Kepler's equation

Last updated
Kepler's equation solutions for five different eccentricities between 0 and 1 Kepler equation solutions.svg
Kepler's equation solutions for five different eccentricities between 0 and 1

In orbital mechanics, Kepler's equation relates various geometric properties of the orbit of a body subject to a central force.

Contents

It was derived by Johannes Kepler in 1609 in Chapter 60 of his Astronomia nova , [1] [2] and in book V of his Epitome of Copernican Astronomy (1621) Kepler proposed an iterative solution to the equation. [3] [4] This equation and its solution, however, first appeared in a 9th-century work by Habash al-Hasib al-Marwazi, which dealt with problems of parallax. [5] [6] [7] [8] The equation has played an important role in the history of both physics and mathematics, particularly classical celestial mechanics.

Equation

Kepler's equation is

where is the mean anomaly, is the eccentric anomaly, and is the eccentricity.

The 'eccentric anomaly' is useful to compute the position of a point moving in a Keplerian orbit. As for instance, if the body passes the periastron at coordinates , , at time , then to find out the position of the body at any time, you first calculate the mean anomaly from the time and the mean motion by the formula , then solve the Kepler equation above to get , then get the coordinates from:

where is the semi-major axis, the semi-minor axis.

Kepler's equation is a transcendental equation because sine is a transcendental function, and it cannot be solved for algebraically. Numerical analysis and series expansions are generally required to evaluate .

Alternate forms

There are several forms of Kepler's equation. Each form is associated with a specific type of orbit. The standard Kepler equation is used for elliptic orbits (). The hyperbolic Kepler equation is used for hyperbolic trajectories (). The radial Kepler equation is used for linear (radial) trajectories (). Barker's equation is used for parabolic trajectories ().

When , the orbit is circular. Increasing causes the circle to become elliptical. When , there are four possibilities:

A value of slightly above 1 results in a hyperbolic orbit with a turning angle of just under 180 degrees. Further increases reduce the turning angle, and as goes to infinity, the orbit becomes a straight line of infinite length.

Hyperbolic Kepler equation

The Hyperbolic Kepler equation is:

where is the hyperbolic eccentric anomaly. This equation is derived by redefining M to be the square root of −1 times the right-hand side of the elliptical equation:

(in which is now imaginary) and then replacing by .

Radial Kepler equations

The Radial Kepler equation for the case where the object does not have enough energy to escape is:

where is proportional to time and is proportional to the distance from the centre of attraction along the ray and attains the value 1 at the maximum distance. This equation is derived by multiplying Kepler's equation by 1/2 and setting to 1:

and then making the substitution

The radial equation for when the object has enough energy to escape is:

When the energy is exactly the minimum amount needed to escape, then the time is simply proportional to the distance to the power 3/2.

Inverse problem

Calculating for a given value of is straightforward. However, solving for when is given can be considerably more challenging. There is no closed-form solution. Solving for is more or less equivalent to solving for the true anomaly, or the difference between the true anomaly and the mean anomaly, which is called the "Equation of the center".

One can write an infinite series expression for the solution to Kepler's equation using Lagrange inversion, but the series does not converge for all combinations of and (see below).

Confusion over the solvability of Kepler's equation has persisted in the literature for four centuries. [9] Kepler himself expressed doubt at the possibility of finding a general solution:

I am sufficiently satisfied that it [Kepler's equation] cannot be solved a priori, on account of the different nature of the arc and the sine. But if I am mistaken, and any one shall point out the way to me, he will be in my eyes the great Apollonius.

Johannes Kepler [10]

Fourier series expansion (with respect to ) using Bessel functions is [11] [12] [13]

With respect to , it is a Kapteyn series.

Inverse Kepler equation

The inverse Kepler equation is the solution of Kepler's equation for all real values of :

Evaluating this yields:

These series can be reproduced in Mathematica with the InverseSeries operation.

InverseSeries[Series[M-Sin[M],{M,0,10}]]
InverseSeries[Series[M-eSin[M],{M,0,10}]]

These functions are simple Maclaurin series. Such Taylor series representations of transcendental functions are considered to be definitions of those functions. Therefore, this solution is a formal definition of the inverse Kepler equation. However, is not an entire function of at a given non-zero . Indeed, the derivative

goes to zero at an infinite set of complex numbers when the nearest to zero being at and at these two points

(where inverse cosh is taken to be positive), and goes to infinity at these values of . This means that the radius of convergence of the Maclaurin series is and the series will not converge for values of larger than this. The series can also be used for the hyperbolic case, in which case the radius of convergence is The series for when converges when .

While this solution is the simplest in a certain mathematical sense,[ which? ], other solutions are preferable for most applications. Alternatively, Kepler's equation can be solved numerically.

The solution for was found by Karl Stumpff in 1968, [14] but its significance wasn't recognized. [15] [ clarification needed ]

One can also write a Maclaurin series in . This series does not converge when is larger than the Laplace limit (about 0.66), regardless of the value of (unless is a multiple of ), but it converges for all if is less than the Laplace limit. The coefficients in the series, other than the first (which is simply ), depend on in a periodic way with period .

Inverse radial Kepler equation

The inverse radial Kepler equation () for the case in which the object does not have enough energy to escape can similarly be written as:

Evaluating this yields:

To obtain this result using Mathematica:

InverseSeries[Series[ArcSin[Sqrt[t]]-Sqrt[(1-t)t],{t,0,15}]]

Numerical approximation of inverse problem

Newton's method

For most applications, the inverse problem can be computed numerically by finding the root of the function:

This can be done iteratively via Newton's method:

Note that and are in units of radians in this computation. This iteration is repeated until desired accuracy is obtained (e.g. when < desired accuracy). For most elliptical orbits an initial value of is sufficient. For orbits with , a initial value of can be used. Numerous works developed accurate (but also more complex) start guesses. [16] If is identically 1, then the derivative of , which is in the denominator of Newton's method, can get close to zero, making derivative-based methods such as Newton-Raphson, secant, or regula falsi numerically unstable. In that case, the bisection method will provide guaranteed convergence, particularly since the solution can be bounded in a small initial interval. On modern computers, it is possible to achieve 4 or 5 digits of accuracy in 17 to 18 iterations. [17] A similar approach can be used for the hyperbolic form of Kepler's equation. [18] :66–67 In the case of a parabolic trajectory, Barker's equation is used.

Fixed-point iteration

A related method starts by noting that . Repeatedly substituting the expression on the right for the on the right yields a simple fixed-point iteration algorithm for evaluating . This method is identical to Kepler's 1621 solution. [4]

functionE(e,M,n)E=Mfork=1tonE=M+e*sinEnextkreturnE

The number of iterations, , depends on the value of . The hyperbolic form similarly has .

This method is related to the Newton's method solution above in that

To first order in the small quantities and ,

.

See also

Related Research Articles

<span class="mw-page-title-main">Euler's formula</span> Complex exponential in terms of sine and cosine

Euler's formula, named after Leonhard Euler, is a mathematical formula in complex analysis that establishes the fundamental relationship between the trigonometric functions and the complex exponential function. Euler's formula states that, for any real number x, one has where e is the base of the natural logarithm, i is the imaginary unit, and cos and sin are the trigonometric functions cosine and sine respectively. This complex exponential function is sometimes denoted cis x. The formula is still valid if x is a complex number, and is also called Euler's formula in this more general case.

In classical mechanics, a harmonic oscillator is a system that, when displaced from its equilibrium position, experiences a restoring force F proportional to the displacement x: where k is a positive constant.

<span class="mw-page-title-main">Kepler's laws of planetary motion</span> Laws describing the motion of planets

In astronomy, Kepler's laws of planetary motion, published by Johannes Kepler between 1609 and 1619, describe the orbits of planets around the Sun. The laws modified the heliocentric theory of Nicolaus Copernicus, replacing its circular orbits and epicycles with elliptical trajectories, and explaining how planetary velocities vary. The three laws state that:

  1. The orbit of a planet is an ellipse with the Sun at one of the two foci.
  2. A line segment joining a planet and the Sun sweeps out equal areas during equal intervals of time.
  3. The square of a planet's orbital period is proportional to the cube of the length of the semi-major axis of its orbit.
<span class="mw-page-title-main">Simple harmonic motion</span> To-and-fro periodic motion in science and engineering

In mechanics and physics, simple harmonic motion is a special type of periodic motion an object experiences by means of a restoring force whose magnitude is directly proportional to the distance of the object from an equilibrium position and acts towards the equilibrium position. It results in an oscillation that is described by a sinusoid which continues indefinitely.

<span class="mw-page-title-main">Spherical coordinate system</span> Coordinates comprising a distance and two angles

In mathematics, a spherical coordinate system is a coordinate system for three-dimensional space where the position of a given point in space is specified by three real numbers: the radial distancer along the radial line connecting the point to the fixed point of origin; the polar angleθ between the radial line and a polar axis; and the azimuthal angleφ as the angle of rotation of the radial line around the polar axis. (See graphic re the "physics convention".) Once the radius is fixed, the three coordinates (r, θ, φ), known as a 3-tuple, provide a coordinate system on a sphere, typically called the spherical polar coordinates.

<span class="mw-page-title-main">Trigonometric functions</span> Functions of an angle

In mathematics, the trigonometric functions are real functions which relate an angle of a right-angled triangle to ratios of two side lengths. They are widely used in all sciences that are related to geometry, such as navigation, solid mechanics, celestial mechanics, geodesy, and many others. They are among the simplest periodic functions, and as such are also widely used for studying periodic phenomena through Fourier analysis.

<span class="mw-page-title-main">Orbital mechanics</span> Field of classical mechanics concerned with the motion of spacecraft

Orbital mechanics or astrodynamics is the application of ballistics and celestial mechanics to the practical problems concerning the motion of rockets, satellites, and other spacecraft. The motion of these objects is usually calculated from Newton's laws of motion and the law of universal gravitation. Orbital mechanics is a core discipline within space-mission design and control.

<span class="mw-page-title-main">Chebyshev polynomials</span> Polynomial sequence

The Chebyshev polynomials are two sequences of polynomials related to the cosine and sine functions, notated as and . They can be defined in several equivalent ways, one of which starts with trigonometric functions:

<span class="mw-page-title-main">Spherical harmonics</span> Special mathematical functions defined on the surface of a sphere

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields. The table of spherical harmonics contains a list of common spherical harmonics.

<span class="mw-page-title-main">Projectile motion</span> Motion of launched objects due to gravity

Projectile motion is a form of motion experienced by an object or particle that is projected in a gravitational field, such as from Earth's surface, and moves along a curved path under the action of gravity only. In the particular case of projectile motion on Earth, most calculations assume the effects of air resistance are passive and negligible.

In orbital mechanics, the eccentric anomaly is an angular parameter that defines the position of a body that is moving along an elliptic Kepler orbit. The eccentric anomaly is one of three angular parameters ("anomalies") that define a position along an orbit, the other two being the true anomaly and the mean anomaly.

<span class="mw-page-title-main">Euler–Bernoulli beam theory</span> Method for load calculation in construction

Euler–Bernoulli beam theory is a simplification of the linear theory of elasticity which provides a means of calculating the load-carrying and deflection characteristics of beams. It covers the case corresponding to small deflections of a beam that is subjected to lateral loads only. By ignoring the effects of shear deformation and rotatory inertia, it is thus a special case of Timoshenko–Ehrenfest beam theory. It was first enunciated circa 1750, but was not applied on a large scale until the development of the Eiffel Tower and the Ferris wheel in the late 19th century. Following these successful demonstrations, it quickly became a cornerstone of engineering and an enabler of the Second Industrial Revolution.

In mathematics, a change of variables is a basic technique used to simplify problems in which the original variables are replaced with functions of other variables. The intent is that when expressed in new variables, the problem may become simpler, or equivalent to a better understood problem.

A banked turn is a turn or change of direction in which the vehicle banks or inclines, usually towards the inside of the turn. For a road or railroad this is usually due to the roadbed having a transverse down-slope towards the inside of the curve. The bank angle is the angle at which the vehicle is inclined about its longitudinal axis with respect to the horizontal.

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support such that it freely swings back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

In classical mechanics, holonomic constraints are relations between the position variables that can be expressed in the following form:

<span class="mw-page-title-main">Elastic instability</span>

Elastic instability is a form of instability occurring in elastic systems, such as buckling of beams and plates subject to large compressive loads.

<span class="mw-page-title-main">Kepler orbit</span> Celestial orbit whose trajectory is a conic section in the orbital plane

In celestial mechanics, a Kepler orbit is the motion of one body relative to another, as an ellipse, parabola, or hyperbola, which forms a two-dimensional orbital plane in three-dimensional space. A Kepler orbit can also form a straight line. It considers only the point-like gravitational attraction of two bodies, neglecting perturbations due to gravitational interactions with other objects, atmospheric drag, solar radiation pressure, a non-spherical central body, and so on. It is thus said to be a solution of a special case of the two-body problem, known as the Kepler problem. As a theory in classical mechanics, it also does not take into account the effects of general relativity. Keplerian orbits can be parametrized into six orbital elements in various ways.

The Binet equation, derived by Jacques Philippe Marie Binet, provides the form of a central force given the shape of the orbital motion in plane polar coordinates. The equation can also be used to derive the shape of the orbit for a given force law, but this usually involves the solution to a second order nonlinear, ordinary differential equation. A unique solution is impossible in the case of circular motion about the center of force.

In Euclidean geometry, for a plane curve C and a given fixed point O, the pedal equation of the curve is a relation between r and p where r is the distance from O to a point on C and p is the perpendicular distance from O to the tangent line to C at the point. The point O is called the pedal point and the values r and p are sometimes called the pedal coordinates of a point relative to the curve and the pedal point. It is also useful to measure the distance of O to the normal pc (the contrapedal coordinate) even though it is not an independent quantity and it relates to (r, p) as

References

  1. Kepler, Johannes (1609). "LX. Methodus, ex hac Physica, hoc est genuina & verissima hypothesi, extruendi utramque partem æquationis, & distantias genuinas: quorum utrumque simul per vicariam fieri hactenus non potuit. argumentum falsæ hypotheseos". Astronomia Nova Aitiologētos, Seu Physica Coelestis, tradita commentariis De Motibus Stellæ Martis, Ex observationibus G. V. Tychonis Brahe (in Latin). pp. 299–300.
  2. Aaboe, Asger (2001). Episodes from the Early History of Astronomy. Springer. pp. 146–147. ISBN   978-0-387-95136-2.
  3. Kepler, Johannes (1621). "Libri V. Pars altera.". Epitome astronomiæ Copernicanæ usitatâ formâ Quæstionum & Responsionum conscripta, inq; VII. Libros digesta, quorum tres hi priores sunt de Doctrina Sphæricâ (in Latin). pp. 695–696.
  4. 1 2 Swerdlow, Noel M. (2000). "Kepler's Iterative Solution to Kepler's Equation". Journal for the History of Astronomy . 31 (4): 339–341. Bibcode:2000JHA....31..339S. doi:10.1177/002182860003100404. S2CID   116599258.
  5. Colwell, Peter (1993). Solving Kepler's Equation Over Three Centuries. Willmann-Bell. p. 4. ISBN   978-0-943396-40-8.
  6. Dutka, J. (1997-07-01). "A note on "Kepler's equation"". Archive for History of Exact Sciences. 51 (1): 59–65. Bibcode:1997AHES...51...59D. doi:10.1007/BF00376451. S2CID   122568981.
  7. North, John (2008-07-15). Cosmos: An Illustrated History of Astronomy and Cosmology. University of Chicago Press. ISBN   978-0-226-59441-5.
  8. Livingston, John W. (2017-12-14). The Rise of Science in Islam and the West: From Shared Heritage to Parting of The Ways, 8th to 19th Centuries. Routledge. ISBN   978-1-351-58926-0.
  9. It is often claimed that Kepler's equation "cannot be solved analytically"; see for example here. Other authors claim that it cannot be solved at all; see for example Madabushi V. K. Chari; Sheppard Joel Salon; Numerical Methods in Electromagnetism, Academic Press, San Diego, CA, USA, 2000, ISBN   0-12-615760-X, p. 659
  10. "Mihi ſufficit credere, ſolvi a priori non poſſe, propter arcus & ſinus ετερογενειαν. Erranti mihi, quicumque viam monſtraverit, is erit mihi magnus Apollonius." Hall, Asaph (May 1883). "Kepler's Problem". Annals of Mathematics. 10 (3): 65–66. doi:10.2307/2635832. JSTOR   2635832.
  11. Fitzpatrick, Philip Matthew (1970). Principles of celestial mechanics. Academic Press. ISBN   0-12-257950-X.
  12. Colwell, Peter (January 1992). "Bessel Functions and Kepler's Equation". The American Mathematical Monthly. 99 (1): 45–48. doi:10.2307/2324547. ISSN   0002-9890. JSTOR   2324547.
  13. Boyd, John P. (2007). "Rootfinding for a transcendental equation without a first guess: Polynomialization of Kepler's equation through Chebyshev polynomial equation of the sine". Applied Numerical Mathematics. 57 (1): 12–18. doi:10.1016/j.apnum.2005.11.010.
  14. Stumpff, Karl (1 June 1968). "On The application of Lie-series to the problems of celestial mechanics". NASA Technical Note D-4460.{{cite journal}}: Cite journal requires |journal= (help)
  15. Colwell, Peter (1993). Solving Kepler's Equation Over Three Centuries. Willmann–Bell. p. 43. ISBN   0-943396-40-9.
  16. Odell, A. W.; Gooding, R. H. (1986). "Procedures for solving Kepler's equation". Celestial Mechanics. 38 (4). Springer Science and Business Media LLC: 307–334. Bibcode:1986CeMec..38..307O. doi:10.1007/bf01238923. ISSN   1572-9478. S2CID   120179781.
  17. Keister, Adrian. "The Numerical Analysis of Finding the Height of a Circular Segment". Wineman Technology. Wineman Technology, Inc. Retrieved 28 December 2019.
  18. Pfleger, Thomas; Montenbruck, Oliver (1998). Astronomy on the Personal Computer (Third ed.). Berlin, Heidelberg: Springer. ISBN   978-3-662-03349-4.