular distance (∆(θ)) away from the point P ,
A. Existence of stable circular orbit

r(∆θ)traj < r(∆θ)tan (19) For the case of a power law central potential

which means that the orbit bends more sharply than the k nk
V (r) = , f (r) = − (21)
tangent and stays closer to the inner annulus for nearby rn rn+1
points (Fig. 8). Hence the second type of orbits shown
From the stability condition Eq. (17),
in some texts (Fig. 7) are not possible for central forces.
For the outer annulus the analysis with respect to tan- ∂f 3f (r0 )
gent is not very meaningful as one can set up an even |r0 < −
∂r r0
stronger bound for its orbit, directly considering the po-
tential Ṽ (r). In this case the effective potential has a we find,
positive slope with respect to the radius, and hence a  
(n + 1)nk k 1
negative effective force (f˜(r) < 0). The orbit should not < −3 −
only remain nearer than the tangent, but even nearer r0n+2 r0n+1 r0
than the outer annulus. This condition is confirmed if n < 2 (22)
8

40
For the point of extremum
amn
r0 n−2 = .
30 l2
l^2/(2m r^2)
Finding the second derivative of Ṽ with respect to
r at the point r0 ,
20

d2 Ṽ an(n + 1) 3l2
(a) unstable circular orbit 2
= − n+2
+
dr r mr4
10
an(n + 1) 3l2

(b) unbounded orbit 1
= 4 − +
~ r rn−2 m
V(r)
l2 3l2
 
0 1
= 4 −(n + 1) +
r0 m m
1 l2
-10
V(r) ~ - 1/r^3 = 4 (2 − n) < 0
r0 m

Hence it is a point of maximum. For E > Ṽ (r0 ) we


-20
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 always have an unbounded orbit. For E < Ṽ (r0 ) the
orbit is semibounded, bounded above or bounded below,
FIG. 10: Effective potential for f (r) ∼ −1/r 4 force according to its initial state. The particle either spirals in
or spirals out. For E = Ṽ (r0 ) we get an unstable circular
orbit.
Hence the circular orbit is stable for an attractive power
law potential that varies slower than inverse square (or a
force that varies slower than inverse cube). 2. V (r) = −a/r n where n = 2

The effective radial potential is Ṽ (r) = −a/r2 +


B. Study of boundedness of orbits l /(2mr2 ). This is essentially an attractive or repulsive
2

inverse square term.


1. V (r) = −a/r n where n > 2
ã l2
Ṽ (r) = − , ã > 0 if a >
Consider the case when n > 2. We have the effective r2 2m
potential, l2
ã < 0 if a <
2m
a l2
Ṽ (r) = − + (23) This potential cannot give circular orbit ever. If the ef-
rn 2mr2
fective potential is attractive and E < 0 it has an upper
1. Ṽ (r) = −1/rn (a − l2 · rn−2 /(2m)), and as r → 0 bound of radial distance. A typical orbit would therefore
we may neglect l2 · rn−2 /(2m) in comparison to a, be an inward spiral. For a repulsive effective potential
we will have outward spiral moving to infinite radial dis-
thus limr→0 Ṽ (r) = −∞.
tance.
2. limr→∞ Ṽ (r) = 0.
3. V (r) = −a/r n where 2 > n > 0
3. For large but finite r, Ṽ (r) = −1/r2 (a/rn−2 −
l2 /(2m)), and a/rn−2 is negligible compared to
l2 /(2m), and hence Ṽ (r) is positive. In this case the effective potential is,

4. Ṽ (r) intersects the r axis at r∗ where (r∗ )n−2 = a l2


Ṽ (r) = − + (24)
2am/l2. rn 2mr2
We have the following properties of Ṽ (r)
5. Ṽ (r) has a maximum at r0 where r0n−2 = anm/l2.
For a maximum point of Ṽ we should have 1. Ṽ (r) = 1/r2 (l2 /2m−a/rn−2 ) and as r → 0 we may
∂ Ṽ /∂r |r0 = 0 and ∂ 2 Ṽ /∂r2 |r0 < 0. neglect a/rn−2 in comparison to l2 /2m, and thus
limr→0 Ṽ (r) → +∞.
∂ Ṽ an l2
= n+1 − =0. 2. limr→∞ Ṽ (r) = 0.
∂r r mr3
9

30 30

20 20

l^2/(2m r^2) l^2/(2m r^2)

10 10
~
V(r)

0 0

-10 ~ -10
V(r)

-20 -20
V(r) ~ ln(r)

V(r) ~ -1/r
-30 -30

-40 -40
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

FIG. 11: Effective potential for f (r) ∼ −1/r 2 force FIG. 12: Effective potential for f (r) ∼ −1/r force

3. Ṽ (r) = −1/rn (a − l2 /2mr2−n ) and for large but 4. V (r) = a ln r


finite values of r, l2 /2mr2−n is negligible compared
to a, hence Ṽ (r) is negative. This corresponds to one over r force f (r) ∼ −a/r. The
effective potential is given by,
4. Ṽ (r) intersects the r axis at r∗ where (r∗ )2−n =
l2 /(2am). l2
Ṽ (r) = a ln r + (25)
2mr2
5. Ṽ (r) reaches a minimum at r0 where r0 2−n =
l2 /amn. We have the following properties of Ṽ (r)
For a minimum point of Ṽ we should have
∂ Ṽ /∂r |r0 = 0 and ∂ 2 Ṽ /∂r2 |r0 > 0. 1. limr→0 Ṽ (r) → +∞.

∂ Ṽ an l2 2. limr→∞ Ṽ (r) → +∞.


= n+1 − =0.
∂r r mr3 3. There is no point of intersection with the r axis,
For the point of extremum and Ṽ (r) is always positive.

l2 4. Ṽ (r) reaches a minimum at r0 = l/ am.
r0 2−n = . For a minimum point of Ṽ we should have
amn
∂ Ṽ /∂r |r0 = 0 and ∂ 2 Ṽ /∂r2 |r0 > 0.
Finding the second derivative of Ṽ with respect to
r at the point r0 , ∂ Ṽ a l2
= −
∂r r mr3
d2 Ṽ an(n + 1) 3l2
2
= − n+2
+ for the point of extremum
dr r mr4
an(n + 1) 3l2

1 l2
= 4 − + r02 =
r rn−2 m am
2
3l2
 
1 l
= 4 −(n + 1) + The second derivative of Ṽ at r = r0 ,
r0 m m
1 l 2 d2 Ṽ a 3l2
= 4 (2 − n) > 0 2
= − 2+
r0 m dr r mr4
3l2

1
= 2 −a +
Hence it is a point of minimum. For this case the orbit r mr2
can be bounded or unbounded depending on the total 1
energy E of the particle. = 2 · 2a > 0
r
10

30
The second derivative of Ṽ with respect to r,
d2 Ṽ n−2 3l2
25 = an(n − 1)r +
dr2 mr4
3l2
 
1
= 4 an(n − 1)rn+2 +
20 r m
1 l2
= 4 (2 + n) > 0
15 r m
Hence it is a point of minimum. The orbit is always
10 V(r) ~ r^2 bounded.
~
V(r) The findings for the general power law potential can
be summarized as follows,
5
V (r) = sign(n)arn n 6= 0 (27)
l^2/(2m r^2)
f (r) = −abs(n)arn−1 (28)
0

and
-5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
V (r) = b ln r (29)
b
f (r) = − (30)
FIG. 13: Effective potential for f (r) ∼ −r force r

Hence it is a point of minimum. The orbit is always TABLE I: Dependence of boundedness on power law.
bounded. n of V (r) ∼ ar n nature of boundedness

) − ∞, −2( always unbounded

−2 spiralling orbit
5. V (r) = ar n where n > 0
) − 2, 0( bounded or unbounded depending on E
For this potential the effective potential is given by,
0 (ln r) always bounded
2
l
Ṽ (r) = arn + n>0 (26) )0, +∞( always bounded
2mr2

We have the following properties of Ṽ (r)


C. Stable bounded orbit, geometric shape
1. limr→0 Ṽ (r) → +∞.
The stable bounded orbits in a power law central field
2. limr→∞ Ṽ (r) → +∞. can have the following forms
1. V (r) = −a/rn , 0 < n ≤ 2
3. There is no point of intersection with the r axis,
and Ṽ (r) is always positive. 2. V (r) = b log r
3. V (r) = a · rn , n > 0
4. Ṽ (r) has a minimum at r0 , r0n+2 = l2 /amn.
For a minimum point of Ṽ we should have In all the above cases the derivative of V (r) with respect
∂ Ṽ /∂r |r0 = 0 and ∂ 2 Ṽ /∂r2 |r0 > 0. to r is always positive, and hence the force is necessarily
attractive (f (r) = −∂V (r)/∂r < 0).
∂ Ṽ l2 In the first case the effective one dimensional potential
= anrn−1 − Ṽ (r) goes to infinity as r → 0. Ṽ (r) has a minimum at
∂r mr3
some r = r0 , and it has a negative slope for all r < r0 ,
for the point of extremum and positive slope for all r > r0 . For E = Ṽ (r0 ) we get
stable circular orbit, and for Ṽ (r0 ) < E < 0. the orbits
l2 are still bounded. For cases (2) and (3) the circular orbit
r0 n+2 = is stable, and orbits are always bounded for any energy.
amn
11

orbits are sometime mistakenly identified with diagrams


TABLE II: Dependence of stability of circular orbits on power
of the form shown in Fig. 711 . The above expression in
law.
fact corresponds to a class of orbits that look generically
n of V (r) ∼ ar n stability of circular orbits
like those shown in Fig. 6.
) − ∞, −2) unstable
The generic features of central force orbits discussed
here have been verified by computer simulation for a large
) − 2, ∞( stable class of central force fields. The figures shown here (Fig.
2, 3, 5 and 6) were generated by these simulations. Stu-
dents interested in studying and generating such orbits
VI. CONCLUSION will find Ref.12 helpful.

Existence of bounded orbit for a large class of attrac-


Acknowledgement
tive central field has been discussed. The generic nature
of central field bounded orbits is analytically derived.
Certain class of these orbits (Fig. 7) presented in popular The authors wish to express their indebtedness to the
texts1 , are shown to be non-feasible. well known texts by Goldstein1 , Landau2 and Arnold9 .
Small deviation from circularity in the case of central They also acknowledge their teachers in related graduate
field is often expressed in terms of inverse of radial dis- courses at Stony Brook, Prof. Max Dresden, Prof. A. S.
tance (u = 1/r). Goldhaber and Prof. Leon A. Takhtajan.
Authors gratefully acknowledge the encouragement re-
u = u0 + a · cos(βθ) ceived from Prof. Shyamal SenGupta of Presidency Col-
r0 lege, Calcutta. The material presented here was used in
r = a graduate level classical mechanics course at Jadavpur
1 + a · r0 · cos(βθ)
University during 1998-2001. SR wishes to thank his stu-
where r0 = 1/u0 . It is interesting to note that these dents for stimulating discussions.

8

Electronic address: subho@juphys.ernet.in L.A. Pars, Introduction to Dynamics, Cambridge, 1953.
1 9
H. Goldstein, Classical Mechanics, Addison-Wesley, 1980. V.I. Arnold, Mathematical Methods of Classical Mechanics,
2
L.D. Landau, E.M. Lifshitz, Mechanics, Pergamon Press, Springer Verlag, 1989.
10
1976. R. Abraham, and J.E. Marsden, Foundations of Mechan-
3
K.R. Symon, Mechanics, Addison-Wesley, 1971. ics, Benjamin-Cummings Publ.Co.Inc., 1978.
4 11
D.T. Greenwood, Classical Dynamics, Prentice Hall, 1977. H. Goldstein, Fig. 3-7, Fig. 3-13 and Eq. 3-45 in Classical
5
J.L. Synge, and B.A. Griffith, Principles of Mechanics, Mc- Mechanics, Addison-Wesley, 1980.
12
GrawHill, 1970. H. Gould, J. Tobochnik, An Introduction to Computer
6
A. Sommerfeld, Mechanics, Academic Press, 1952. Simulation Methods, Addison-Wesley, 1988.
7
E.T. Whittaker, A Treatise on the Analytical Dynamics of
Particles and Rigid Bodies, Dover, 1944.

You might also like