Equation of the center

Last updated
Simulated view of an object in an elliptic orbit, as seen from the focus of the orbit. The view rotates with the mean anomaly, so the object appears to oscillate back and forth across this mean position with the equation of the center. The object also appears to become smaller and larger as it moves farther away and nearer because of the eccentricity of the orbit. A marker (red) shows the position of the periapsis. Eoc anim.gif
Simulated view of an object in an elliptic orbit, as seen from the focus of the orbit. The view rotates with the mean anomaly, so the object appears to oscillate back and forth across this mean position with the equation of the center. The object also appears to become smaller and larger as it moves farther away and nearer because of the eccentricity of the orbit. A marker (red) shows the position of the periapsis.

In two-body, Keplerian orbital mechanics, the equation of the center is the angular difference between the actual position of a body in its elliptical orbit and the position it would occupy if its motion were uniform, in a circular orbit of the same period. It is defined as the difference true anomaly, ν, minus mean anomaly, M, and is typically expressed a function of mean anomaly, M, and orbital eccentricity, e. [1]

Contents

Discussion

Since antiquity, the problem of predicting the motions of the heavenly bodies has been simplified by reducing it to one of a single body in orbit about another. In calculating the position of the body around its orbit, it is often convenient to begin by assuming circular motion. This first approximation is then simply a constant angular rate multiplied by an amount of time. However, the actual solution, assuming Newtonian physics, is an elliptical orbit (a Keplerian orbit). For these, it is easy to find the mean anomaly (and hence the time) for a given true anomaly (the angular position of the planet around the sun), by converting true anomaly to "eccentric anomaly":

where atan2(y, x) is the angle from the x-axis of the ray from (0, 0) to (x, y), having the same sign as y (note that the arguments are often reversed in spreadsheets), and then using Kepler's equation to find the mean anomaly:

If is known and we wish to find and then Kepler's equation can be solved by numerical methods, but there are also series solutions involving sine of .

In cases of small eccentricity, the position given by a truncated series solution may be quite accurate. Many orbits of interest, such as those of bodies in the Solar System or of artificial Earth satellites, have these nearly-circular orbits. As eccentricity becomes greater, and orbits more elliptical, the accuracy of a given truncation of the series declines. If the series is taken as a power series in eccentricity then it fails to converge at high eccentricities.

The series in its modern form can be truncated at any point, and even when limited to just the most important terms it can produce an easily calculated approximation of the true position when full accuracy is not important. Such approximations can be used, for instance, as starting values for iterative solutions of Kepler's equation, [1] or in calculating rise or set times, which due to atmospheric effects cannot be predicted with much precision.

The ancient Greeks, in particular Hipparchus, knew the equation of the center as prosthaphaeresis , although their understanding of the geometry of the planets' motion was not the same. [2] The word equation (Latin, aequatio, -onis) in the present sense comes from astronomy. It was specified and used by Kepler, as that variable quantity determined by calculation which must be added or subtracted from the mean motion to obtain the true motion. In astronomy, the term equation of time has a similar meaning. [3] The equation of the center in modern form was developed as part of perturbation analysis, that is, the study of the effects of a third body on two-body motion. [4] [5]

Series expansion

Maximum error of the series expansion of the equation of the center, in radians, as a function of orbital eccentricity (bottom axis) and the power of e at which the series is truncated (right axis). Note that at low eccentricity (left-hand side of the graph), the series does not need to be carried to high order to produce accurate results. Eoc max err.png
Maximum error of the series expansion of the equation of the center, in radians, as a function of orbital eccentricity (bottom axis) and the power of e at which the series is truncated (right axis). Note that at low eccentricity (left-hand side of the graph), the series does not need to be carried to high order to produce accurate results.
Series-expanded equation of the center as a function of mean anomaly for various eccentricities, with the equation of the center truncated at e for all curves. Note that the truncated equation fails at high eccentricity and produces an oscillating curve. But this is because the coefficients of the Fourier series are inaccurate due to truncation in their calculation. Eoc vs e.png
Series-expanded equation of the center as a function of mean anomaly for various eccentricities, with the equation of the center truncated at e for all curves. Note that the truncated equation fails at high eccentricity and produces an oscillating curve. But this is because the coefficients of the Fourier series are inaccurate due to truncation in their calculation.

In Keplerian motion, the coordinates of the body retrace the same values with each orbit, which is the definition of a periodic function. Such functions can be expressed as periodic series of any continuously increasing angular variable, [6] and the variable of most interest is the mean anomaly, M. Because it increases uniformly with time, expressing any other variable as a series in mean anomaly is essentially the same as expressing it in terms of time. Although the true anomaly is an analytic function of M, it is not an entire function so a power series in M will have a limited range of convergence. But as a periodic function, a Fourier series will converge everywhere. The coefficients of the series are built from Bessel functions depending on the eccentricity e. Note that while these series can be presented in truncated form, they represent a sum of an infinite number of terms. [7]

The series for ν, the true anomaly can be expressed most conveniently in terms of M, e and Bessel functions of the first kind, [8]

where

are the Bessel functions and
[9]

The result is in radians.

The Bessel functions can be expanded in powers of x by, [10]

and βm by, [11]

Substituting and reducing, the equation for ν becomes (truncated at order e7), [8]

and by the definition, moving M to the left-hand side,

gives an approximation for the equation of the center. However, it is not a good approximation when e is high (see graph). If the coefficients are calculated from the Bessel functions then the approximation is much better when going up to the same frequency (such as ).

This formula is sometimes presented in terms of powers of e with coefficients in functions of sin M (here truncated at order e6),

which is similar to the above form. [12] [13] This presentation, when not truncated, contains the same infinite set of terms, but implies a different order of adding them up. Because of this, for small e, the series converges rapidly but if e exceeds the "Laplace limit" of 0.6627... then it diverges for all values of M (other than multiples of π), a fact discovered by Francesco Carlini and Pierre-Simon Laplace. [12] [14]

Examples

The equation of the center attains its maximum when the eccentric anomaly is the true anomaly is the mean anomaly is and the equation of the center is Here are some examples:

Orbital
eccentricity
[15]
True valueMaximum equation of the center (series truncated as shown)
e7e3e2
Venus 0.0067770.7766°0.7766°0.7766°0.7766°
Earth 0.016711.915°1.915°1.915°1.915°
Saturn 0.053866.173°6.174°6.174°6.186°
Mars 0.0933910.71°10.71°10.71°10.77°
Mercury 0.205623.64°23.68°23.77°23.28°

See also

Related Research Articles

<span class="mw-page-title-main">Bessel function</span> Families of solutions to related differential equations

Bessel functions, first defined by the mathematician Daniel Bernoulli and then generalized by Friedrich Bessel, are canonical solutions y(x) of Bessel's differential equation for an arbitrary complex number , which represents the order of the Bessel function. Although and produce the same differential equation, it is conventional to define different Bessel functions for these two values in such a way that the Bessel functions are mostly smooth functions of .

<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.

Orbital elements are the parameters required to uniquely identify a specific orbit. In celestial mechanics these elements are considered in two-body systems using a Kepler orbit. There are many different ways to mathematically describe the same orbit, but certain schemes, each consisting of a set of six parameters, are commonly used in astronomy and orbital mechanics.

<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">Equation of time</span> Apparent solar time minus mean solar time

The equation of time describes the discrepancy between two kinds of solar time. The word equation is used in the medieval sense of "reconciliation of a difference". The two times that differ are the apparent solar time, which directly tracks the diurnal motion of the Sun, and mean solar time, which tracks a theoretical mean Sun with uniform motion along the celestial equator. Apparent solar time can be obtained by measurement of the current position of the Sun, as indicated by a sundial. Mean solar time, for the same place, would be the time indicated by a steady clock set so that over the year its differences from apparent solar time would have a mean of zero.

<span class="mw-page-title-main">Mean anomaly</span> Specifies the orbit of an object in space

In celestial mechanics, the mean anomaly is the fraction of an elliptical orbit's period that has elapsed since the orbiting body passed periapsis, expressed as an angle which can be used in calculating the position of that body in the classical two-body problem. It is the angular distance from the pericenter which a fictitious body would have if it moved in a circular orbit, with constant speed, in the same orbital period as the actual body in its elliptical orbit.

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">True anomaly</span> Parameter of Keplerian orbits

In celestial mechanics, true anomaly is an angular parameter that defines the position of a body moving along a Keplerian orbit. It is the angle between the direction of periapsis and the current position of the body, as seen from the main focus of the ellipse.

<span class="mw-page-title-main">Parabolic trajectory</span> Type of orbit

In astrodynamics or celestial mechanics a parabolic trajectory is a Kepler orbit with the eccentricity equal to 1 and is an unbound orbit that is exactly on the border between elliptical and hyperbolic. When moving away from the source it is called an escape orbit, otherwise a capture orbit. It is also sometimes referred to as a C3 = 0 orbit (see Characteristic energy).

<span class="mw-page-title-main">Hyperbolic trajectory</span> Concept in astrodynamics

In astrodynamics or celestial mechanics, a hyperbolic trajectory or hyperbolic orbit is the trajectory of any object around a central body with more than enough speed to escape the central object's gravitational pull. The name derives from the fact that according to Newtonian theory such an orbit has the shape of a hyperbola. In more technical terms this can be expressed by the condition that the orbital eccentricity is greater than one.

<span class="mw-page-title-main">Elliptic orbit</span> Kepler orbit with an eccentricity of less than one

In astrodynamics or celestial mechanics, an elliptic orbit or elliptical orbit is a Kepler orbit with an eccentricity of less than 1; this includes the special case of a circular orbit, with eccentricity equal to 0. In a stricter sense, it is a Kepler orbit with the eccentricity greater than 0 and less than 1. In a wider sense, it is a Kepler orbit with negative energy. This includes the radial elliptic orbit, with eccentricity equal to 1.

<span class="mw-page-title-main">Spacecraft flight dynamics</span> Application of mechanical dynamics to model the flight of space vehicles

Spacecraft flight dynamics is the application of mechanical dynamics to model how the external forces acting on a space vehicle or spacecraft determine its flight path. These forces are primarily of three types: propulsive force provided by the vehicle's engines; gravitational force exerted by the Earth and other celestial bodies; and aerodynamic lift and drag.

The two-body problem in general relativity is the determination of the motion and gravitational field of two bodies as described by the field equations of general relativity. Solving the Kepler problem is essential to calculate the bending of light by gravity and the motion of a planet orbiting its sun. Solutions are also used to describe the motion of binary stars around each other, and estimate their gradual loss of energy through gravitational radiation.

In mathematics, the Laplace limit is the maximum value of the eccentricity for which a solution to Kepler's equation, in terms of a power series in the eccentricity, converges. It is approximately

<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.

In orbital mechanics, the universal variable formulation is a method used to solve the two-body Kepler problem. It is a generalized form of Kepler's Equation, extending it to apply not only to elliptic orbits, but also parabolic and hyperbolic orbits common for spacecraft departing from a planetary orbit. It is also applicable to ejection of small bodies in Solar System from the vicinity of massive planets, during which processes the approximating two-body orbits can have widely varying eccentricities, almost always e ≥ 1 .

In celestial mechanics, Lambert's problem is concerned with the determination of an orbit from two position vectors and the time of flight, posed in the 18th century by Johann Heinrich Lambert and formally solved with mathematical proof by Joseph-Louis Lagrange. It has important applications in the areas of rendezvous, targeting, guidance, and preliminary orbit determination.

In astrodynamics and celestial mechanics a radial trajectory is a Kepler orbit with zero angular momentum. Two objects in a radial trajectory move directly towards or away from each other in a straight line.

<span class="mw-page-title-main">Kepler's equation</span> Orbital mechanics term

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

<span class="mw-page-title-main">Position of the Sun</span> Calculating the Suns location in the sky at a given time and place

The position of the Sun in the sky is a function of both the time and the geographic location of observation on Earth's surface. As Earth orbits the Sun over the course of a year, the Sun appears to move with respect to the fixed stars on the celestial sphere, along a circular path called the ecliptic.

References

  1. 1 2 Vallado, David A. (2001). Fundamentals of Astrodynamics and Applications (second ed.). Microcosm Press, El Segundo, CA. p. 82. ISBN   1-881883-12-4.
  2. Narrien, John (1833). An Historical Account of the Origin and Progress of Astronomy. Baldwin and Cradock, London. pp.  230–231.
  3. Capderou, Michel (2005). Satellites Orbits and Missions . Springer-Verlag. p.  23. ISBN   978-2-287-21317-5.
  4. Moulton, Forest Ray (1914). An Introduction to Celestial Mechanics (second revised ed.). Macmillan Co., New York. p. 165. ISBN   9780598943972., at Google books
  5. Smart, W. M. (1953). Celestial Mechanics. Longmans, Green and Co., London. p. 26.
  6. Brouwer, Dirk; Clemence, Gerald M. (1961). Methods of Celestial Mechanics . Academic Press, New York and London. p.  60.
  7. Vallado, David A. (2001). p. 80
  8. 1 2 Brouwer, Dirk; Clemence, Gerald M. (1961). p. 77.
  9. Brouwer, Dirk; Clemence, Gerald M. (1961). p. 62.
  10. Brouwer, Dirk; Clemence, Gerald M. (1961). p. 68.
  11. Smart, W. M. (1953). p. 32.
  12. 1 2 Moulton, Forest Ray (1914). pp. 171–172.
  13. Danby, J.M.A. (1988). Fundamentals of Celestial Mechanics. Willmann-Bell, Inc., Richmond, VA. pp. 199–200. ISBN   0-943396-20-4.
  14. Plummer, H. C. (1918). An Introductory Treatise on Dynamical Astronomy (PDF). Cambridge University Press. pp.  46–47.
  15. Seidelmann, P. Kenneth; Urban, Sean E., eds. (2013). Explanatory Supplement to the Astronomical Almanac (3rd ed.). University Science Books, Mill Valley, CA. p. 338. ISBN   978-1-891389-85-6.

Further reading