Kepler's equation Contents Equation Alternate forms Inverse problem Numerical approximation of inverse...

Kepler conjectureKepler's laws of planetary motionKepler orbitKepler triangleKepler's equationKepler polyhedraKepler's SupernovaKeplerian telescopeKatharina KeplerJakob BartschKeplerKepler space telescopeJohannes Kepler ATVList of things


Johannes KeplerOrbits


orbital mechanicscentral forceJohannes Keplercelestial mechanicsmean anomalyeccentric anomalyeccentricitymean motionsemi-major axissemi-minor axistranscendental equationsinetranscendental functionalgebraicallyNumerical analysisseriesBarker's equationsquare root of −1closed-form solutioninfinite seriesLagrange inversionMathematicaMaclaurin seriesentire functionKarl StumpffLaplace limitMathematicarootNewton's methodBarker's equationfixed-point iterationNewton's method







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.


It was first 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] The equation has played an important role in the history of both physics and mathematics, particularly classical celestial mechanics.




Contents






  • 1 Equation


  • 2 Alternate forms


    • 2.1 Hyperbolic Kepler equation


    • 2.2 Radial Kepler equation




  • 3 Inverse problem


    • 3.1 Inverse Kepler equation


    • 3.2 Inverse radial Kepler equation




  • 4 Numerical approximation of inverse problem


    • 4.1 Fixed-point iteration




  • 5 See also


  • 6 References


  • 7 External links





Equation


Kepler's equation is



M=E−esin⁡E{displaystyle M=E-esin E}



where M is the mean anomaly, E is the eccentric anomaly, and e is the eccentricity.


The 'eccentric anomaly' E 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 x = a(1 − e), y = 0, at time t = t0, then to find out the position of the body at any time, you first calculate the mean anomaly M from the time and the mean motion n by the formula M = n(tt0), then solve the Kepler equation above to get E, then get the coordinates from:



x=a(cos⁡E−e)y=bsin⁡E{displaystyle {begin{array}{lcl}x&=&a(cos E-e)\y&=&bsin Eend{array}}}



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


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



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 (0 ≤ e < 1). The hyperbolic Kepler equation is used for hyperbolic trajectories (e ≫ 1). The radial Kepler equation is used for linear (radial) trajectories (e = 1). Barker's equation is used for parabolic trajectories (e = 1).


When e = 0, the orbit is circular. Increasing e causes the circle to become elliptical. When e = 1, there are three possibilities:



  • a parabolic trajectory,

  • a trajectory going in or out along an infinite ray emanating from the centre of attraction,

  • or a trajectory that goes back and forth along a line segment from the centre of attraction to a point at some distance away.


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



Hyperbolic Kepler equation


The Hyperbolic Kepler equation is:



M=esinh⁡H−H{displaystyle M=esinh H-H}



where H 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:


M=i(E−esin⁡E){displaystyle M=ileft(E-esin Eright)}

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



Radial Kepler equation


The Radial Kepler equation is:



t(x)=sin−1⁡(x)−x(1−x){displaystyle t(x)=sin ^{-1}({sqrt {x}})-{sqrt {x(1-x)}}}



where t is proportional to time and x is proportional to the distance from the centre of attraction along the ray. This equation is derived by multiplying Kepler's equation by 1/2 and setting e to 1:


t(x)=12[E−sin⁡(E)].{displaystyle t(x)={frac {1}{2}}left[E-sin(E)right].}

and then making the substitution


E=2sin−1⁡(x).{displaystyle E=2sin ^{-1}({sqrt {x}}).}


Inverse problem


Calculating M for a given value of E is straightforward. However, solving for E when M is given can be considerably more challenging. There is no closed-form solution.


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 e and M (see below).


Confusion over the solvability of Kepler's equation has persisted in the literature for four centuries.[5] Kepler himself expressed doubt at the possibility of finding a general solution.











Inverse Kepler equation


The inverse Kepler equation is the solution of Kepler's equation for all real values of e{displaystyle e}:


E={∑n=1∞Mn3n!limθ0+(dn−1dθn−1((θθsin⁡)3)n)),e=1∑n=1∞Mnn!limθ0+(dn−1dθn−1((θθesin⁡))n)),e≠1{displaystyle E={begin{cases}displaystyle sum _{n=1}^{infty }{frac {M^{frac {n}{3}}}{n!}}lim _{theta to 0^{+}}!{Bigg (}{frac {mathrm {d} ^{,n-1}}{mathrm {d} theta ^{,n-1}}}{bigg (}{bigg (}{frac {theta }{sqrt[{3}]{theta -sin(theta )}}}{bigg )}^{!!!n}{bigg )}{Bigg )},&e=1\displaystyle sum _{n=1}^{infty }{frac {M^{n}}{n!}}lim _{theta to 0^{+}}!{Bigg (}{frac {mathrm {d} ^{,n-1}}{mathrm {d} theta ^{,n-1}}}{bigg (}{Big (}{frac {theta }{theta -esin(theta )}}{Big )}^{!n}{bigg )}{Bigg )},&eneq 1end{cases}}}

Evaluating this yields:


E={s+160s3+11400s5+125200s7+4317248000s9+12137207200000s11+15143912713500800000s13+⋯ with s=(6M)1/3,e=111−eM−e(1−e)4M33!+(9e2+e)(1−e)7M55!−(225e3+54e2+e)(1−e)10M77!+(11025e4+4131e3+243e2+e)(1−e)13M99!+⋯,e≠1{displaystyle E={begin{cases}displaystyle s+{frac {1}{60}}s^{3}+{frac {1}{1400}}s^{5}+{frac {1}{25200}}s^{7}+{frac {43}{17248000}}s^{9}+{frac {1213}{7207200000}}s^{11}+{frac {151439}{12713500800000}}s^{13}+cdots {text{ with }}s=(6M)^{1/3},&e=1\\displaystyle {frac {1}{1-e}}M-{frac {e}{(1-e)^{4}}}{frac {M^{3}}{3!}}+{frac {(9e^{2}+e)}{(1-e)^{7}}}{frac {M^{5}}{5!}}-{frac {(225e^{3}+54e^{2}+e)}{(1-e)^{10}}}{frac {M^{7}}{7!}}+{frac {(11025e^{4}+4131e^{3}+243e^{2}+e)}{(1-e)^{13}}}{frac {M^{9}}{9!}}+cdots ,&eneq 1end{cases}}}

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



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

InverseSeries[Series[M - e Sin[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, E is not an entire function of M at a given non-zero e. The derivative


dM/dE=1−ecos⁡E{displaystyle dM/dE=1-ecos E}

goes to zero at an infinite set of complex numbers when e<1. There are solutions at E=±icosh−1⁡(1/e),{displaystyle E=pm icosh ^{-1}(1/e),} and at those values


M=E−esin⁡E=±i(cosh−1⁡(1/e)−1−e2){displaystyle M=E-esin E=pm ileft(cosh ^{-1}(1/e)-{sqrt {1-e^{2}}}right)}

(where inverse cosh is taken to be positive), and dE/dM goes to infinity at these points. This means that the radius of convergence of the Maclaurin series is cosh−1⁡(1/e)−1−e2{displaystyle cosh ^{-1}(1/e)-{sqrt {1-e^{2}}}} and the series will not converge for values of M larger than this. The series can also be used for the hyperbolic case, in which case the radius of convergence is cos−1⁡(1/e)−e2−1.{displaystyle cos ^{-1}(1/e)-{sqrt {e^{2}-1}}.} The series for when e = 1 converges when m < 2π.


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 e ≠ 1 was found by Karl Stumpff in 1968,[7] but its significance wasn't recognized.[8][clarification needed]


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



Inverse radial Kepler equation


The inverse radial Kepler equation (e = 1) can also be written as:


x(t)=∑n=1∞[limr→0+(t23nn!dn−1drn−1(rn(32(sin−1⁡(r)−r−r2))−23n))]{displaystyle x(t)=sum _{n=1}^{infty }left[lim _{rto 0^{+}}left({frac {t^{{frac {2}{3}}n}}{n!}}{frac {mathrm {d} ^{,n-1}}{mathrm {d} r^{,n-1}}}!left(r^{n}left({frac {3}{2}}{Big (}sin ^{-1}({sqrt {r}})-{sqrt {r-r^{2}}}{Big )}right)^{!-{frac {2}{3}}n}right)right)right]}

Evaluating this yields:


x(t)=p−15p2−3175p3−237875p4−18943931875p5−329321896875p6−241809262077640625p7− ⋯ |p=(32t)2/3{displaystyle x(t)=p-{frac {1}{5}}p^{2}-{frac {3}{175}}p^{3}-{frac {23}{7875}}p^{4}-{frac {1894}{3931875}}p^{5}-{frac {3293}{21896875}}p^{6}-{frac {2418092}{62077640625}}p^{7}- cdots {bigg |}{p=left({tfrac {3}{2}}tright)^{2/3}}}

To obtain this result using Mathematica:


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


Numerical approximation of inverse problem


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


f(E)=E−esin⁡(E)−M(t){displaystyle f(E)=E-esin(E)-M(t)}

This can be done iteratively via Newton's method:


En+1=En−f(En)f′(En)=En−En−esin⁡(En)−M(t)1−ecos⁡(En){displaystyle E_{n+1}=E_{n}-{frac {f(E_{n})}{f'(E_{n})}}=E_{n}-{frac {E_{n}-esin(E_{n})-M(t)}{1-ecos(E_{n})}}}

Note that E and M are in units of radians in this computation. This iteration is repeated until desired accuracy is obtained (e.g. when f(E) < desired accuracy). For most elliptical orbits an initial value of E0 = M(t) is sufficient. For orbits with e > 0.8, an initial value of E0 = π should be used. A similar approach can be used for the hyperbolic form of Kepler's equation.[9]:66–67 In the case of a parabolic trajectory, Barker's equation is used.



Fixed-point iteration


A related method starts by noting that E=M+esin⁡E{displaystyle E=M+esin {E}}. Repeatedly substituting the expression on the right for the E{displaystyle E} on the right yields a simple fixed-point iteration algorithm for evaluating E(e,M){displaystyle E(e,M)}. This method is identical to Kepler's 1621 solution.[4]


function E(e,M,n)
E = M
for k = 1 to n
E = M + e*sin E
next k
return E

The number of iterations, n{displaystyle n}, depends on the value of e{displaystyle e}. The hyperbolic form similarly has H=esinh⁡H−M{displaystyle H=esinh H-M}.


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


En+1=En−En−esin⁡(En)−M(t)1−ecos⁡(En)=En+(M+esin⁡En−En)(1+ecos⁡En)1−e2(cos⁡En)2{displaystyle E_{n+1}=E_{n}-{frac {E_{n}-esin(E_{n})-M(t)}{1-ecos(E_{n})}}=E_{n}+{frac {(M+esin {E_{n}}-E_{n})(1+ecos {E_{n}})}{1-e^{2}(cos {E_{n}})^{2}}}}

To first order in the small quantities M−En{displaystyle M-E_{n}} and e{displaystyle e},



En+1≈M+esin⁡En{displaystyle E_{n+1}approx M+esin {E_{n}}}.


See also



  • Kepler's laws of planetary motion

  • Kepler problem

  • Kepler problem in general relativity

  • Radial trajectory



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..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  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. ^ ab Swerdlow, N. M. (2000). "Kepler's Iterative Solution to Kepler's Equation". Journal for the History of Astronomy. 31: 339–341. Bibcode:2000JHA....31..339S. doi:10.1177/002182860003100404.


  5. ^ It is often claimed that Kepler's equation "cannot be solved analytically"; see for example here. Whether this is true or not depends on whether one considers an infinite series (or one which does not always converge) to be an analytical solution. Other authors make the absurd claim that it cannot be solved at all; see for example M. V. K. Chari, Sheppard Joel Salon 2000 Technology & Engineering.


  6. ^ "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.


  7. ^ Stumpff, Karl (1 June 1968). "On The application of Lie-series to the problems of celestial mechanics". NASA Technical Note D-4460.


  8. ^ Colwell, Peter (1993). Solving Kepler's Equation Over Three Centuries. Willmann–Bell. p. 43. ISBN 0-943396-40-9.


  9. ^ Pfleger, Oliver Montenbruck, Thomas (1998). Astronomy on the Personal Computer (Third edition. ed.). Berlin, Heidelberg: Springer Berlin Heidelberg. ISBN 978-3-662-03349-4.




External links








  • Danby, J. M.; Burkardt, T. M. (1983). "The solution of Kepler's equation. I". Cel. Mech. 31: 95–107. Bibcode:1983CeMec..31...95D. doi:10.1007/BF01686811.


  • Conway, B. A. (1986). An improved algorithm due to Laguere for the solution of Kepler's equation. doi:10.2514/6.1986-84.


  • Mikkola, Seppo (1987). "A cubic approximation for Kepler's equation". Cel. Mech. 40 (3). Bibcode:1987CeMec..40..329M. doi:10.1007/BF01235850.


  • Nijenhuis, Albert (1991). "Solving Kepler's equation with high efficiency and accuracy". Cel. Mech. Dyn. Astr. 51 (4): 319–330. Bibcode:1991CeMDA..51..319N. doi:10.1007/BF00052925.


  • Fukushima, Toshio (1996). "A method solving kepler's equation without transcendental function evaluations". Cel. Mech. Dyn. Astron. 66 (3): 309–319. Bibcode:1996CeMDA..66..309F. doi:10.1007/BF00049384.


  • Charles, E. D.; Tatum, J. B. (1997). "The convergence of Newton-Raphson iteration with Kepler's equation". Cel. Mech. Dyn. Astr. 69 (4): 357–372. Bibcode:1997CeMDA..69..357C. doi:10.1023/A:1008200607490.


  • Stumpf, Laura (1999). "Chaotic behaviour in the newton iterative function associated with kepler's equation". Cel. Mech. Dyn. Astr. 74 (2): 95–109. doi:10.1023/A:1008339416143.


  • Palacios, M. (2002). "Kepler equation and accelerated Newton method". J. Comput. Appl. Math. 138: 335–346. Bibcode:2002JCoAM.138..335P. doi:10.1016/S0377-0427(01)00369-7.


  • 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". Appl. Num. Math. 57 (1): 12–18. doi:10.1016/j.apnum.2005.11.010.


  • Kepler's Equation at Wolfram Mathworld








Popular posts from this blog

Why do type traits not work with types in namespace scope?What are POD types in C++?Why can templates only be...

Will tsunami waves travel forever if there was no land?Why do tsunami waves begin with the water flowing away...

Should I use Docker or LXD?How to cache (more) data on SSD/RAM to avoid spin up?Unable to get Windows File...