Circle criterion: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Engineer Bob
define acronym and link
 
en>Michael Hardy
 
Line 1: Line 1:
Today, there are several other types of web development and blogging software available to design and host your website blogs online and that too in minutes, if not hours. You can either install Word - Press yourself or use free services offered on the web today. The effect is to promote older posts by moving them back onto the front page and into the rss feed. Hosted by Your Domain on Another Web Host - In this model, you first purchase multiple-domain webhosting, and then you can build free Wordpress websites on your own domains, taking advantage of the full power of Wordpress. If you are happy with your new look then click "Activate 'New Theme'" in the top right corner. <br><br>Most Word - Press web developers can provide quality CMS website solutions and they price their services at reasonable rates. WPTouch is among the more well known Word - Press smartphone plugins which is currently in use by thousands of users. This plugin allows a blogger get more Facebook fans on the related fan page. From my very own experiences, I will let you know why you should choose WPZOOM Live journal templates. Moreover, many Word - Press themes need to be purchased and designing your own WP site can be boring. <br><br>You can down load it here at this link:  and utilize your FTP software program to upload it to your Word - Press Plugin folder. The only problem with most is that they only offer a monthly plan, you never own the software and you can’t even install the software on your site, you must go to another website to manage your list and edit your autoresponder. Those who cannot conceive with donor eggs due to some problems can also opt for surrogacy option using the services of surrogate mother. To turn the Word - Press Plugin on, click Activate on the far right side of the list. Socrates: (link to  ) Originally developed for affiliate marketers, I've used this theme again and again to develop full-fledged web sites that include static pages, squeeze pages, and a blog. <br><br>A built-in widget which allows you to embed quickly video from popular websites. * Robust CRM to control and connect with your subscribers. re creating a Word - Press design yourself, the good news is there are tons of Word - Press themes to choose from. If you are looking for Hire Wordpress Developer then just get in touch with him. If your blog employs the permalink function, This gives your SEO efforts a boost, and your visitors will know firsthand what's in the post when seeing the URL. <br><br>More it extends numerous opportunities where your firm is at comfort and rest assured of no risks & errors. Automated deal feed integration option to populate your blog with relevant deals. If you loved this article and you would certainly like to obtain even more facts regarding [http://www.twinurl.com/wordpressbackupplugin86245 wordpress backup] kindly visit our webpage. As a result, it is really crucial to just take aid of some experience when searching for superior quality totally free Word - Press themes, Word - Press Premium Themes for your web site. )  Remote Login: With the process of PSD to Wordpress conversion comes the advantage of flexibility. Get started today so that people searching for your type of business will be directed to you.
[[File:OrbitalEccentricityDemo.svg|thumb|300px|A diagram of the various forms of the '''Kepler Orbit''' and their eccentricities. Blue is a hyperbolic trajectory (''e'' > 1). Green is a parabolic trajectory (''e'' = 1). Red is an elliptical orbit (''e'' < 1). Grey is a circular orbit (''e'' = 0).]]
 
:''For further closely relevant mathematical developments see also [[Two-body problem]], also [[Gravitational two-body problem]], and [[Kepler problem]].''
 
In [[celestial mechanics]], a '''Kepler orbit''' (or '''Keplerian orbit''') describes the motion of an orbiting body as an [[ellipse]], [[parabola]], or [[hyperbola]], which forms a two-dimensional [[Orbital plane (astronomy)|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 [[Perturbation (astronomy)|perturbations]] due to gravitational interactions with other objects, [[Drag (physics)|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 [[Parametrization|parametrized]] into six [[orbital elements]] in various ways.
 
In most applications, there is a large central body, the center of mass of which is assumed to be the center of mass of the entire system. By decomposition, the orbits of two objects of similar mass can be described as Kepler orbits around their common center of mass, their [[Barycentric coordinates (astronomy)|barycenter]].
 
==Introduction==
 
From ancient times until the 16th and 17th centuries, the motions of the planets were believed to follow perfectly circular [[geocentricism|geocentric]] paths as taught by the ancient Greek philosophers [[Aristotle]] and [[Ptolemy]]. Variations in the motions of the planets were explained by smaller circular paths overlaid on the larger path (see [[Deferent and epicycle|epicycle)]]. As measurements of the planets became increasingly accurate, revisions to the theory were proposed. In 1543, [[Nicolaus Copernicus]] published a [[heliocentric]] model of the solar system, although he still believed that the planets traveled in perfectly circular paths centered on the sun.{{citation needed|date=August 2012}}
 
===Johannes Kepler===
In 1601, [[Johannes Kepler]] acquired the extensive, meticulous observations of the planets made by [[Tycho Brahe]]. Kepler would spend the next five years trying to fit the observations of the planet Mars to various curves. In 1609, Kepler published the first two of his three [[Kepler's laws of planetary motion|laws of planetary motion]]. The first law states:
 
:"The [[orbit]] of every planet is an [[ellipse]] with the sun at a [[Focus (geometry)|focus]]."
 
More generally, the path of an object undergoing Keplerian motion may also follow a [[parabola]] or a [[hyperbola]], which, along with ellipses, belong to a group of curves known as [[conic sections]]. Mathematically, the distance between a central body and an orbiting body can be expressed as:
 
:<math> r(\nu) = \frac{a(1-e^2)}{1+e\cos(\nu)} </math>
 
where:
*<math>r</math> is the distance
*<math>a</math> is the [[semi-major axis]], which defines the size of the orbit
*<math>e</math> is the [[orbital eccentricity|eccentricity]], which defines the shape of the orbit
*<math>\nu</math> is the [[true anomaly]], which is the angle between the current position of the orbiting object and the location in the orbit at which it is closest to the central body (called the [[periapsis]])
Alternately, the equation can be expressed as:
 
:<math> r(\nu) = \frac{p}{1+e\cos(\nu)} </math>
 
Where <math>p</math> is called the [[conic section#Features|semi-latus rectum]] of the curve. This form of the equation is particularly useful when dealing with parabolic trajectories, for which the semi-major axis is infinite.
 
Despite developing these laws from observations, Kepler was never able to develop a theory to explain these motions.<ref>Bate, Mueller, White. pp 177–181</ref>
 
===Isaac Newton===
Between 1665 to 1666, [[Isaac Newton]] developed several concepts related to motion, gravitation and differential calculus. However, these concepts were not published until 1687 in the [[Philosophiæ Naturalis Principia Mathematica|Principia]], in which he outlined his [[Newton's laws of motion|laws of motion]] and his [[Newton's law of universal gravitation|law of universal gravitation]]. His second of his three laws of motion states:
 
<blockquote>
The [[acceleration]] a of a body is parallel and directly proportional to the net [[force]] acting on the body, is in the direction of the net force, and is inversely proportional to the [[mass]] of the body:
 
:<math> \mathbf{F} = m\mathbf{a} = m\frac{d^2\mathbf{r}}{dt^2}</math>
 
Where:
*<math>\mathbf{F}</math> is the force vector
*<math>m</math> is the mass of the body on which the force is acting
*<math>\mathbf{a}</math> is the acceleration vector, the second time derivative of the position vector <math>\mathbf{r}</math>
</blockquote>
Strictly speaking, this form of the equation only applies to an object of constant mass, which holds true based on the simplifying assumptions made below.
 
[[Image:NewtonsLawOfUniversalGravitation.svg|thumb|right|300px|The mechanisms of Newton's law of universal gravitation; a point mass ''m''<sub>1</sub> attracts another point mass ''m''<sub>2</sub> by a force ''F''<sub>2</sub> which is proportional to the product of the two masses and inversely proportional to the square of the distance (''r'') between them. Regardless of masses or distance, the magnitudes of <nowiki>|</nowiki>''F''<sub>1</sub><nowiki>|</nowiki> and <nowiki>|</nowiki>''F''<sub>2</sub><nowiki>|</nowiki> will always be equal. ''G'' is the [[gravitational constant]].]]
 
Newton's law of gravitation states:
 
<blockquote>
Every [[point mass]] attracts every other point mass by a [[force]] pointing along the line intersecting both points. The force is [[Proportionality (mathematics)|proportional]] to the product of the two masses and inversely proportional to the square of the distance between the point masses:
 
:<math>F = G \frac{m_1 m_2}{r^2}</math>
 
where:
 
* <math>F</math> is the magnitude of the gravitational force between the two point masses
* <math>G</math> is the [[gravitational constant]]
* <math>m_1</math> is the mass of the first point mass
* <math>m_2</math> is the mass of the second point mass
* <math>r</math> is the distance between the two point masses
 
</blockquote>
 
From the laws of motion and the law of universal gravitation, Newton was able to derive Kepler's laws, demonstrating consistency between observation and theory. The laws of Kepler and Newton formed the basis of modern [[celestial mechanics]] until [[Albert Einstein]] introduced the concepts of [[special relativity|special]] and [[general relativity|general]] relativity in the early 20th century. For most applications, Keplerian motion approximates the motions of planets and satellites to relatively high degrees of accuracy and is used extensively in [[astronomy]] and [[astrodynamics]].
 
==Simplified two body problem==
To solve for the motion of an object in a [[two-body problem|two body system]], two simplifying assumptions can be made:
:1. The bodies are spherically symmetric and can be treated as point masses.
:2. There are no external or internal forces acting upon the bodies other than their mutual gravitation.
 
The shapes of large celestial bodies are close to spheres. By symmetry, the net gravitational force attracting a mass point towards a homogeneous sphere must be directed towards its centre. The [[shell theorem]] (also proven by Isaac Newton) states that the magnitude of this force is the same as if all mass was concentrated in the middle of the sphere, even if the density of the sphere varies with depth (as it does for most celestial bodies). From this immediately follows that the attraction between two homogeneous spheres is as if both had its mass concentrated to its center.
 
Smaller objects, like [[asteroid]]s or [[spacecraft]] often have a shape strongly deviating from a sphere. But the gravitational forces produced by these irregularities are generally small compared to the gravity of the central body. The difference between an irregular shape and a perfect sphere also diminishes with distances, and most orbital distances are very large when compared with the diameter of a small orbiting body. Thus for some applications, shape irregularity can be neglected without significant impact on accuracy.
 
Planets rotate at varying rates and thus may take a slightly oblate shape because of the centrifugal force. With such an oblate shape, the gravitational attraction will deviate somewhat from that of a homogeneous sphere. This phenomenon is quite noticeable for artificial Earth satellites, especially those in low orbits. At larger distances the effect of this oblateness becomes negligible. Planetary motions in the Solar System can be computed with sufficient precision if they are treated as point masses.
 
Two point mass objects with masses <math>m_1</math> and <math>m_2</math> and position vectors <math>\mathbf{r}_1</math> and <math>\mathbf{r}_2</math> relative to some [[Inertial frame of reference|inertial reference frame]] experience gravitational forces:
 
:<math> m_1 \ddot{\mathbf{r}}_1 = \frac{-G m_1 m_2}{r^2} \mathbf{\hat{r}}</math>
 
:<math> m_2 \ddot{\mathbf{r}}_2 = \frac{G m_1 m_2}{r^2} \mathbf{\hat{r}}</math>
 
where <math>\mathbf{r}</math> is the relative position vector of mass 1 with respect to mass 2, expressed as:
 
:<math> \mathbf{r} = \mathbf{r}_1 - \mathbf{r}_2 </math>
 
and <math>\mathbf{\hat{r}}</math> is the [[unit vector]] in that direction and <math>r</math> is the [[Euclidean vector#Length|length]] of that vector.
 
Dividing by their respective masses and subtracting the second equation from the first yields the equation of motion for the acceleration of the first object with respect to the second:
 
{{NumBlk|:|<math> \ddot{\mathbf{r}} = - \frac{\mu}{r^2} \mathbf{\hat{r}}</math>|{{EquationRef|1}}}}
 
where <math>\mu</math> is the gravitational parameter and is equal to
 
:<math> \mu = G(m_1 + m_2)</math>
 
In many applications, a third simplifying assumption can be made:
:3. When compared to the central body, the mass of the orbiting body is insignificant. Mathematically, ''m''<sub>1</sub> >> ''m''<sub>2</sub>, so ''&mu;'' = ''G'' (''m''<sub>1</sub>&nbsp;+&nbsp;''m''<sub>2</sub>) &asymp; ''Gm''<sub>1</sub>.
This assumption is not necessary to solve the simplified two body problem, but it simplifies calculations, particularly with Earth-orbiting satellites and planets orbiting the sun. Even [[Jupiter]]'s mass is less than the sun's by a factor of 1047,<ref>http://ssd.jpl.nasa.gov</ref> which would constitute an error of 0.096% in the value of &mu;. Notable exceptions include the Earth-moon system (mass ratio of 81.3), the Pluto-Charon system (mass ratio of 8.9) and binary star systems.
 
Under these assumptions the differential equation for the two body case can be completely solved mathematically and the resulting orbit which follows Kepler's laws of planetary motion is called a "Kepler orbit". The orbits of all planets are to high accuracy Kepler orbits around the Sun. The small deviations are due to the much weaker gravitational attractions between the planets, and in the case of [[Mercury (planet)|Mercury]], due to [[general relativity]]. The orbits of the artificial satellites around the Earth are, with a fair approximation, Kepler orbits with small perturbations due to the gravitational attraction of the sun, the moon and the oblateness of the Earth. In high accuracy applications for which the equation of motion must be integrated numerically with all gravitational and non-gravitational forces (such as [[solar radiation pressure]] and [[Drag (physics)|atmospheric drag]]) being taken into account, the Kepler orbit concepts are of paramount importance and heavily used.
 
===Keplerian elements===
[[Image:Orbit1.svg|thumb|300px|right|Keplerian [[orbital elements]].]]
{{Main|Keplerian elements}}
 
It is worth mentioning that any Keplerian trajectory can be defined by six parameters. The motion of an object moving in three-dimensional space is characterized by a position vector and a velocity vector. Each vector has three components, so the total number of values needed to define a trajectory through space is six. An orbit is generally defined by six elements (known as ''Keplerian elements'') that can be computed from position and velocity, three of which have already been discussed. These elements are convenient in that of the six, five are unchanging for an unperturbed orbit (a stark contrast to two constantly changing vectors). The future location of an object within its orbit can be predicted and its new position and velocity can be easily obtained from the orbital elements.
 
Two define the size and shape of the trajectory:
*[[Semimajor axis]] (<math>a\,\!</math>)
*[[Eccentricity (orbit)|Eccentricity]] (<math>e\,\!</math>)
 
Three define the orientation of the [[Orbital plane (astronomy)|orbital plane]]:
*[[Inclination]] (<math>i\,\!</math>) defines the angle between the orbital plane and the reference plane.
*[[Longitude of the ascending node]] (<math>\Omega\,\!</math>) defines the angle between the reference direction and the upward crossing of the orbit on the reference plane (the ascending node).
*[[Argument of periapsis]] (<math>\omega\,\!</math>) defines the angle between the ascending node and the periapsis.
 
And finally:
*[[True anomaly]] (<math>\nu</math>) defines the position of the orbiting body along the trajectory, measured from periapsis. Several alternate values can be used instead of true anomaly, the most common being <math>M</math> the [[mean anomaly]] and <math>T</math>, the time since periapsis.
 
Because <math>i</math>, <math>\Omega</math> and <math>\omega</math> are simply angular measurements defining the orientation of the trajectory in the reference frame, they are not strictly necessary when discussing the motion of the object within the orbital plane. They have been mentioned here for completeness, but are not required for the proofs below.
 
==Mathematical solution of the differential equation ({{EquationNote|1}}) above==
For movement under any central force, i.e. a force parallel to '''r''', the [[specific relative angular momentum]] <math> \bold{H} = \bold{r} \times {\dot{\bold{r}}} </math> stays constant:<br/>
<math> \dot {\bold{H}} = \frac{d}{dt}\left(\bold{r} \times {\dot{\bold{r}}}\right) = \dot{\bold{r}} \times {\dot{\bold{r}}} + \bold{r} \times {\ddot{\bold{r}}} =\bold{0} + \bold{0} = \bold{0}</math><br/>
 
Since the cross product of the position vector and its velocity stays constant, they must lie in the same plane, orthogonal to <math> \bold{H} </math>. This implies the vector function is a [[plane curve]]. <br/>
Because the equation has symmetry around its origin, it is easier to solve in polar coordinates. However, it is important to note that equation ({{EquationNote|1}}) refers to linear acceleration <math> \left (\ddot{\bold{r}} \right )</math>, as opposed to angular <math> \left (\ddot{\theta} \right )</math> or radial <math> \left (\ddot{r} \right )</math> acceleration. Therefore, one must be cautious when transforming the equation.
Introducing a cartesian coordinate system <math>(\hat{\bold{x}} \ , \ \hat{\bold{y}})</math> and [[unit vector#Cylindrical coordinates|polar unit vectors]] <math>(\hat{\bold{r}} \ , \ \hat{\boldsymbol\theta})</math> in the plane orthogonal to <math> \bold{H} </math>:<br/>
 
<math> \hat{\bold{r}}=\cos(\theta)\hat{\bold{x}} + \sin(\theta)\hat{\bold{y}} </math><br/>
<math> \hat{\boldsymbol\theta}=-\sin(\theta)\hat{\bold{x}} + \cos(\theta)\hat{\bold{y}} </math><br/>
 
We can now rewrite the vector function <math>\bold{r}</math> and its derivatives as:<br/>
 
<math> \bold{r} =r ( \cos\theta \hat{x} + \sin \theta \hat{y}) = r\hat{\mathbf{r}} </math><br/>
 
<math> \dot{\bold{r}} = \dot r \hat {\mathbf r} + r \dot \theta \hat {\boldsymbol{\theta}} </math><br/>
 
<math> \ddot{\bold{r}} =  (\ddot r - r\dot\theta^2)\hat{\mathbf{r}} + (r \ddot\theta + 2 \dot r \dot\theta) \hat{\boldsymbol\theta} </math><br/>
 
(see "[[Polar coordinates#Vector calculus]]"). Substituting these into ({{EquationNote|1}}), we find:<br/>
 
<math> (\ddot r - r\dot\theta^2)\hat{\mathbf{r}} + (r \ddot\theta + 2 \dot r \dot\theta) \hat{\boldsymbol\theta} = \left (-\frac{\mu}{r^2}\right )\hat{\mathbf{r}} + (0)\hat{\boldsymbol\theta}</math><br/>
 
This gives the non-ordinary polar differential equation:
 
# {{NumBlk|:|<math>\ddot{r} - r {\dot{\theta}}^2 = - \frac {\mu} {r^2}</math><br/>|{{EquationRef|2}}}}
 
In order to solve this equation, we must first eliminate all time derivatives. We find that:
 
<math> H = |\bold{r} \times  {\dot{\bold{r}}}| = |(r\cos(\theta), r\sin(\theta), 0) \times (\dot{r}\cos(\theta)-r\sin(\theta)\dot{\theta}, \dot{r}\sin(\theta)+r\cos(\theta)\dot{\theta}, 0)| = |(0,0,r^2\dot\theta)| = r^2\dot\theta</math>
# {{NumBlk|:|<math> \dot\theta = \frac{H}{r^2} </math>|{{EquationRef|3}}}}
 
Taking the time derivative of ({{EquationNote|3}}), we get
 
# {{NumBlk|:|<math>\ddot{\theta}  = - \frac {2 \cdot H \cdot \dot{r}} {r^3}</math>|{{EquationRef|4}}}}
 
Equations ({{EquationNote|3}}) and ({{EquationNote|4}}) allow us to eliminate the time derivatives of  <math>\theta</math>. In order to eliminate the time derivatives of  <math>r</math>, we must use the chain rule to find appropriate substitutions:
 
# {{NumBlk|:|<math>\dot{r}  =  \frac {dr} {d\theta} \cdot \dot {\theta}</math>|{{EquationRef|5}}}}
 
# {{NumBlk|:|<math>\ddot{r}  =  \frac {d^2r} {d\theta^2} \cdot {\dot {\theta}}^2 + \frac {dr} {d\theta} \cdot \ddot {\theta}</math>|{{EquationRef|6}}}}
 
Using these four substitutions, all time derivatives in ({{EquationNote|2}}) can be eliminated, yielding an [[ordinary differential equation]] for <math>r</math> as function of <math>\theta\,</math>.<br/>
 
<math>\ddot{r} - r {\dot{\theta}}^2 = - \frac {\mu} {r^2}</math><br/>
 
<math>\frac {d^2r} {d\theta^2} \cdot {\dot {\theta}}^2 + \frac {dr} {d\theta} \cdot \ddot {\theta} - r {\dot{\theta}}^2 = - \frac {\mu} {r^2}</math><br/>
 
<math>\frac {d^2r} {d\theta^2} \cdot \left (\frac{H}{r^2} \right )^2 + \frac {dr} {d\theta} \cdot \left (- \frac {2 \cdot H \cdot \dot{r}} {r^3} \right ) - r\left (\frac{H}{r^2} \right )^2 = - \frac {\mu} {r^2}</math>
# {{NumBlk|:|<math>\frac {H^2} {r^4} \cdot \left ( \frac{d^2 r} {d\theta ^2} - 2 \cdot \frac{\left (\frac {dr} {d\theta} \right ) ^2}
{r} - r\right )= - \frac {\mu} {r^2}</math>|{{EquationRef|7}}}}
The differential equation ({{EquationNote|7}}) can be solved analytically by the variable substitution
# {{NumBlk|:|<math>r=\frac{1} {s}</math>|{{EquationRef|8}}}}
 
Using the chain rule for differentiation one gets:
 
# {{NumBlk|:|<math>\frac {dr} {d\theta} = -\frac {1} {s^2} \cdot \frac {ds} {d\theta}</math>|{{EquationRef|9}}}}
# {{NumBlk|:|<math>\frac {d^2r} {d\theta^2} = \frac {2} {s^3} \cdot \left (\frac {ds} {d\theta}\right )^2 -
\frac {1} {s^2} \cdot \frac {d^2s} {d\theta^2}</math>|{{EquationRef|10}}}}
 
Using the expressions ({{EquationNote|10}}) and ({{EquationNote|9}}) for <math>\frac {d^2r} {d\theta^2}</math> and <math>\frac {dr} {d\theta}</math>
one gets
# {{NumBlk|:|<math>H^2 \cdot  \left ( \frac {d^2s} {d\theta^2} + s \right ) = \mu</math>|{{EquationRef|11}}}}
with the general solution
# {{NumBlk|:|<math>s = \frac {\mu} {H^2} \cdot \left ( 1 + e \cdot \cos (\theta-\theta_0)\right )</math>|{{EquationRef|12}}}}
 
where ''e'' and <math>\theta_0\,</math> are constants of integration depending on the initial values for ''s'' and <math>\frac {ds} {d\theta}</math>.
 
Instead of using the constant of integration <math>\theta_0\,</math> explicitly one introduces the convention that the unit vectors <math> \hat{x} \ , \ \hat{y}</math> defining the coordinate system in the orbital plane are selected such that <math> \theta_0\,</math> takes the value zero and ''e'' is positive. This then means that <math>\theta\,</math> is zero at the point where <math> s</math> is maximal and therefore <math> r= \frac {1}{s}</math> is minimal. Defining the parameter p as <math> \frac {H^2}{\mu}</math> one has that<br/>
<math>r = \frac {1}{s} = \frac {p}{1 + e \cdot \cos \theta}</math>
 
===Alternate derivation===
Another way to solve this equation without the use of polar differential equations is as follows:<br/>Define a unit vector <math>\bold{u}</math> such that <math>\bold{r} = r\bold{u}</math> and <math> \ddot{\bold{r}} = -\frac{\mu}{r^2}\bold{u}</math>. It follows that<br/>
 
<math>\bold{H} = \bold{r} \times \dot{\bold{r}} = r\bold{u} \times \frac{d}{dt}(r\bold{u}) = r\bold{u} \times (r\dot{\bold{u}}+\dot{r}\bold{u}) = r^2(\bold{u} \times \dot{\bold{u}}) + r\dot{r}(\bold{u} \times \bold{u}) = r^2\bold{u} \times \dot{\bold{u}}</math><br/>
 
Now consider<br/>
 
<math>\ddot{\bold{r}} \times \bold{H} = -\frac{\mu}{r^2}\bold{u} \times (r^2\bold{u} \times \dot{\bold{u}}) = -\mu\bold{u} \times (\bold{u} \times \dot{\bold{u}}) = -\mu[(\bold{u}\cdot\dot{\bold{u}})\bold{u}-(\bold{u}\cdot\bold{u})\dot{\bold{u}}]</math><br/>
 
(see [[Triple product#Vector triple product]]). Notice that<br/>
 
<math>\bold{u}\cdot\bold{u} = |\bold{u}|^2 = 1</math><br/>
 
<math>\bold{u}\cdot\dot{\bold{u}} = \frac{1}{2}(\bold{u}\cdot\dot{\bold{u}} + \dot{\bold{u}}\cdot\bold{u}) = \frac{1}{2}\frac{d}{dt}(\bold{u}\cdot\bold{u}) = 0 </math><br/>
 
Substituting these values into the previous equation, one gets:<br/>
 
<math>\ddot{\bold{r}}\times\bold{H}=\mu\dot{\bold{u}}</math><br/>
 
Integrating both sides:<br/>
 
<math>\dot{\bold{r}}\times\bold{H}=\mu\bold{u} + \bold{c}</math><br/>
 
Where '''c''' is a constant vector. Dotting this with '''r''' yields an interesting result:<br/>
 
<math> \bold{r}\cdot(\dot{\bold{r}}\times\bold{H})=\bold{r}\cdot(\mu\bold{u} + \bold{c}) = \mu\bold{r}\cdot\bold{u} + \bold{r}\cdot\bold{c} = \mu r(\bold{u}\cdot\bold{u})+rc\cos(\theta)=r(\mu + c\cos(\theta))</math><br/>
 
Where <math>\theta</math> is the angle between <math>\bar{r}</math> and <math>\bar{c}</math>. Solving for r: <br/>
 
<math> r = \frac{\bold{r}\cdot(\dot{\bold{r}}\times\bold{H})}{\mu + c\cos(\theta)} = \frac{(\bold{r}\times\dot{\bold{r}})\cdot\bold{H}}{\mu + c\cos(\theta)} = \frac{|\bold{H}|^2}{\mu + c\cos(\theta)}</math><br/>
 
Notice that <math>(r,\theta)</math> are effectively the polar coordinates of the vector function. Making the substitutions <math>p=\frac{|\bold{H}|^2}{\mu}</math> and <math>e=\frac{c}{\mu}</math>, we again arrive at the equation
# {{NumBlk|:|<math>r = \frac {p}{1 + e \cdot \cos \theta}</math>|{{EquationRef|13}}}}
 
This is the equation in polar coordinates for a [[conic section]] with origin in a focal point. The argument <math>\theta\,</math> is called "true anomaly".<br/>
 
===Properties of trajectory equation===
For <math>e\ =\ 0\,</math> this is a circle with radius ''p''.
 
For <math>0\ < e\ <\ 1\,</math> this is an [[ellipse]] with
# {{NumBlk|:|<math>a = \frac {p}{1-e^2}</math>|{{EquationRef|14}}}}
# {{NumBlk|:|<math>b = \frac {p}{\sqrt{1-e^2}} = a \cdot \sqrt{1-e^2}</math>|{{EquationRef|15}}}}
 
For <math>e\ =\ 1\,</math> this is a [[parabola]] with focal length <math>\frac {p}{2}</math>
 
For <math>e\ >\ 1\,</math> this is a [[hyperbola]] with
# {{NumBlk|:|<math>a = \frac {p}{e^2-1}</math>|{{EquationRef|16}}}}
# {{NumBlk|:|<math>b = \frac {p}{\sqrt{e^2-1}} = a \cdot \sqrt{e^2-1}</math>|{{EquationRef|17}}}}
 
The following image illustrates an ellipse (red), a parabola (green) and a hyperbola (blue)
 
[[File:Kepler orbits.svg|thumb|An elliptic Kepler orbit with an eccentricity of 0.7, a parabolic Kepler orbit and a hyperbolic Kepler orbit with an eccentricity of 1.3. The distance to the focal point is a function of the polar angle relative to the horizontal line as given by the equation ({{EquationNote|13}})]]
 
The point on the horizontal line going out to the right from the focal point is the point with <math>\theta = 0\,</math> for which the distance to the focus takes the minimal value <math>\frac {p}{1 + e}</math>, the pericentre. For the ellipse there is also an apocentre for which the distance to the focus takes the maximal value <math>\frac {p}{1 - e}</math>. For the hyperbola the range for <math>\theta\,</math> is
:<math>\left [ -\cos^{-1}\left(-\frac{1}{e}\right) < \theta < \cos^{-1}\left(-\frac{1}{e}\right)\right ]</math>
and for a parobola the range is
:<math>\left [ -\pi < \theta < \pi \right ]</math>
 
Using the chain rule for differentiation ({{EquationNote|5}}), the equation ({{EquationNote|2}}) and the definition of ''p'' as <math>\frac {H^2}{\mu}</math> one gets that the radial velocity component is
# {{NumBlk|:|<math>V_r = \dot{r} = \frac {H}{p} \cdot e \cdot \sin \theta =  \sqrt{\frac {\mu}{p}} \cdot e \cdot \sin \theta</math>|{{EquationRef|18}}}}
 
and that the tangential component (velocity component perpendicular to <math>V_r</math>) is
# {{NumBlk|:|<math>V_t = r \cdot \dot{\theta} = \frac {H}{r} =  \sqrt{\frac {\mu}{p}} \cdot (1 + e \cdot \cos \theta)</math>|{{EquationRef|19}}}}
 
The connection between the polar argument <math>\theta\,</math> and time ''t'' is slightly different for elliptic and hyperbolic orbits.
 
For an elliptic orbit one switches to the "[[eccentric anomaly]]" ''E'' for which
# {{NumBlk|:|<math>x = a \cdot (\cos E -e)</math>|{{EquationRef|20}}}}
# {{NumBlk|:|<math>y = b \cdot \sin E</math>|{{EquationRef|21}}}}
and consequently
# {{NumBlk|:|<math>\dot{x} = -a \cdot \sin E \cdot \dot{E}</math>|{{EquationRef|22}}}}
# {{NumBlk|:|<math>\dot{y} = b \cdot \cos E \cdot \dot{E}</math>|{{EquationRef|23}}}}
 
and the angular momentum ''H'' is
# {{NumBlk|:|<math>H = x \cdot \dot{y} - y \cdot \dot{x}=a \cdot b \cdot ( 1 - e \cdot \cos E) \cdot \dot{E}</math>|{{EquationRef|24}}}}
 
Integrating with respect to time ''t'' one gets
# {{NumBlk|:|<math>H \cdot t = a \cdot b \cdot ( E - e \cdot \sin E)</math>|{{EquationRef|25}}}}
 
under the assumption that time <math>t=0</math> is selected such that the integration constant is zero.
 
As by definition of ''p'' one has
# {{NumBlk|:|<math>H = \sqrt{\mu \cdot p}</math>|{{EquationRef|26}}}}
this can be written
# {{NumBlk|:|<math>t = a \cdot \sqrt{\frac{a} {\mu}} ( E - e \cdot \sin E)</math>|{{EquationRef|27}}}}
 
For a hyperbolic orbit one uses the [[hyperbolic functions]] for the parameterisation
# {{NumBlk|:|<math>x = a \cdot (e - \cosh E)</math>|{{EquationRef|28}}}}
# {{NumBlk|:|<math>y = b \cdot \sinh E</math>|{{EquationRef|29}}}}
for which one has
# {{NumBlk|:|<math>\dot{x} = -a \cdot  \sinh E \cdot \dot{E}</math>|{{EquationRef|30}}}}
# {{NumBlk|:|<math>\dot{y} = b \cdot \cosh E \cdot \dot{E}</math>|{{EquationRef|31}}}}
and the angular momentum ''H'' is
# {{NumBlk|:|<math>H = x \cdot \dot{y} - y \cdot \dot{x}=a \cdot b \cdot (  e \cdot \cosh E-1) \cdot \dot{E}</math>|{{EquationRef|32}}}}
Integrating with respect to time ''t'' one gets
# {{NumBlk|:|<math>H \cdot t= a \cdot b \cdot ( e \cdot \sinh E-E)</math>|{{EquationRef|33}}}}
i.e.
# {{NumBlk|:|<math>t = a \cdot \sqrt{\frac{a} {\mu}} (e \cdot \sinh E-E)</math>|{{EquationRef|34}}}}
 
To find what time t that corresponds to a certain true anomaly <math>\theta\,</math> one computes corresponding parameter ''E'' connected to time with relation ({{EquationNote|27}}) for an elliptic and with relation ({{EquationNote|34}}) for a hyperbolic orbit.
 
Note that the relations ({{EquationNote|27}}) and ({{EquationNote|34}}) define a mapping between the ranges
:<math>\left [ -\infin < t < \infin\right ]  \longleftrightarrow \left [-\infin < E < \infin \right ] </math>
 
==Some additional formulae==
''See also [[Equation of the center#Analytical expansions|Equation of the center &ndash; Analytical expansions]]''
 
For an elliptic orbit one gets from ({{EquationNote|20}}) and ({{EquationNote|21}}) that
# {{NumBlk|:|<math>r = a \cdot (1-e \cdot \cos E)</math>|{{EquationRef|35}}}}
and therefore that
# {{NumBlk|:|<math>\cos \theta = \frac{x} {r} =\frac{\cos E-e}{1-e \cdot \cos E}</math>|{{EquationRef|36}}}}
From ({{EquationNote|36}}) then follows that
:<math>
\tan^2 \frac{\theta}{2} =
\frac{1-\cos \theta}{1+\cos \theta}=
\frac{1-\frac{\cos E-e}{1-e \cdot \cos E}}{1+\frac{\cos E-e}{1-e \cdot \cos E}}=
\frac{1-e \cdot \cos E  -  \cos E+e}{1-e \cdot \cos E  +  \cos E-e}=
\frac{1+e}{1-e} \ \cdot\ \frac{1-\cos E}{1+\cos E}=
\frac{1+e}{1-e} \ \cdot\ \tan^2 \frac{E}{2}
</math>
 
From the geometrical construction defining the [[eccentric anomaly]] it is clear that the vectors <math>(\ \cos E\ ,\ \sin E\ )</math> and
<math>(\ \cos \theta\ ,\ \sin \theta\ )</math> are on the same side of the ''x''-axis. From this then follows that the vectors
<math>\left( \cos\frac{E}{2}\ ,\ \sin\frac{E}{2} \right)</math> and <math>\left( \cos\frac{\theta}{2}\ ,\ \sin\frac{\theta}{2} \right)</math> are in the same quadrant. One therefore has that
# {{NumBlk|:|<math>\tan \frac{\theta}{2} = \sqrt{\frac{1+e}{1-e}} \cdot \tan \frac{E}{2}</math>|{{EquationRef|37}}}}
 
and that
 
# {{NumBlk|:|<math>\theta = 2 \cdot \operatorname{arg}\left(\sqrt{1-e} \ \cdot\  \cos \frac{E}{2}\ ,\ \sqrt{1+e} \ \cdot\  \sin\frac{E}{2}\right)+ n\cdot 2\pi</math>|{{EquationRef|38}}}}
# {{NumBlk|:|<math>E = 2 \cdot \operatorname{arg}\left(\sqrt{1+e} \ \cdot\  \cos \frac{\theta}{2}\ ,\ \sqrt{1-e} \ \cdot\  \sin\frac{\theta}{2}\right)+ n\cdot 2\pi</math>|{{EquationRef|39}}}}
 
where "<math>\operatorname{arg}(x\ ,\ y)</math>" is the polar argument of the vector <math>(\ x\ ,\ y\ )</math> and ''n'' is selected such that <math>\left |E
-\theta\right| < \pi </math>
 
For the numerical computation of <math>\operatorname{arg}(x\ ,\ y)</math> the standard function [[Atan2|ATAN2(y,x)]]
(or in [[double precision]] DATAN2(y,x)) available in for example the programming language [[Fortran|FORTRAN]] can be used.
 
Note that this is a mapping between the ranges
 
:<math>\left [ -\infin < \theta < \infin\right ]  \longleftrightarrow \left [-\infin < E < \infin \right ] </math>
 
For an hyperbolic orbit one gets from ({{EquationNote|28}}) and ({{EquationNote|29}}) that
# {{NumBlk|:|<math>r = a \cdot (e \cdot \cosh E-1)</math>|{{EquationRef|40}}}}
and therefore that
# {{NumBlk|:|<math>\cos \theta = \frac{x} {r} =\frac{e-\cosh E}{e \cdot \cosh E-1}</math>|{{EquationRef|41}}}}
 
As
:<math>
\tan^2 \frac{\theta}{2} =
\frac{1-\cos\theta}{1+\cos \theta}=
\frac{1-\frac{e-\cosh E}{e \cdot \cosh E-1}}{1+\frac{e-\cosh E}{e \cdot \cosh E-1}}=
\frac{e \cdot \cosh E - e +\cosh E}{e \cdot \cosh E + e -\cosh E}=
\frac{e+1}{e-1}\ \cdot\ \frac{\cosh E-1}{\cosh E+1}=
\frac{e+1}{e-1}\ \cdot\ \tanh^2 \frac{E}{2}
</math>
and as <math> \tan \frac{\theta}{2}</math> and <math>\tanh \frac{E}{2}</math> have the same sign it follows that
# {{NumBlk|:|<math>\tan \frac{\theta}{2} = \sqrt{\frac{e+1}{e-1}} \cdot \tanh \frac{E}{2}</math>|{{EquationRef|42}}}}
This relation is convenient for passing between "true anomaly" and the parameter
''E'', the latter being connected to time through relation ({{EquationNote|34}}). Note that this is a mapping between the ranges
 
:<math>\left [ -\cos^{-1}\left(-\frac{1}{e}\right) < \theta < \cos^{-1}\left(-\frac{1}{e}\right)\right ]  \longleftrightarrow \left [-\infin < E < \infin \right ] </math>
 
and that <math>\frac{E}{2}</math> can be computed using the relation
 
:<math>\tanh ^{-1}x=\frac{1}{2}\ln \left( \frac{1+x}{1-x} \right)</math>
 
From relation ({{EquationNote|27}}) follows that the orbital period ''P'' for an elliptic orbit is
# {{NumBlk|:|<math>P =  2\pi \cdot a \cdot \sqrt{\frac{a} {\mu}}</math>|{{EquationRef|43}}}}
 
As the potential energy corresponding to the force field of relation ({{EquationNote|1}}) is
:<math>-\frac {\mu} {r}</math>
it follows from ({{EquationNote|13}}), ({{EquationNote|14}}), ({{EquationNote|18}}) and ({{EquationNote|19}}) that the sum of the kinetic and the potential energy
 
:<math>\frac{{V_r}^2+{V_t}^2}{2}-\frac {\mu} {r}</math>
 
for an elliptic orbit is
 
# {{NumBlk|:|<math>-\frac {\mu} {2 \cdot a}</math>|{{EquationRef|44}}}}
and from ({{EquationNote|13}}), ({{EquationNote|16}}), ({{EquationNote|18}}) and ({{EquationNote|19}}) that the sum of the kinetic and the potential energy for a hyperbolic orbit is
# {{NumBlk|:|<math>\frac {\mu} {2 \cdot a}</math>|{{EquationRef|45}}}}
 
Relative the inertial coordinate system
 
:<math> \hat{x} \ , \ \hat{y}</math>
 
in the orbital plane with <math> \hat{x}</math> towards pericentre one gets from ({{EquationNote|18}}) and ({{EquationNote|19}}) that the velocity components are
 
# {{NumBlk|:|<math>V_x = \cos \theta \cdot V_r - \sin \theta \cdot V_t = -\sqrt{\frac {\mu}{p}} \cdot \sin \theta</math>|{{EquationRef|46}}}}
# {{NumBlk|:|<math>V_y = \sin \theta \cdot V_r + \cos \theta \cdot V_t = \sqrt{\frac {\mu}{p}} \cdot (e +\cos \theta)</math>|{{EquationRef|47}}}}
 
==Determination of the Kepler orbit that corresponds to a given initial state==
 
This is the "[[initial value problem]]" for the differential equation ({{EquationNote|1}}) which is a first order equation
for the 6-dimensional "state vector" <math>(\ \bar{r}\ ,\bar{v}\ )</math> when written as
 
# {{NumBlk|:|<math>\dot {\bar{v}} = -\mu \cdot \frac {\hat{r}} {r^2}</math>|{{EquationRef|48}}}}
# {{NumBlk|:|<math>\dot {\bar{r}} = \bar{v}</math>|{{EquationRef|49}}}}
 
For any values for the initial "state vector" <math>(\ \bar{r_0}\ ,\bar{v_0}\ )</math> the Kepler orbit corresponding
to the solution of this initial value problem can be found with the following algorithm:
 
Define the orthogonal unit vectors <math>(\hat{r}\ ,\ \hat{t})</math> through
 
# {{NumBlk|:|<math>\bar{r_0} = r \cdot \hat{r}</math>|{{EquationRef|50}}}}
# {{NumBlk|:|<math>\bar{v_0} = V_r \cdot \hat{r} + V_t \cdot \hat{t}</math>|{{EquationRef|51}}}}
 
with <math>r > 0</math> and <math>V_t > 0</math>
 
From ({{EquationNote|13}}), ({{EquationNote|18}}) and ({{EquationNote|19}}) follows that by setting
 
# {{NumBlk|:|<math>p = \frac{{(r \cdot V_t)}^2}{\mu }</math>|{{EquationRef|52}}}}
 
and by defining <math>e \ge 0</math> and <math>\theta</math> such that
 
# {{NumBlk|:|<math>e \cdot \cos \theta = \frac{V_t} {V_0} - 1</math>|{{EquationRef|53}}}}
# {{NumBlk|:|<math>e \cdot \sin \theta = \frac{V_r} {V_0}</math>|{{EquationRef|54}}}}
 
where
 
# {{NumBlk|:|<math>V_0 = \sqrt{\frac{\mu}{p}}</math>|{{EquationRef|55}}}}
 
one gets a Kepler orbit that for true anomaly <math>\theta</math> has the same ''r'', <math>V_r</math> and <math>V_t</math> values as those defined by ({{EquationNote|50}}) and ({{EquationNote|51}}).
 
If this Kepler orbit then also has the same <math>(\hat{r}\ ,\ \hat{t})</math> vectors for this true anomaly <math>\theta</math> as the ones defined by ({{EquationNote|50}}) and ({{EquationNote|51}}) the state vector <math>(\bar{r}\ ,\ \bar{v})</math> of the Kepler orbit takes the desired values <math>(\ \bar{r_0}\ ,\bar{v_0}\ )</math> for true anomaly <math>\theta</math>.
 
The standard inertially fixed coordinate system <math>(\hat{x}\ ,\ \hat{y})</math> in the orbital plane (with <math>\hat{x}</math> directed from the centre of the homogeneous sphere to the pericentre) defining the orientation of the conical section (ellipse, parabola or hyperbola) can then be determined with the relation
 
# {{NumBlk|:|<math>\hat{x} = \cos \theta \cdot \hat{r} - \sin \theta \cdot \hat{t}</math>|{{EquationRef|56}}}}
# {{NumBlk|:|<math>\hat{y} = \sin \theta \cdot \hat{r} + \cos \theta \cdot \hat{t}</math>|{{EquationRef|57}}}}
 
Note that the relations ({{EquationNote|53}}) and ({{EquationNote|54}}) has a singularity when <math>V_r=0</math> and
 
:<math>V_t=V_0=\sqrt{\frac{\mu}{p}}=\sqrt{\frac{\mu}{\frac{{(r \cdot V_t)}^2}{\mu }}}</math>
 
i.e.
 
# {{NumBlk|:|<math>V_t=\sqrt{\frac{\mu}{r}}</math>|{{EquationRef|58}}}}
 
which is the case that it is a circular orbit that is fitting the initial state <math>(\ \bar{r_0}\ ,\bar{v_0}\ )</math>
 
==The osculating Kepler orbit==
{{Main|Osculating orbit}}
For any state vector <math>(\bar{r} , \bar{v} )</math> the Kepler orbit corresponding to this state can be computed with the algorithm defined above.
First the parameters <math>p ,  e ,  \theta</math> are determined from <math>r ,  V_r ,  V_t</math> and then
the orthogonal unit vectors in the orbital plane <math> \hat{x} , \hat{y}</math> using the relations ({{EquationNote|56}}) and ({{EquationNote|57}}).
 
If now the equation of motion is
 
# {{NumBlk|:|<math>\ddot {\bar{r}} = \operatorname{\bar{F}}(\bar{r},\dot {\bar{r}},t)</math>|{{EquationRef|59}}}}
 
where
 
:<math>\operatorname{\bar{F}}(\bar{r},\dot {\bar{r}},t)</math>
 
is a function other than
 
:<math>-\mu \cdot  \frac {\hat{r}} {r^2}</math>
 
the resulting parameters
 
<math>p ,\,  e ,\,  \theta,\,  \hat{x} ,\,  \hat{y}</math>
 
defined by <math>\bar{r},\dot {\bar{r}}</math> will all vary with time as opposed to the case of a Kepler orbit for which only the parameter
<math>\theta</math> will vary
 
The Kepler orbit computed in this way having the same "state vector" as the solution to the "equation of motion" ({{EquationNote|59}}) at time ''t'' is said to be "osculating" at this time.
 
This concept is for example useful in case
:<math>\operatorname{\bar{F}}(\bar{r},\dot {\bar{r}},t)=-\mu \cdot \frac {\hat{r}} {r^2}+\operatorname{\bar{f}}(\bar{r},\dot {\bar{r}},t)</math>
 
where
 
:<math>\operatorname{\bar{f}}(\bar{r},\dot {\bar{r}},t)</math>
 
is a small "perturbing force" due to for example a faint gravitational pull from other celestial bodies. The parameters of the osculating Kepler orbit will then only slowly change and the osculating Kepler orbit is a good approximation to the real orbit for a considerable time period before and after the time of osculation.
 
This concept can also be useful for a rocket during powered flight as it then tells which Kepler orbit
the rocket would continue in in case the thrust is switched-off.
 
For a "close to circular" orbit the concept "[[eccentricity vector]]" defined as <math>\bar{e}=e \cdot \hat{x}</math> is useful. From ({{EquationNote|53}}), ({{EquationNote|54}}) and ({{EquationNote|56}}) follows that
 
# {{NumBlk|:|<math>\bar{e}=\frac{(V_t-V_0) \cdot \hat{r} - V_r \cdot \hat{t}}{V_0}</math>|{{EquationRef|60}}}}
 
i.e. <math>\bar{e}</math> is a smooth differentiable function of the state vector <math>( \bar{r} ,\bar{v} )</math> also if this state corresponds to a circular orbit.
 
== See also ==
{{Portal|Physics}}
* [[Kepler's laws of planetary motion]]
* [[Elliptic orbit]]
* [[Hyperbolic trajectory]]
* [[Parabolic trajectory]]
* [[Radial trajectory]]
* [[Orbit modeling]]
 
==Citations==
{{Reflist}}
 
==References==
 
* El'Yasberg "Theory of flight of artificial earth satellites", '''Israel program for Scientific Translations (1967)'''
*{{cite book
| last1 = Bate
| first1 = Roger
| last2 = Mueller
| first2 = Donald
| last3 = White
| first3 = Jerry
| title = Fundamentals of Astrodynamics
| publisher = Dover Publications, Inc., New York
| year = 1971
| ISBN = 0-486-60061-0}}
 
==External links==
* [http://wgpqqror.homepage.t-online.de/work.html JAVA applet animating the orbit of a satellite] in an elliptic Kepler orbit around the Earth with any value for semi-major axis and eccentricity.
 
[[Category:Orbits]]
[[Category:Johannes Kepler]]

Latest revision as of 03:29, 4 November 2013

A diagram of the various forms of the Kepler Orbit and their eccentricities. Blue is a hyperbolic trajectory (e > 1). Green is a parabolic trajectory (e = 1). Red is an elliptical orbit (e < 1). Grey is a circular orbit (e = 0).
For further closely relevant mathematical developments see also Two-body problem, also Gravitational two-body problem, and Kepler problem.

In celestial mechanics, a Kepler orbit (or Keplerian orbit) describes the motion of an orbiting body 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 most applications, there is a large central body, the center of mass of which is assumed to be the center of mass of the entire system. By decomposition, the orbits of two objects of similar mass can be described as Kepler orbits around their common center of mass, their barycenter.

Introduction

From ancient times until the 16th and 17th centuries, the motions of the planets were believed to follow perfectly circular geocentric paths as taught by the ancient Greek philosophers Aristotle and Ptolemy. Variations in the motions of the planets were explained by smaller circular paths overlaid on the larger path (see epicycle). As measurements of the planets became increasingly accurate, revisions to the theory were proposed. In 1543, Nicolaus Copernicus published a heliocentric model of the solar system, although he still believed that the planets traveled in perfectly circular paths centered on the sun.Potter or Ceramic Artist Truman Bedell from Rexton, has interests which include ceramics, best property developers in singapore developers in singapore and scrabble. Was especially enthused after visiting Alejandro de Humboldt National Park.

Johannes Kepler

In 1601, Johannes Kepler acquired the extensive, meticulous observations of the planets made by Tycho Brahe. Kepler would spend the next five years trying to fit the observations of the planet Mars to various curves. In 1609, Kepler published the first two of his three laws of planetary motion. The first law states:

"The orbit of every planet is an ellipse with the sun at a focus."

More generally, the path of an object undergoing Keplerian motion may also follow a parabola or a hyperbola, which, along with ellipses, belong to a group of curves known as conic sections. Mathematically, the distance between a central body and an orbiting body can be expressed as:

where:

Alternately, the equation can be expressed as:

Where is called the semi-latus rectum of the curve. This form of the equation is particularly useful when dealing with parabolic trajectories, for which the semi-major axis is infinite.

Despite developing these laws from observations, Kepler was never able to develop a theory to explain these motions.[1]

Isaac Newton

Between 1665 to 1666, Isaac Newton developed several concepts related to motion, gravitation and differential calculus. However, these concepts were not published until 1687 in the Principia, in which he outlined his laws of motion and his law of universal gravitation. His second of his three laws of motion states:

The acceleration a of a body is parallel and directly proportional to the net force acting on the body, is in the direction of the net force, and is inversely proportional to the mass of the body:

Where:

Strictly speaking, this form of the equation only applies to an object of constant mass, which holds true based on the simplifying assumptions made below.

The mechanisms of Newton's law of universal gravitation; a point mass m1 attracts another point mass m2 by a force F2 which is proportional to the product of the two masses and inversely proportional to the square of the distance (r) between them. Regardless of masses or distance, the magnitudes of |F1| and |F2| will always be equal. G is the gravitational constant.

Newton's law of gravitation states:

Every point mass attracts every other point mass by a force pointing along the line intersecting both points. The force is proportional to the product of the two masses and inversely proportional to the square of the distance between the point masses:

where:

From the laws of motion and the law of universal gravitation, Newton was able to derive Kepler's laws, demonstrating consistency between observation and theory. The laws of Kepler and Newton formed the basis of modern celestial mechanics until Albert Einstein introduced the concepts of special and general relativity in the early 20th century. For most applications, Keplerian motion approximates the motions of planets and satellites to relatively high degrees of accuracy and is used extensively in astronomy and astrodynamics.

Simplified two body problem

To solve for the motion of an object in a two body system, two simplifying assumptions can be made:

1. The bodies are spherically symmetric and can be treated as point masses.
2. There are no external or internal forces acting upon the bodies other than their mutual gravitation.

The shapes of large celestial bodies are close to spheres. By symmetry, the net gravitational force attracting a mass point towards a homogeneous sphere must be directed towards its centre. The shell theorem (also proven by Isaac Newton) states that the magnitude of this force is the same as if all mass was concentrated in the middle of the sphere, even if the density of the sphere varies with depth (as it does for most celestial bodies). From this immediately follows that the attraction between two homogeneous spheres is as if both had its mass concentrated to its center.

Smaller objects, like asteroids or spacecraft often have a shape strongly deviating from a sphere. But the gravitational forces produced by these irregularities are generally small compared to the gravity of the central body. The difference between an irregular shape and a perfect sphere also diminishes with distances, and most orbital distances are very large when compared with the diameter of a small orbiting body. Thus for some applications, shape irregularity can be neglected without significant impact on accuracy.

Planets rotate at varying rates and thus may take a slightly oblate shape because of the centrifugal force. With such an oblate shape, the gravitational attraction will deviate somewhat from that of a homogeneous sphere. This phenomenon is quite noticeable for artificial Earth satellites, especially those in low orbits. At larger distances the effect of this oblateness becomes negligible. Planetary motions in the Solar System can be computed with sufficient precision if they are treated as point masses.

Two point mass objects with masses and and position vectors and relative to some inertial reference frame experience gravitational forces:

where is the relative position vector of mass 1 with respect to mass 2, expressed as:

and is the unit vector in that direction and is the length of that vector.

Dividing by their respective masses and subtracting the second equation from the first yields the equation of motion for the acceleration of the first object with respect to the second:

Template:NumBlk

where is the gravitational parameter and is equal to

In many applications, a third simplifying assumption can be made:

3. When compared to the central body, the mass of the orbiting body is insignificant. Mathematically, m1 >> m2, so μ = G (m1 + m2) ≈ Gm1.

This assumption is not necessary to solve the simplified two body problem, but it simplifies calculations, particularly with Earth-orbiting satellites and planets orbiting the sun. Even Jupiter's mass is less than the sun's by a factor of 1047,[2] which would constitute an error of 0.096% in the value of μ. Notable exceptions include the Earth-moon system (mass ratio of 81.3), the Pluto-Charon system (mass ratio of 8.9) and binary star systems.

Under these assumptions the differential equation for the two body case can be completely solved mathematically and the resulting orbit which follows Kepler's laws of planetary motion is called a "Kepler orbit". The orbits of all planets are to high accuracy Kepler orbits around the Sun. The small deviations are due to the much weaker gravitational attractions between the planets, and in the case of Mercury, due to general relativity. The orbits of the artificial satellites around the Earth are, with a fair approximation, Kepler orbits with small perturbations due to the gravitational attraction of the sun, the moon and the oblateness of the Earth. In high accuracy applications for which the equation of motion must be integrated numerically with all gravitational and non-gravitational forces (such as solar radiation pressure and atmospheric drag) being taken into account, the Kepler orbit concepts are of paramount importance and heavily used.

Keplerian elements

Keplerian orbital elements.

Mining Engineer (Excluding Oil ) Truman from Alma, loves to spend time knotting, largest property developers in singapore developers in singapore and stamp collecting. Recently had a family visit to Urnes Stave Church.

It is worth mentioning that any Keplerian trajectory can be defined by six parameters. The motion of an object moving in three-dimensional space is characterized by a position vector and a velocity vector. Each vector has three components, so the total number of values needed to define a trajectory through space is six. An orbit is generally defined by six elements (known as Keplerian elements) that can be computed from position and velocity, three of which have already been discussed. These elements are convenient in that of the six, five are unchanging for an unperturbed orbit (a stark contrast to two constantly changing vectors). The future location of an object within its orbit can be predicted and its new position and velocity can be easily obtained from the orbital elements.

Two define the size and shape of the trajectory:

Three define the orientation of the orbital plane:

And finally:

Because , and are simply angular measurements defining the orientation of the trajectory in the reference frame, they are not strictly necessary when discussing the motion of the object within the orbital plane. They have been mentioned here for completeness, but are not required for the proofs below.

Mathematical solution of the differential equation (Template:EquationNote) above

For movement under any central force, i.e. a force parallel to r, the specific relative angular momentum stays constant:

Since the cross product of the position vector and its velocity stays constant, they must lie in the same plane, orthogonal to . This implies the vector function is a plane curve.
Because the equation has symmetry around its origin, it is easier to solve in polar coordinates. However, it is important to note that equation (Template:EquationNote) refers to linear acceleration , as opposed to angular or radial acceleration. Therefore, one must be cautious when transforming the equation. Introducing a cartesian coordinate system and polar unit vectors in the plane orthogonal to :



We can now rewrite the vector function and its derivatives as:




(see "Polar coordinates#Vector calculus"). Substituting these into (Template:EquationNote), we find:


This gives the non-ordinary polar differential equation:

  1. Template:NumBlk

In order to solve this equation, we must first eliminate all time derivatives. We find that:

  1. Template:NumBlk

Taking the time derivative of (Template:EquationNote), we get

  1. Template:NumBlk

Equations (Template:EquationNote) and (Template:EquationNote) allow us to eliminate the time derivatives of . In order to eliminate the time derivatives of , we must use the chain rule to find appropriate substitutions:

  1. Template:NumBlk
  1. Template:NumBlk

Using these four substitutions, all time derivatives in (Template:EquationNote) can be eliminated, yielding an ordinary differential equation for as function of .



  1. Template:NumBlk

The differential equation (Template:EquationNote) can be solved analytically by the variable substitution

  1. Template:NumBlk

Using the chain rule for differentiation one gets:

  1. Template:NumBlk
  2. Template:NumBlk

Using the expressions (Template:EquationNote) and (Template:EquationNote) for and one gets

  1. Template:NumBlk

with the general solution

  1. Template:NumBlk

where e and are constants of integration depending on the initial values for s and .

Instead of using the constant of integration explicitly one introduces the convention that the unit vectors defining the coordinate system in the orbital plane are selected such that takes the value zero and e is positive. This then means that is zero at the point where is maximal and therefore is minimal. Defining the parameter p as one has that

Alternate derivation

Another way to solve this equation without the use of polar differential equations is as follows:
Define a unit vector such that and . It follows that


Now consider


(see Triple product#Vector triple product). Notice that



Substituting these values into the previous equation, one gets:


Integrating both sides:


Where c is a constant vector. Dotting this with r yields an interesting result:


Where is the angle between and . Solving for r:


Notice that are effectively the polar coordinates of the vector function. Making the substitutions and , we again arrive at the equation

  1. Template:NumBlk

This is the equation in polar coordinates for a conic section with origin in a focal point. The argument is called "true anomaly".

Properties of trajectory equation

For this is a circle with radius p.

For this is an ellipse with

  1. Template:NumBlk
  2. Template:NumBlk

For this is a parabola with focal length

For this is a hyperbola with

  1. Template:NumBlk
  2. Template:NumBlk

The following image illustrates an ellipse (red), a parabola (green) and a hyperbola (blue)

An elliptic Kepler orbit with an eccentricity of 0.7, a parabolic Kepler orbit and a hyperbolic Kepler orbit with an eccentricity of 1.3. The distance to the focal point is a function of the polar angle relative to the horizontal line as given by the equation (Template:EquationNote)

The point on the horizontal line going out to the right from the focal point is the point with for which the distance to the focus takes the minimal value , the pericentre. For the ellipse there is also an apocentre for which the distance to the focus takes the maximal value . For the hyperbola the range for is

and for a parobola the range is

Using the chain rule for differentiation (Template:EquationNote), the equation (Template:EquationNote) and the definition of p as one gets that the radial velocity component is

  1. Template:NumBlk

and that the tangential component (velocity component perpendicular to ) is

  1. Template:NumBlk

The connection between the polar argument and time t is slightly different for elliptic and hyperbolic orbits.

For an elliptic orbit one switches to the "eccentric anomaly" E for which

  1. Template:NumBlk
  2. Template:NumBlk

and consequently

  1. Template:NumBlk
  2. Template:NumBlk

and the angular momentum H is

  1. Template:NumBlk

Integrating with respect to time t one gets

  1. Template:NumBlk

under the assumption that time is selected such that the integration constant is zero.

As by definition of p one has

  1. Template:NumBlk

this can be written

  1. Template:NumBlk

For a hyperbolic orbit one uses the hyperbolic functions for the parameterisation

  1. Template:NumBlk
  2. Template:NumBlk

for which one has

  1. Template:NumBlk
  2. Template:NumBlk

and the angular momentum H is

  1. Template:NumBlk

Integrating with respect to time t one gets

  1. Template:NumBlk

i.e.

  1. Template:NumBlk

To find what time t that corresponds to a certain true anomaly one computes corresponding parameter E connected to time with relation (Template:EquationNote) for an elliptic and with relation (Template:EquationNote) for a hyperbolic orbit.

Note that the relations (Template:EquationNote) and (Template:EquationNote) define a mapping between the ranges

Some additional formulae

See also Equation of the center – Analytical expansions

For an elliptic orbit one gets from (Template:EquationNote) and (Template:EquationNote) that

  1. Template:NumBlk

and therefore that

  1. Template:NumBlk

From (Template:EquationNote) then follows that

From the geometrical construction defining the eccentric anomaly it is clear that the vectors and are on the same side of the x-axis. From this then follows that the vectors and are in the same quadrant. One therefore has that

  1. Template:NumBlk

and that

  1. Template:NumBlk
  2. Template:NumBlk

where "" is the polar argument of the vector and n is selected such that

For the numerical computation of the standard function ATAN2(y,x) (or in double precision DATAN2(y,x)) available in for example the programming language FORTRAN can be used.

Note that this is a mapping between the ranges

For an hyperbolic orbit one gets from (Template:EquationNote) and (Template:EquationNote) that

  1. Template:NumBlk

and therefore that

  1. Template:NumBlk

As

and as and have the same sign it follows that

  1. Template:NumBlk

This relation is convenient for passing between "true anomaly" and the parameter E, the latter being connected to time through relation (Template:EquationNote). Note that this is a mapping between the ranges

and that can be computed using the relation

From relation (Template:EquationNote) follows that the orbital period P for an elliptic orbit is

  1. Template:NumBlk

As the potential energy corresponding to the force field of relation (Template:EquationNote) is

it follows from (Template:EquationNote), (Template:EquationNote), (Template:EquationNote) and (Template:EquationNote) that the sum of the kinetic and the potential energy

for an elliptic orbit is

  1. Template:NumBlk

and from (Template:EquationNote), (Template:EquationNote), (Template:EquationNote) and (Template:EquationNote) that the sum of the kinetic and the potential energy for a hyperbolic orbit is

  1. Template:NumBlk

Relative the inertial coordinate system

in the orbital plane with towards pericentre one gets from (Template:EquationNote) and (Template:EquationNote) that the velocity components are

  1. Template:NumBlk
  2. Template:NumBlk

Determination of the Kepler orbit that corresponds to a given initial state

This is the "initial value problem" for the differential equation (Template:EquationNote) which is a first order equation for the 6-dimensional "state vector" when written as

  1. Template:NumBlk
  2. Template:NumBlk

For any values for the initial "state vector" the Kepler orbit corresponding to the solution of this initial value problem can be found with the following algorithm:

Define the orthogonal unit vectors through

  1. Template:NumBlk
  2. Template:NumBlk

with and

From (Template:EquationNote), (Template:EquationNote) and (Template:EquationNote) follows that by setting

  1. Template:NumBlk

and by defining and such that

  1. Template:NumBlk
  2. Template:NumBlk

where

  1. Template:NumBlk

one gets a Kepler orbit that for true anomaly has the same r, and values as those defined by (Template:EquationNote) and (Template:EquationNote).

If this Kepler orbit then also has the same vectors for this true anomaly as the ones defined by (Template:EquationNote) and (Template:EquationNote) the state vector of the Kepler orbit takes the desired values for true anomaly .

The standard inertially fixed coordinate system in the orbital plane (with directed from the centre of the homogeneous sphere to the pericentre) defining the orientation of the conical section (ellipse, parabola or hyperbola) can then be determined with the relation

  1. Template:NumBlk
  2. Template:NumBlk

Note that the relations (Template:EquationNote) and (Template:EquationNote) has a singularity when and

i.e.

  1. Template:NumBlk

which is the case that it is a circular orbit that is fitting the initial state

The osculating Kepler orbit

Mining Engineer (Excluding Oil ) Truman from Alma, loves to spend time knotting, largest property developers in singapore developers in singapore and stamp collecting. Recently had a family visit to Urnes Stave Church. For any state vector the Kepler orbit corresponding to this state can be computed with the algorithm defined above. First the parameters are determined from and then the orthogonal unit vectors in the orbital plane using the relations (Template:EquationNote) and (Template:EquationNote).

If now the equation of motion is

  1. Template:NumBlk

where

is a function other than

the resulting parameters

defined by will all vary with time as opposed to the case of a Kepler orbit for which only the parameter will vary

The Kepler orbit computed in this way having the same "state vector" as the solution to the "equation of motion" (Template:EquationNote) at time t is said to be "osculating" at this time.

This concept is for example useful in case

where

is a small "perturbing force" due to for example a faint gravitational pull from other celestial bodies. The parameters of the osculating Kepler orbit will then only slowly change and the osculating Kepler orbit is a good approximation to the real orbit for a considerable time period before and after the time of osculation.

This concept can also be useful for a rocket during powered flight as it then tells which Kepler orbit the rocket would continue in in case the thrust is switched-off.

For a "close to circular" orbit the concept "eccentricity vector" defined as is useful. From (Template:EquationNote), (Template:EquationNote) and (Template:EquationNote) follows that

  1. Template:NumBlk

i.e. is a smooth differentiable function of the state vector also if this state corresponds to a circular orbit.

See also

Sportspersons Hyslop from Nicolet, usually spends time with pastimes for example martial arts, property developers condominium in singapore singapore and hot rods. Maintains a trip site and has lots to write about after touring Gulf of Porto: Calanche of Piana.

Citations

43 year old Petroleum Engineer Harry from Deep River, usually spends time with hobbies and interests like renting movies, property developers in singapore new condominium and vehicle racing. Constantly enjoys going to destinations like Camino Real de Tierra Adentro.

References

  • El'Yasberg "Theory of flight of artificial earth satellites", Israel program for Scientific Translations (1967)
  • 20 year-old Real Estate Agent Rusty from Saint-Paul, has hobbies and interests which includes monopoly, property developers in singapore and poker. Will soon undertake a contiki trip that may include going to the Lower Valley of the Omo.

    My blog: http://www.primaboinca.com/view_profile.php?userid=5889534

External links

  1. Bate, Mueller, White. pp 177–181
  2. http://ssd.jpl.nasa.gov