Sie sind auf Seite 1von 17

119

5 Central Force Motion



(Most of the material presented in this chapter is taken from Thornton and Marion, Chap.
8, and Goldstein, Poole, and Safko Chap. 3)
5.1 Introduction
In this chapter we will study the problem of two bodies moving under the influence of
each other. This problem is one which appears widely in physics. The orbits of planets
around the Sun are an example of the motion in the case of an attractive force (gravity).
The scattering of a beam of energetic protons by a solid (or more particularly, by their
nuclei) is an example.of this phenomenon where there is a repulsive force
(electromagnetic in nature, between the incoming proton and the (positive) target
nucleus).
This type of problem was solved by Newton for two bodies, but the problem of three or
more bodies interacting in this way has no simple analytical closed form. This means that
the motion cannot be expressed as simple finite equation, but only as an infinite series of
terms (but not one where one can confidently say that many or most of them are small,
and thus ignore them). This n-body problem can be usefully studied in certain special
cases however, or by using computer integration techniques (where we compute and plot
the motion and study it that way). However, here we will restrict ourselves to the two-
body problem, which can be solved analytically.
5.2 Reduction to the Equivalent One-body Problem the Reduced
Mass
We consider a system consisting of two point masses,
1 2
and m m , when the only forces
are those due to an interaction potential U . This potential is presumed to be generated by
the particles involved, and not some external phenomenon.

We will assume that U is a function only of the distance between the two particles
2 1
r = = r r r where
1 2
and r r are defined with respect to some arbitrary origin. So our
assumption at this point has not specified the form of the force except to say that it is
directly radially from one particle. Think about it: if the potential is only a function of r,
then the gradient (force) can only be directed directly towards or away from the particles.
Gravity and the forces between electric charges are examples of such forces.

Such a system has six degrees of freedom, three for each particle, since can each move in
three dimensions. We could choose, for example, the three components of each of the
two vectors,
1
r and
2
r (see Figure 5-1). However, since the potential energy is solely a
function of the distance between the two particles, i.e., ( ) U U r = , it is to our advantage if
we also express the kinetic energy as function of r (that is, of r ) rather than
1 2
and r r .
How do we do that? Well, lets cleverly first define the center of mass R of the system
as
120

( )
1 2 1 1 2 2
. m m m m + = + R r r (5.1)

We also consider the distance between each particle and the center of mass


1 1 2 2 2
1 1
1 2 1 2
1 1 2 2 1
2 2
1 2 1 2
,
m m m
m m m m
m m m
m m m m
+
= =
+ +
+
= =
+ +
r r
r R r r
r r
r R r r
(5.2)

with
2 1
= r r r .


Figure 5-1 The different vectors involved in the two-body problem.


We now calculate the kinetic energy of the system


( ) ( )
2 2
1 1 2 2
2 2
1 1 2 2
1 1
2 2
1 1
,
2 2
T m m
m m
= +
( (
= + + +

r r
R r R R r R



(5.3)

and inserting equations (5.2) in equation (5.3) we get


121

( )
2 2
2 1
1 2
1 2 1 2
2
2 2 2 2
1
1 2 1 2
2
2 2 1 1
2
1 2 1 2
2 2 1 2
1 2
1 2
1 1
2 2
1
2
2
1
2
2
1 1
.
2 2
m m
T m m
m m m m
m m
m
m m m m
m m
m
m m m m
mm
m m
m m
( (
= + +
( (
+ +

(
| |
( +
|
+ +
(
\

(
| |
( + + +
|
+ +
(
\

= + +
+
R r R r
R r R r
R r R r
R r



i

i


(5.4)

We introduce a new quantity , the reduced mass, defined as


1 2
1 2
,
mm
m m
=
+
(5.5)

which can alternatively be written as


1 2
1 1 1
.
m m
= + (5.6)

We can use equation (5.5) to write the kinetic energy as

( )
2 2
1 2
1 1
.
2 2
T m m = + + R r

(5.7)

We have, therefore, succeeded in expressing the kinetic energy as a function of r . So our
cleverness above has paid off! But wait, the kinetic energy actually has two components.
The first term on the right is the kinetic energy associated with the whole system moving
at the speed of the center of mass, the second term is the orbital kinetic energy on top of
that. That will be OK so we proceed. We are now in a position to write down the
Lagrangian for the central force problem, and in a form which will turn out to be
relatively easy to deal with.


( ) ( )
2 2
1 2
1 1
,
2 2
L T U
m m U r
=
= + + R r


(5.8)

where the potential energy ( ) U r is yet undefined, except for the fact that it is solely a
function of the distance between the two particles. However, seen from equation (5.8)
that the three components of the center of mass vector are cyclic. More precisely,

122
0,
L L L
X Y Z

= = =

(5.9)

with
x y z
X Y Z = + + R e e e (the centre of mass position). This means that the generalized
momentum associated with it is conserved. Wait you say! This is only true if a
generalized coordinate is cyclic in the Hamiltonian, not the Lagrangian! This is true,
however, if you find the Hamiltonian youll find that it is just T+U (the time t does not
appear explicitly anywhere), and the coordinates in question are cyclic.

From the perspective of our Lagrange formulation here we observe that Lagranges
equation is

( )
1 2
1 2
0
0
( ) 0
( ) constant
L d L
X dt X
d L L
dt X X
d
m m X
dt
m m X

=


= =

+ =
+ =

(5.10)

This equation is true for Y and Z as well. From this, we conclude that the center of mass
is therefore either at rest or moving uniformly. The equations of motion for , , and X Y Z
can be combined into the following vector relation

( )
1 2
0. m m + = R

(5.11)

Since the center of mass vector (and its derivative) does not appear anywhere else in the
Lagrangian, we can drop the first term of the right hand side of equation (5.8) in all
subsequent analysis and only consider the remaining three degrees of freedom. We keep
in mind that the whole system may have some overall constant motion of its centre of
mass, but its motion does not affect the orbital motion which we discuss next. In fact,
weve already solved three of the equations of motion, it just turns out they have very
simple solutions. The remaining Lagrangian is therefore

( )
2
1
.
2
L U r = r (5.12)

What is left of the Lagrangian is exactly what would be expected if we were dealing with
a problem of a single particle of mass subjected to a central force towards a central
particle which is fixed (immovable). [Think about it: what would the kinetic energy be of
the one particle that can move? What would the potential be? So what is L=T-U?]
Of course, this one-body problem is not really physical: there are no immovable bodies.
But the central force motion of two particles about their common center of mass is
123
reducible to a mathematically equivalent one-body problem, which may provide a useful
approximation to the case where one body is much more massive than the other, say.
However, despite the mathematical equivalence of the two-body problem to this one-
body problem, we need to keep in mind that the central body is not immobile in the two-
body case.
5.3 The First Integrals of Motion
Since we are dealing with a problem where the force involved is conservative, where the
potential is a function of the distance r of the reduced mass to the force center alone, the
system has spherical symmetry. Now consider that there is some relative motion between
the two particles (if they have zero relative velocity, they will simply attract or repel each
other along a straight line between them, but this is not a case that is usually of much
interest). Consider the velocity of one particle about the other. Since the force is always
directed radially, there is no force perpendicular to the plane defined by the particles
motion at their initial conditions. Therefore the bodies always remain in that same plane,
though they can of course move around in that plane. Equivalently, we could consider the
torque exerted by a radial force.
( ) 0 K

= =
N r F
N r r

where K is some multiplicative factor, positive (for repulsive forces) or negative (for
attractive ones). Because the force is radial, the torque is always zero and the (vector)
angular momentum is conserved. Since the vector angular momentum is always
perpendicular to the radius vector and the velocity, the constant of the angular momentum
means that the velocity always has to be in the same plane.

Since we have deduced that the motion is at all time confined to the aforementioned
plane we are, therefore, fully justified to use polar coordinates as the two remaining
generalized coordinates for this problem. Or equivalently, we can choose cylindrical
coordinates and set z (perpendicular to the plane) to be a constant since the motion is
restricted to a plane. Note that this is not a constraint in the sense weve used it before. It
falls out of the physics. Technically, the result that z=constant is one of the integrals of
motion or first integrals of motion. An integral of the motion reduces the number of
dimensions of the problem to be solved in the same way as a constraint does though, by
providing an expression which can replace one of the variables with a constant by simple
algebraic substitution.

Since the three components of the angular momentum vector are all constant, the angular
momentum actually provide three integrals of the motion. The constancy of the three
components of the linear momentum of the centre of mass (described earlier in this
chapter) provides another six integrals of motion (three for the velocity and three for the
initial position). The existence of these integrals is very important to the process of
solving these problems, again because they reduce the dimension of the equations to be
solved.

So, lets express our Lagrangian in polar coordinates. Since
r
r = r e , we have
124


.
r r
r
r r
r r

= +
= +
r e e
e e

(5.13)

(remember in taking the derivative of a vector, you have to take the derivative of the
basis vectors as well). In the last equation we have used the following relation (with the
usual transformation between the ( ) , r polar and the ( ) , x y Cartesian coordinates)


( ) ( )
( ) ( )
cos sin
sin cos ,
r x y
x y


= +
= +
e e e
e e e
(5.14)

from which it can be verified that

.
r r
r
r
r


= + =

e e
e e

(5.15)

We can now rewrite the Lagrangian as a function of and r


( ) ( )
2 2 2
1
.
2
L r r U r = +

(5.16)

We notice from equation (5.16) that is a cyclic variable. The corresponding
generalized momentum is therefore conserved, that is


2
constant.
L
p r

= =

(5.17)

The momentum p

is a constant and therefore is another integral of motion and is seen


(from elementary physics) to be equal to the magnitude of the angular momentum vector.
It is customarily written as


2
constant l r =

(5.18)

Since this just comes out of the conservation of angular momentum, this doesnt provide
another integral, its just another way of expressing one of the three integrals associated
with the angular momentum. But it will prove useful.

Before continuing to solve the motion of the particles, lets consider some of the
implications of conservation of angular momentum. Consider the problem in its one-body
guise. As the mobile particle moves along its trajectory through an infinitesimal angular
displacement d within an amount of time dt , the area dA swept out by its radius
vector r is given by
125


2 2
1 1
.
2 2
dA r d r dt = =

(5.19)

Alternatively, we can define the areal velocity as


2
1
2
constant.
2
dA
r
dt
l

=
= =

(5.20)

Thus, the areal velocity is constant in time. This result, discovered by Kepler for
planetary motion, is called Keplers Second Law. In essence, a radial line extending
between the two bodies sweeps out equal areas in equal amounts of time. It is important
to realize that the conservation of the areal velocity (or angular momentum) is a general
property of central force motion and is not restricted to the inverse-square law force
involved in planetary motion.
Another integral of motion (the only one remaining) concerns the conservation of energy.
The conservation is insured because we are considering conservative systems and there is
no explicit t in anything. Writing E for the energy we have


( ) ( )
2 2 2
1
,
2
E T U
r r U r
= +
= + +

(5.21)

or using equation (5.18)

( )
2
2
2
1 1
2 2
l
E r U r
r

= + + (5.22)

So we have ten integrals of motion so far, and it turns out there are another two. Since our
initial problem had six degrees of freedom (three per particle), we have a twelve
dimensional problem to solve, either six second-order differential equations (Lagranges
method), or 12 first-order differential equations (Hamiltons method), in order to get
position and velocity for each degree of freedom. The presence of 12 integrals of motion
will helps in the solution of the problem, as weve seen.

The presence or absence of integrals of the motion is one of the key aspects of whether
one can solve the problem at hand. For example, the three body problem has only 10
integrals of motion, but has 9 degrees of freedom and hence 18 dimensions, 18-10 leaves
8. As a result, a partial explanation of the difficulty of solving the three body problem is
that there arent enough integrals of motion. But well leave this topic for the time being.
126
5.4 The Equations of Motion
We will use two different ways for the derivation of the equations of motion. The first
one consists of inverting equation (5.22) and express r as a function of ( ) , , and E l U r
such that

( )
2
2 2
2
,
dr l
r E U r
dt r
= = (

(5.23)

or alternatively


( )
2
2 2
.
2
dr
dt
l
E U r
r
=
(

(5.24)

It turns out that equation (5.24) can be solved (that is, both sides can be integrated), once
the potential energy ( ) U r is defined, to yield the solution ( ) t t r = , or after inversion
( ) r r t = . This is now Newton solved the problem.

The second method considered here for solving the equations of motion uses the
Lagrange equations


0
0.
L d L
r dt r
L d L
dt

=


=

(5.25)

The second of these equations was already used to get equation (5.18) for the
conservation of angular momentum, so weve already solved it. We only have the second
one to solve now, but its a little tricky. Applying the first of equations (5.25) to the
Lagrangian (equation (5.16)) gives


( ) ( )
2
.
U
r r F r
r


=

(5.26)

So Lagranges equations give us the equation of motion quite easily. Solving it is a
problem which remains, and which Lagranges method doesnt help with, we have to use
the regular methods for solving differential equations. So to do that we now cleverly
modify this equation by making the following change of variable


1
. u
r
(5.27)

127
We calculate the first two derivatives of u relative to


2 2
1
2 2
1 1
,
du dr dr dt
d r d r dt d
r l
r
r r l

= =
| |
= =
|
\

(5.28)

where we have used the fact that
2
l r =

, and for the second derivative




2
2
2
2
2
.
d u d dt d r
r r
d d l d dt l l
r r
l

| | | |
= = =
| |
\ \
=

(5.29)

From this equation we have


2 2
2
2 2
,
l d u
r u
d
= (5.30)

and from equation (5.18)


2
2
2 3
2 2
.
l l
r r u
r


| |
= =
|
\

(5.31)

Inserting equations (5.30) and (5.31) in equation (5.26) yields


2
2 2 2
1 1
,
d u
u F
d l u u

| |
+ =
|
\
(5.32)

which can be rewritten as

( )
2 2
2 2
1 1 d r
F r
d r r l

| |
+ =
|
\
(5.33)

Though you might not know it to look at it, we now have an equation which is in a form
we can solve, but it still needs solving, and the solution will depend on the nature of the
force F(r).

128
5.5 Centrifugal Energy and the Effective Potential
In equations (5.23) to Error! Reference source not found. for , , and dr dt ,
respectively, appeared a common term containing different energies (i.e., the total and
potential energies and an extra term)

( )
2
2
.
2
l
E U r
r
(5.34)

The last term has the units of energy but what is it exactly? We first note that

2
2 2
2
1
.
2 2
l
r
r

=

(5.35)

which is the rotational kinetic energy, though NOT the total kinetic energy (which would
have to include a term with r ). So lets call this the energy of rotation.

It is interesting to note that if we arbitrarily define this quantity as a type of potential
energy
c
U , we can derive a conservative force from it. That is, if we set


2
2
,
2
c
l
U
r
(5.36)

then the force associated with it is


2
2
3
.
c
c
U l
F r
r r

= = =


(5.37)

The force defined by equation (5.37) is the so-called centrifugal force. It would,
therefore, be probably better to call
c
U is the centrifugal potential energy and to
include it with the potential energy ( ) U r to form the effective potential energy ( ) V r
defined as

( ) ( )
2
2
2
l
V r U r
r
+ (5.38)

This effective potential energy acts mathematically like a potential energy and so we can
usefully pretend that it is a potential for some purposes, which well examine below. In
fact, this is a good example of how forces are closely tied to energy and why the
Lagrangian and Hamiltonian formalisms work.

If we take for example the case of an inverse-square-law (e.g., gravity or electrostatic),
we have

129
( )
2
,
k
F r
r
= (5.39)

and

( ) ( ) .
k
U r F r dr
r
= =

(5.40)

The effective potential is
( )
2
2
.
2
k l
V r
r r
= + (5.41)
It is to be noted that the centrifugal potential reduces the effect of the inverse-square-
law on the particle. This is because the inverse-square-law force is attractive while the
centrifugal force is repulsive. This can be seen in Figure 5-2.
It is also possible to guess some characteristics of potential orbits simply by comparing
the total energy with the effective potential energy at different values of r . From Figure
5-3 we can see that the motion of the particle is unbounded if the total energy (
1
E ) is
greater than the effective potential energy for
1
r r when ( )
1 1
E V r = . This is because
the positions for which
1
r r < are not allowed since the value under the square root in
equation (5.34) becomes negative, which from equation (5.23) would imply an imaginary
velocity. For the same reason, the orbit will be bounded with
2 4
r r r for a total energy
2
E , where ( ) ( )
2 2 4
V r E V r . The turning points
2
r and
4
r are called the apsidal
distances. Finally, the orbit is circular with
3
r r = when the total energy
3
E is such that
( )
3 3
E V r = .






130

Figure 5-2 - Curves for the centrifugal, effective, and gravitational potential energies.

Figure 5-3 Depending on the total energy, different orbits are found.

131
5.6 Planetary Motion Keplers Problem
The equation for a planetary orbit can be calculated from equation
Error! Reference source not found. when the functional form for the gravitational
potential is substituted for ( ) U r

( )
( )
2
2
2
constant.
2
2
l r dr
r
k l
E
r r

= +
| |
+
|
\

(5.42)

This equation can be integrated if we first make the change of variable 1 u r = , the
integral then becomes


( )
( )
2 2
2
2
constant
2
2
constant,
2
l du
u
l u
E ku
du
E ku u
l

= +
| |
+
|
\
= +
+

(5.43)

the sign in front of the integral has not changed as it is assumed that the limits of the
integral were inverted when going from equation (5.42) to (5.43). We further transform
equation (5.43) by manipulating the denominator


( )
2
2 2
4 2 2
2
2 2
2 2
2
constant
2
constant
2
1 1
constant,
1
1
du
u
k E k
u
l l l
du
k El ul
l k k
du
u

= +
| |
+
|
\
= +
| | | |
+
| |
\ \
= +
| |

|
\

(5.44)

with


2 2
2
2
and 1 .
l El
k k


+ (5.45)

132
Now, to find the solution to the integral of equation (5.44), lets cleverly consider a
function ( ) f u such that

( ) ( ) cos . f u u = (

(5.46)

Taking a derivative relative to u we get

( ) sin ,
df d
du du

= (5.47)

or


( )
( )
2
2
1
sin
1
1 cos
1
.
1
d df
du du
df
du
df
du
f

=
=

(5.48)

Returning to equation (5.44), and identifying ( ) f u with the following

( )
1
,
u
f u

= (5.49)

we find that

( )
1
1
cos ,
u
u

| |
= +
|
\
(5.50)

or, since ( ) ( ) cos cos = ,

( ) 1 cos ,
r

= + + (5.51)

where is some constant. If we choose r to be minimum when 0 = , then 0 = and
we finally have our solution!

( ) 1 cos
r

= + (5.52)

133
with


2 2
2
2
and 1 .
l El
k k


+ (5.53)

Equation (5.52) is that of a conic section (circle, ellipse, parabola or hyperbola) with one
focus at the origin. So he vector from one body to the other draws out a conic section
over time. Note that this might involve both bodies actually moving significantly (if, for
example, they have similar masses). The motion of both bodies around the centre of
mass turns out to be the same conic section, but the more massive body does a
proportionately smaller ones. If one body is much more massive than the other (eg a
planet around the Sun) as is often the case, then the massive body barely moves and the
smaller body effectively traces an elliptical path around it (if its bound to the Sun, E<0,
of course).
The quantities and 2 are called the eccentricity and the latus rectum of the orbit,
respectively. The minimum value of r (when 0 = ) is called the pericenter, and the
maximum value for the radius is the apocenter. The turning points are apsides. The
corresponding terms for the turning points of the motion about the sun are perihelion and
aphelion, and for the earth, perigee and apogee.
As was stated when discussing the results shown in Figure 5-3, the energy of the orbit
will determine its shape. For example, we found that the radius of the orbit is constant
when
min
E V = . We see from, equation (5.53) that this also implies that the eccentricity is
zero (i.e., 0 = ). In fact, the value of the eccentricity is used to classify the orbits
according to different conic sections (see Figure 5-4):

1 > 0 E > Hyperbola
1 = 0 E = Parabola
0 1 < <
min
0 V E < <
Ellipse
0 =
min
E V =
Circle

For planetary motion, we can determine the length of the major and minor axes
(designated by 2 and 2 a b ) using equations (5.52) and (5.53). [The semimajor axis a and
the semiminor axis b are the terms usually used in astronomy rather than major and minor
axes]. The semimajor axis is half the long axis of the ellipse, and so we have


min max
2
1
2 2 1 1
.
1 2
r r
a
k
E

+ | |
= = +
|
+
\
= =

(5.54)

From this do you notice anything interesting? If two orbits have the same semimajor axis,
what is the different between their energies? What other factors come into play?

134

Figure 5-4 The shape of orbits as a function of the eccentricity.

If the eccentricity is zero, then the orbit is a circle and the semimajor axis is just the
radius of the circle. As increases, the orbit becomes more and more elongated. Note
that the Sun (or central body) is NOT at the centre of the ellipse, but at one focus. The
other focus is empty.

The period of the orbit can be evaluated using equation (5.20) for the areal velocity


2
. dt dA
l

= (5.55)

Since the entire area enclosed by the ellipse will be swept during the duration of a period
, we have


0 0
2
,
A
dt dA
l


=

(5.56)

or


2 2
, A ab
l l

= = (5.57)

where we have the fact that the area of an ellipse is given by ab . Now, substituting the
first of equations (5.53) and equation Error! Reference source not found. in equation
(5.57) we get

135

2
2 3
4
a
k

= (5.58)

In the case of the motion of a solar system planet about the Sun we have

,
p s
k Gm m = (5.59)

where , , and
p s
G m m are the universal gravitational constant, the mass of the planet, and
the mass of the Sun, respectively. We therefore get


( )
2
2 3
2
3
2 3
2
4
4
4
,
p s
p s p s
p s
s
m m
a
Gm m m m
a
G m m
a
Gm

=
+
=
+

(5.60)

since
s p
m m . We find that the square of the period is proportional to the semi-major
axis to the third, with the same proportionality constant for every planet. This
(approximate) result is known as Keplers Third Law. We end by summarizing Keplers
Laws
I. Planets move in elliptical orbits about the Sun with the Sun at one focus.
II. The area per unit time swept out by a radius vector from the Sun to a planet is
constant.
III. The square of a planets period is proportional to the cube of the major axis of the
planets orbit.

Das könnte Ihnen auch gefallen