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

Not logged in Talk Contributions Create account Log in

Article Talk Read Edit View history Search Wikipedia

Classical central-force problem


From Wikipedia, the free encyclopedia

Main page In classical potential theory, the central-force problem is to determine the motion of a particle in a single central
Contents potential field. A central force is a force (possibly negative) that points from the particle directly towards a fixed
Featured content point in space, the center, and whose magnitude only depends on the distance of the object to the center. In
Current events
many important cases, the problem can be solved analytically, i.e., in terms of well-studied functions such as
Random article
trigonometric functions.
Donate to Wikipedia
Wikipedia store The solution of this problem is important to classical mechanics, since many naturally occurring forces are
central. Examples include gravity and electromagnetism as described by Newton's law of universal gravitation
Interaction
and Coulomb's law, respectively. The problem is also important because some more complicated problems in
Help classical physics (such as the two-body problem with forces along the line connecting the two bodies) can be
About Wikipedia
reduced to a central-force problem. Finally, the solution to the central-force problem often makes a good initial
Community portal
approximation of the true motion, as in calculating the motion of the planets in the Solar System.
Recent changes
Contact page Contents [hide]

Tools 1 Basics
What links here 1.1 Definition of a central force
Related changes 1.2 Potential energy
Upload file 1.3 One-dimensional problem
Special pages 1.4 Uniform circular motion
Permanent link
1.5 Relation to the classical two-body problem
Page information
2 Qualitative properties
Wikidata item
2.1 Planar motion
Cite this page
2.2 Polar coordinates
Print/export 2.3 Specific angular momentum
Download as PDF 2.4 Constant areal velocity
Printable version
2.5 Equivalent parallel force field
Languages 3 General solution
Français 3.1 Binet equation
Edit links 3.2 Orbit of the particle
3.3 Turning points and closed orbits
4 Specific solutions
4.1 Kepler problem
4.2 Central forces with exact solutions
4.3 Revolving orbits
5 Historical development
5.1 Newton's derivation
6 Alternative derivations of the equations of motion
6.1 Lagrangian mechanics
6.2 Hamiltonian mechanics
6.3 Hamilton-Jacobi equation
7 See also
8 Notes
9 References
10 Bibliography
11 External links

Basics [ edit ]

The essence of the central-force problem is to solve for the position r[note 1] of a particle moving under the
influence of a central force F, either as a function of time t or as a function of the angle φ relative to the center of
force and an arbitrary axis.

Definition of a central force [ edit ]


See also: Central force and Newton's notation

A conservative central force F has two defining properties.[1] First, it must


drive particles either directly towards or directly away from a fixed point in
space, the center of force, which is often labeled O. In other words, a central
force must act along the line joining O with the present position of the
particle. Second, a conservative central force depends only on the distance r
between O and the moving particle; it does not depend explicitly on time or
other descriptors of position.

This two-fold definition may be expressed mathematically as follows. The An attractive central force
acting on a body at position P
center of force O can be chosen as the origin of a coordinate system. The (shown in red). By definition, a
vector r joining O to the present position of the particle is known as the central force must point either
position vector. Therefore, a central force must have the mathematical towards a fixed point O (if
attractive) or away from it (if
form[2] repulsive).

where r is the vector magnitude |r| (the distance to the center of force) and r ̂ = r/r is the corresponding unit
vector. According to Newton's second law of motion, the central force F generates a parallel acceleration a
scaled by the mass m of the particle[note 2]

For attractive forces, F(r) is negative, because it works to reduce the distance r to the center. Conversely, for
repulsive forces, F(r) is positive.

Potential energy [ edit ]


See also: Kinetic energy and Potential energy

A central force is always a conservative force; the magnitude F(r) of a central force can always be expressed as
the derivative of a time-independent potential energy function U(r)[3]

Thus, the total energy of the particle—the sum of its kinetic energy and its potential energy U—is a constant;
energy is said to be conserved . To show this, it suffices that the work W done by the force depends only on
initial and final positions, not on the path taken between them.

Equivalently, it suffices that the curl of the force field F is zero; using the formula for the curl in spherical
coordinates,

because the partial derivatives are zero for a central force; the magnitude F does not depend on the angular
spherical coordinates θ and φ.

Since the scalar potential V(r) depends only on the distance r to the origin, it has spherical symmetry. In this
respect, the central-force problem is analogous to the Schwarzschild geodesics in general relativity and to the
quantum mechanical treatments of particles in potentials of spherical symmetry.

One-dimensional problem [ edit ]


If the initial velocity v of the particle is aligned with position vector r, then the motion remains forever on the line
defined by r. This follows because the force—and by Newton's second law, also the acceleration a—is also
aligned with r. To determine this motion, it suffices to solve the equation

One solution method is to use the conservation of total energy

Taking the reciprocal and integrating we get:

For the remainder of the article, it is assumed that the initial velocity v of the particle is not aligned with position
vector r, i.e., that the angular momentum vector L = r × m v is not zero.

Uniform circular motion [ edit ]


Every central force can produce uniform circular motion, provided that the initial radius r and speed v satisfy the
equation for the centripetal force

If this equation is satisfied at the initial moments, it will be satisfied at all later times; the particle will continue to
move in a circle of radius r at speed v forever.

Relation to the classical two-body problem [ edit ]


See also: Two-body problem

The central-force problem concerns an ideal situation (a "one-body problem")


in which a single particle is attracted or repelled from an immovable point O,
the center of force.[4] However, physical forces are generally between two
bodies; and by Newton's third law, if the first body applies a force on the
The positions x1 and x2 of second, the second body applies an equal and opposite force on the first.
two bodies can be expressed in Therefore, both bodies are accelerated if a force is present between them;
terms of their relative separation
r and the position of their center there is no perfectly immovable center of force. However, if one body is
of mass Rcm. overwhelmingly more massive than the other, its acceleration relative to the
other may be neglected; the center of the more massive body may be treated
as approximately fixed.[5] For example, the Sun is overwhelmingly more massive than the planet Mercury;
hence, the Sun may be approximated as an immovable center of force, reducing the problem to the motion of
Mercury in response to the force applied by the Sun. In reality, however, the Sun also moves (albeit only slightly)
in response to the force applied by the planet Mercury.

Such approximations are unnecessary, however. Newton's laws of motion


allow any classical two-body problem to be converted into a corresponding
exact one-body problem.[6] To demonstrate this, let x 1 and x 2 be the
positions of the two particles, and let r = x 1 − x 2 be their relative position.
Then, by Newton's second law,

The final equation derives from Newton's third law; the force of the second
body on the first body (F 21 ) is equal and opposite to the force of the first
Any classical two-body
body on the second (F 12 ). Thus, the equation of motion for r can be written
problem to be converted into an
in the form equivalent one-body problem.
The mass μ of the one
equivalent body equals the
reduced mass of the two original
where is the reduced mass
bodies, and its position r equals
the difference of their positions.

As a special case, the problem of two bodies interacting by a central force can be reduced to a central-force
problem of one body.

Qualitative properties [ edit ]

Planar motion [ edit ]


See also: Angular momentum

The motion of a particle under a central force F always remains in the plane
defined by its initial position and velocity.[7] This may be seen by symmetry.
Since the position r, velocity v and force F all lie in the same plane, there is
never an acceleration perpendicular to that plane, because that would break
the symmetry between "above" the plane and "below" the plane.

To demonstrate this mathematically, it suffices to show that the angular


momentum of the particle is constant. This angular momentum L is defined Illustration of planar motion.
by the equation The angular momentum vector
L is constant; therefore, the
position vector r and velocity
vector v must lie in the yellow
where m is the mass of the particle and p is its linear momentum.[note 3] plane perpendicular to L.
Therefore, the angular momentum vector L is always perpendicular to the
plane defined by the particle's position vector r and velocity vector v.[note 4]

In general, the rate of change of the angular momentum L equals the net torque r × F [8]

The first term m v × v is always zero, because the vector cross product is always zero for any two vectors
pointing in the same or opposite directions. However, when F is a central force, the remaining term r × F is also
zero because the vectors r and F point in the same or opposite directions. Therefore, the angular momentum
vector L is constant. Then

Consequently, the particle's position r (and hence velocity v) always lies in a plane perpendicular to L.[9]

Polar coordinates [ edit ]


See also: Polar coordinate system

Since the motion is planar and the force radial, it is customary to switch to
polar coordinates.[9] In these coordinates, the position vector r is
represented in terms of the radial distance r and the azimuthal angle φ.

Taking the first derivative with respect to time yields the particle's velocity
vector v
The position vector r of a
point P in the plane can be
specified by its distance r from
the center (the origin O) and its Similarly, the second derivative of the particle's position r equals its
azimuthal angle φ. The x and y
Cartesian components of the acceleration a
vector are r cos φ and r sin φ,
respectively.

The velocity v and acceleration a can be expressed in terms of the radial and azimuthal unit vectors. The radial
unit vector is obtained by dividing the position vector r by its magnitude r, as described above

The azimuthal unit vector is given by[note 5]

Thus, the velocity can be written as

whereas the acceleration equals

Specific angular momentum [ edit ]


See also: Specific relative angular momentum

Since F = ma by Newton's second law of motion and since F is a central


force, then only the radial component of the acceleration a can be non-zero;
the angular component a φ must be zero

Therefore,

The specific angular


momentum h equals the speed
This expression in parentheses is usually denoted h v times r⊥, the component of the
position vector r perpendicular
to the velocity vector v. h also
equals the radial distance r
times the azimuthal component
which equals the speed v times r⊥, the component of the radius vector vφ of the velocity. Both of these
perpendicular to the velocity. h is the magnitude of the specific angular formulae are equal to rv cos β.
momentum because it equals the magnitude L of the angular momentum
divided by the mass m of the particle.

For brevity, the angular speed is sometimes written ω

However, it should not be assumed that ω is constant. Since h is constant, ω varies with the radius r according to
the formula[10]

Since h is constant and r2 is positive, the angle φ changes monotonically in any central-force problem, either
continuously increasing (h positive) or continuously decreasing (h negative).[11]

Constant areal velocity [ edit ]


See also: Areal velocity

The magnitude of h also equals twice the areal velocity, which is the rate at
which area is being swept out by the particle relative to the center.[12] Thus,
the areal velocity is constant for a particle acted upon by any type of central
force; this is Kepler's second law.[13] Conversely, if the motion under a
conservative force F is planar and has constant areal velocity for all initial
conditions of the radius r and velocity v, then the azimuthal acceleration a φ
is always zero. Hence, by Newton's second law, F = ma, the force is a
Since the area A equals
½ r⊥vt, the areal velocity dA/dt central force.
(the rate at which A is swept out
The constancy of areal velocity may be illustrated by uniform circular and
by the particle) equals
½ r⊥v = ½h. linear motion. In uniform circular motion, the particle moves with constant
speed v around the circumference of a circle of radius r. Since the angular
velocity ω = v/r is constant, the area swept out in a time Δt equals ω r2 Δt;
hence, equal areas are swept out in equal times Δt. In uniform linear motion (i.e., motion in the absence of a
force, by Newton's first law of motion), the particle moves with constant velocity, that is, with constant speed v
along a line. In a time Δt, the particle sweeps out an area ½vΔtr ⊥ (the impact parameter).[note 6] The distance r⊥
does not change as the particle moves along the line; it represents the distance of closest approach of the line to
the center O (the impact parameter). Since the speed v is likewise unchanging, the areal velocity ½vr ⊥ is a
constant of motion; the particle sweeps out equal areas in equal times.

Equivalent parallel force field [ edit ]


By a transformation of variables,[14] any central-force problem can be
converted into an equivalent parallel-force problem.[note 7] In place of the
ordinary x and y Cartesian coordinates, two new position variables ξ = x/y
and η = 1/y are defined, as is a new time coordinate τ

The corresponding equations of motion for ξ and η are given by


The area A of a circular
sector equals
½ r2φ = ½ r2ωt = ½ r vφt.
Hence, the areal velocity dA/dt
equals ½ r vφ = ½ h. For
uniform circular motion, r and vφ
Since the rate of change of ξ is constant, its second derivative is zero are constant; thus, dA/dt is also
constant.

Since this is the acceleration in the ξ direction and since F=ma by Newton's second law, it follows that the force
in the ξ direction is zero. Hence the force is only along the η direction, which is the criterion for a parallel-force
problem. Explicitly, the acceleration in the η direction equals

because the acceleration in the y-direction equals

Here, F y denotes the y-component of the central force, and y/r equals the cosine of the angle between the y-axis
and the radial vector r.

General solution [ edit ]

Binet equation [ edit ]


See also: Binet equation

Since a central force F acts only along the radius, only the radial component of the acceleration is nonzero. By
Newton's second law of motion, the magnitude of F equals the mass m of the particle times the magnitude of its
radial acceleration[15]

This equation has integration factor

Integrating yields

If h is not zero, the independent variable can be changed from t to [16]

giving the new equation of motion[17]

Making the change of variables to the inverse radius u = 1/r[17] yields

(1 )

where C is a constant of integration and the function G(u) is defined by

This equation becomes quasilinear on differentiating by

This is known as the Binet equation. Integrating (1) yields the solution for [18]

where 0 is another constant of integration. A central-force problem is said to be "integrable" if this final
integration can be solved in terms of known functions.

Orbit of the particle [ edit ]


The total energy of the system E tot equals the sum of the potential energy and the kinetic energy[19]

Since the total energy is constant, the rate of change of r can be calculated[20]

which may be converted (as before) to the derivative of r with respect to the azimuthal angle φ[17]

Integrating and using the angular-momentum formula L=mh yields the formula[21]

which indicates that the angular momentum contributes an effective potential energy[22]

Changing the variable of integration to the inverse radius yields the integral[23]

which expresses the above constants C = 2mE tot /L 2 and G(u) = 2mU(1/u)/L 2 above in terms of the total energy
E tot and the potential energy U(r).

Turning points and closed orbits [ edit ]


Main article: Bertrand's theorem

The rate of change of r is zero whenever the effective potential energy equals the total energy[24]

The points where this equation is satisfied are known as turning points.[24] The orbit on either side of a turning
point is symmetrical; in other words, if the azimuthal angle is defined such that φ = 0 at the turning point, then
the orbit is the same in opposite directions, r(φ) = r(−φ).[25]

If there are two turning points such that the radius r is bounded between rmin and rmax , then the motion is
contained within an annulus of those radii.[24] As the radius varies from the one turning point to the other, the
change in azimuthal angle φ equals[24]

The orbit will close upon itself[note 8] provided that Δφ equals a rational fraction of 2π, i.e.,[24]

where m and n are integers. In that case, the radius oscillates exactly m times while the azimuthal angle φ
makes exactly n revolutions. In general, however, Δφ/2π will not be such a rational number, and thus the orbit
will not be closed. In that case, the particle will eventually pass arbitrarily close to every point within the annulus.
Two types of central force always produce closed orbits: F(r) = αr (a linear force) and F(r) = α/r2 (an inverse-
square law). As shown by Bertrand, these two central forces are the only ones that guarantee closed orbits.[26]

In general, if the angular momentum L is nonzero, the L 2 /2mr 2 term prevents the particle from falling into the
origin, unless the effective potential energy goes to negative infinity in the limit of r going to zero.[27] Therefore, if
there is a single turning point, the orbit generally goes to infinity; the turning point corresponds to a point of
minimum radius.

Specific solutions [ edit ]

Kepler problem [ edit ]


See also: Kepler problem and Laplace–Runge–Lenz vector

In classical physics, many important forces follow an inverse-square law,


such as gravity or electrostatics . The general mathematical form of such
inverse-square central forces is

for a constant , which is negative for an attractive force and positive for a
repulsive one.

This special case of the classical central-force problem is called the Kepler
Classical gravity is a central
problem. For an inverse-square force, the Binet equation derived above is force. Solving that central-force
linear problem shows that a bound
particle follows an elliptical orbit
in which equal areas are swept
out in equal times, as described
by Kepler's second law.
The solution of this equation is

which shows that the orbit is a conic section of eccentricity e; here, φ0 is the initial angle, and the center of force
is at the focus of the conic section. Using the half-angle formula for sine, this solution can also be written as

where u 1 and u 2 are constants, with u 2 larger than u 1 . The two versions of
the solution are related by the equations

and

As for all central forces, the


particle in the Kepler problem Since the sin2 function is always greater than zero, u 2 is the largest possible
sweeps out equal areas in equal value of u and the inverse of the smallest possible value of r, i.e., the
times, as illustrated by the two
blue elliptical sectors. The distance of closest approach (periapsis ). Since the radial distance r cannot
center of force is located at one be a negative number, neither can its inverse u; therefore, u 2 must be a
of the foci of the elliptical orbit.
positive number. If u 1 is also positive, it is the smallest possible value of u,
which corresponds to the largest possible value of r, the distance of furthest
approach (apoapsis ). If u 1 is zero or negative, then the smallest possible value of u is zero (the orbit goes to
infinity); in this case, the only relevant values of φ are those that make u positive.

For an attractive force (α < 0), the orbit is an ellipse , a hyperbola or parabola , depending on whether u 1 is
positive, negative, or zero, respectively; this corresponds to an eccentricity e less than one, greater than one, or
equal to one. For a repulsive force (α > 0), u 1 must be negative, since u 2 is positive by definition and their sum is
negative; hence, the orbit is a hyperbola. Naturally, if no force is present (α=0), the orbit is a straight line.

Central forces with exact solutions [ edit ]


See also: Exact solutions of classical central-force problems

The Binet equation for u(φ) can be solved numerically for nearly any central force F(1/u). However, only a
handful of forces result in formulae for u in terms of known functions. As derived above, the solution for φ can be
expressed as an integral over u

A central-force problem is said to be "integrable" if this integration can be solved in terms of known functions.

If the force is a power law, i.e., if F(r) = α rn , then u can be expressed in terms of circular functions and/or elliptic
functions if n equals 1, -2, -3 (circular functions) and -7, -5, -4, 0, 3, 5, -3/2, -5/2, -1/3, -5/3 and -7/3 (elliptic
functions).[28] Similarly, only six possible linear combinations of power laws give solutions in terms of circular and
elliptic functions[29] [30]

The following special cases of the first two force types always result in circular functions.

The special case

was mentioned by Newton, in corollary 1 to proposition VII of the principia, as the force implied by circular orbits
passing through the point of attraction.

Revolving orbits [ edit ]


Main article: Newton's theorem of revolving orbits

The term r−3 occurs in all the force laws above, indicating that the addition of
the inverse-cube force does not influence the solubility of the problem in
terms of known functions. Newton showed that, with adjustments in the
initial conditions, the addition of such a force does not affect the radial motion
of the particle, but multiplies its angular motion by a constant factor k. An
extension of Newton's theorem was discovered in 2000 by Mahomed and
Vawda.[30]

Assume that a particle is moving under an arbitrary central force F 1 (r), and
Illustration of Newton's
let its radius r and azimuthal angle φ be denoted as r(t) and φ1 (t) as a theorem of revolving orbits. The
function of time t. Now consider a second particle with the same mass m that green planet completes one
shares the same radial motion r(t), but one whose angular speed is k times (subharmonic) orbit for every
three orbits of the blue planet
faster than that of the first particle. In other words, the azimuthal angles of (k=1/3). A GIF version of this
the two particles are related by the equation φ2 (t) = k φ1 (t). Newton showed animation is found here.
that the force acting on the second particle equals the force F 1 (r) acting on
the first particle, plus an inverse-cube central force[31]

where L 1 is the magnitude of the first particle's angular momentum.

If k 2 is greater than one, F 2 −F 1 is a negative number; thus, the added inverse-cube force is attractive .
Conversely, if k 2 is less than one, F 2 −F 1 is a positive number; the added inverse-cube force is repulsive . If k is
an integer such as 3, the orbit of the second particle is said to be a harmonic of the first particle's orbit; by
contrast, if k is the inverse of an integer, such as , the second orbit is said to be a subharmonic of the first
orbit.

Historical development [ edit ]

Newton's derivation [ edit ]


The classical central-force problem was solved geometrically by Isaac
Newton in his Philosophiæ Naturalis Principia Mathematica, in which Newton
introduced his laws of motion. Newton used an equivalent of leapfrog
integration to convert the continuous motion to a discrete one, so that
geometrical methods may be applied. In this approach, the position of the
particle is considered only at evenly spaced time points. For illustration, the
particle in Figure 10 is located at point A at time t = 0, at point B at time
t = Δt, at point C at time t = 2Δt, and so on for all times t = nΔt, where n is an

Figure 10: Newton's integer. The velocity is assumed to be constant between these time points.
geometrical proof that a moving Thus, the vector rAB = rB − rA equals Δt times the velocity vector v AB (red
particle sweeps out equal areas line), whereas rBC = rC − rB equals v BC Δt (blue line). Since the velocity is
in equal times if and only if the
force acting on it at the point B constant between points, the force is assumed to act instantaneously at each
is a central force. Here, the new position; for example, the force acting on the particle at point B instantly
triangle OAB has the same area
changes the velocity from v AB to v BC . The difference vector Δr = rBC − rAB
as the triangles OBC and OBK.
equals ΔvΔt (green line), where Δv = v BC − v AB is the change in velocity
resulting from the force at point B. Since the acceleration a is parallel to Δv
and since F = ma, the force F must be parallel to Δv and Δr. If F is a central force, it must be parallel to the
vector rB from the center O to the point B (dashed green line); in that case, Δr is also parallel to rB .

If no force acts at point B, the velocity is unchanged, and the particle arrives at point K at time t = 2Δt. The areas
of the triangles OAB and OBK are equal, because they share the same base (rAB ) and height (r⊥). If Δr is
parallel to rB , the triangles OBK and OBC are likewise equal, because they share the same base (rB ) and the
height is unchanged. In that case, the areas of the triangles OAB and OBC are the same, and the particle
sweeps out equal areas in equal time. Conversely, if the areas of all such triangles are equal, then Δr must be
parallel to rB , from which it follows that F is a central force. Thus, a particle sweeps out equal areas in equal
times if and only if F is a central force.

Alternative derivations of the equations of motion [ edit ]

Lagrangian mechanics [ edit ]


The formula for the radial force may also be obtained using Lagrangian mechanics. In polar cordinates, the
Lagrangian L of a single particle in a potential energy field U(r) is given by

Then Lagrange's equations of motion

take the form

since the magnitude F(r) of the radial force equals the negative derivative of the potential energy U(r) in the
radial direction.

Hamiltonian mechanics [ edit ]


The radial force formula may also be derived using Hamiltonian mechanics. In polar coordinates, the Hamiltonian
can be written as

Since the azimuthal angle φ does not appear in the Hamiltonian, its conjugate momentum p φ is a constant of the
motion. This conjugate momentum is the magnitude L of the angular momentum, as shown by the Hamiltonian
equation of motion for φ

The corresponding equation of motion for r is

You might also like