Você está na página 1de 10

Kepler orbit

philosophers Aristotle and Ptolemy. Variations in the


e=1 e=2 motions of the planets were explained by smaller circular
paths overlaid on the larger path (see epicycle). As mea-
surements 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.
e=0.5 e=0
F 1.1 Johannes Kepler
In 1601, Johannes Kepler acquired the extensive, metic-
ulous observations of the planets made by Tycho Brahe.
Kepler would spend the next five years trying to fit the ob-
servations of the planet Mars to various curves. In 1609,
Kepler published the first two of his three laws of plane-
tary motion. The first law states:

A diagram of the various forms of the Kepler Orbit and their “The orbit of every planet is an ellipse with the
eccentricities. Blue is a hyperbolic trajectory (e > 1). Green is a
sun at a focus.”
parabolic trajectory (e = 1). Red is an elliptical orbit (0 < e < 1).
Grey is a circular orbit (e = 0).
More generally, the path of an object undergoing Kep-
lerian motion may also follow a parabola or a hyperbola,
In celestial mechanics, a Kepler orbit (or Keplerian or- which, along with ellipses, belong to a group of curves
bit) describes the motion of an orbiting body as an ellipse, known as conic sections. Mathematically, the distance
parabola, or hyperbola, which forms a two-dimensional between a central body and an orbiting body can be ex-
orbital plane in three-dimensional space. (A Kepler or- pressed as:
bit can also form a straight line.) It considers only the
point-like gravitational attraction of two bodies, neglect-
ing perturbations due to gravitational interactions with a(1 − e2 )
other objects, atmospheric drag, solar radiation pressure, r(ν) = 1 + e cos(ν)
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, where:
known as the Kepler problem. As a theory in classical
mechanics, it also does not take into account the effects of • r is the distance
general relativity. Keplerian orbits can be parametrized
• a is the semi-major axis, which defines the size of
into six orbital elements in various ways.
the orbit
In most applications, there is a large central body, the cen-
ter of mass of which is assumed to be the center of mass • e is the eccentricity, which defines the shape of the
of the entire system. By decomposition, the orbits of two orbit
objects of similar mass can be described as Kepler orbits • ν is the true anomaly, which is the angle between the
around their common center of mass, their barycenter. current position of the orbiting object and the loca-
tion in the orbit at which it is closest to the central
body (called the periapsis)
1 Introduction
Alternately, the equation can be expressed as:
From ancient times until the 16th and 17th centuries, the
motions of the planets were believed to follow perfectly p
circular geocentric paths as taught by the ancient Greek r(ν) = 1 + e cos(ν)

1
2 2 SIMPLIFIED TWO BODY PROBLEM

Where p 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 F1 F2
is infinite. m1 m2
Despite developing these laws from observations, Ke-
pler was never able to develop a theory to explain these
motions.[1] r
1.2 Isaac Newton m
_1× m2
F1 = F2 = G r2
Between 1665 to 1666, Isaac Newton developed several
concepts related to motion, gravitation and differential
calculus. However, these concepts were not published un- The mechanisms of Newton’s law of universal gravitation; a point
til 1687 in the Principia, in which he outlined his laws of mass m1 attracts another point mass m2 by a force F2 which is
motion and his law of universal gravitation. His second proportional to the product of the two masses and inversely pro-
of his three laws of motion states: portional to the square of the distance (r) between them. Regard-
less of masses or distance, the magnitudes of |F1 | and |F2 | will
always be equal. G is the gravitational constant.
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, • r is the distance between the two point
and is inversely proportional to the mass of the masses
body:
From the laws of motion and the law of universal gravi-
d2 r tation, Newton was able to derive Kepler’s laws, demon-
F = ma = m 2
dt strating consistency between observation and theory. The
laws of Kepler and Newton formed the basis of mod-
Where:
ern celestial mechanics until Albert Einstein introduced
• F is the force vector the concepts of special and general relativity in the early
• m is the mass of the body on which the 20th century. For most applications, Keplerian motion
force is acting approximates the motions of planets and satellites to rel-
atively high degrees of accuracy and is used extensively
• a is the acceleration vector, the second in astronomy and astrodynamics.
time derivative of the position vector r

Strictly speaking, this form of the equation only applies


to an object of constant mass, which holds true based on 2 Simplified two body problem
the simplifying assumptions made below.
Newton’s law of gravitation states: (See also Orbit Analysis)
To solve for the motion of an object in a two body system,
Every point mass attracts every other point two simplifying assumptions can be made:
mass by a force pointing along the line inter-
secting both points. The force is proportional 1. The bodies are spherically symmetric and
to the product of the two masses and inversely can be treated as point masses.
proportional to the square of the distance be-
tween the point masses: 2. There are no external or internal forces act-
ing upon the bodies other than their mutual
gravitation.
m1 m2
F =G
r2
The shapes of large celestial bodies are close to spheres.
where: By symmetry, the net gravitational force attracting a mass
• F is the magnitude of the gravitational point towards a homogeneous sphere must be directed
force between the two point masses towards its centre. The shell theorem (also proven by
• G is the gravitational constant Isaac Newton) states that the magnitude of this force is
the same as if all mass was concentrated in the middle of
• m1 is the mass of the first point mass the sphere, even if the density of the sphere varies with
• m2 is the mass of the second point mass depth (as it does for most celestial bodies). From this
2.1 Keplerian elements 3

immediately follows that the attraction between two ho- This assumption is not necessary to solve the simplified
mogeneous spheres is as if both had its mass concentrated two body problem, but it simplifies calculations, particu-
to its center. larly with Earth-orbiting satellites and planets orbiting the
Smaller objects, like asteroids or spacecraft often have sun. Even[2]Jupiter's mass is less than the sun’s by a factor
a shape strongly deviating from a sphere. But the grav- of 1047, which would constitute an error of 0.096%
itational forces produced by these irregularities are gen- in the value of μ. Notable exceptions include the Earth-
erally small compared to the gravity of the central body. moon system (mass ratio of 81.3), the Pluto-Charon sys-
The difference between an irregular shape and a perfect tem (mass ratio of 8.9) and binary star systems.
sphere also diminishes with distances, and most orbital Under these assumptions the differential equation for the
distances are very large when compared with the diame- two body case can be completely solved mathematically
ter of a small orbiting body. Thus for some applications, and the resulting orbit which follows Kepler’s laws of
shape irregularity can be neglected without significant planetary motion is called a “Kepler orbit”. The orbits of
impact on accuracy. all planets are to high accuracy Kepler orbits around the
Planets rotate at varying rates and thus may take a slightly Sun. The small deviations are due to the much weaker
oblate shape because of the centrifugal force. With such gravitational attractions between the planets, and in the
an oblate shape, the gravitational attraction will deviate case of Mercury, due to general relativity. The orbits of
somewhat from that of a homogeneous sphere. This phe- the artificial satellites around the Earth are, with a fair ap-
nomenon is quite noticeable for artificial Earth satellites, proximation, Kepler orbits with small perturbations due
especially those in low orbits. At larger distances the ef- to the gravitational attraction of the sun, the moon and the
fect of this oblateness becomes negligible. Planetary mo- oblateness of the Earth. In high accuracy applications for
tions in the Solar System can be computed with sufficient which the equation of motion must be integrated numer-
precision if they are treated as point masses. ically with all gravitational and non-gravitational forces
(such as solar radiation pressure and atmospheric drag)
Two point mass objects with masses m1 and m2 and po- being taken into account, the Kepler orbit concepts are
sition vectors r1 and r2 relative to some inertial reference of paramount importance and heavily used.
frame experience gravitational forces:

−Gm1 m2 2.1 Keplerian elements


m1 r̈1 = ^r
r2
Gm1 m2 Celestial body
m2 r̈2 = ^
r
r2
where r is the relative position vector of mass 1 with re-
spect to mass 2, expressed as:

True anomaly
ν
r = r1 − r2 Argument of periapsis
ω
and ^
r is the unit vector in that direction and r is the length Ω

of that vector.
Longitude of ascending node Reference
Dividing by their respective masses and subtracting the direction

second equation from the first yields the equation of mo- Plane o
f refere
tion for the acceleration of the first object with respect to nce i
Inclination
the second: ☊
Ascending node
it
Orb

where µ is the gravitational parameter and is equal to Keplerian orbital elements.

Main article: Keplerian elements


µ = G(m1 + m2 )
In many applications, a third simplifying assumption can It is worth mentioning that any Keplerian trajectory can
be made: be defined by six parameters. The motion of an object
moving in three-dimensional space is characterized by a
3. When compared to the central body, the position vector and a velocity vector. Each vector has
mass of the orbiting body is insignificant. three components, so the total number of values needed
Mathematically, m1 >> m2 , so μ = G (m1 + to define a trajectory through space is six. An orbit is
m2 ) ≈ Gm1 . generally defined by six elements (known as Keplerian el-
4 3 MATHEMATICAL SOLUTION OF THE DIFFERENTIAL EQUATION (1) ABOVE

ements) that can be computed from position and veloc- it is easier to solve in polar coordinates. However, it
ity, three of which have already been discussed. These is important to note that equation (1) refers
( ) to linear
elements are convenient in that of the six, five are un- acceleration (r̈) , as opposed to angular θ̈ or radial
changing for an unperturbed orbit (a stark contrast to two
(r̈) acceleration. Therefore, one must be cautious when
constantly changing vectors). The future location of an
transforming the equation. Introducing a cartesian
object within its orbit can be predicted and its new posi-
coordinate system (x̂ , ŷ) and polar unit vectors (r̂ , θ̂)
tion and velocity can be easily obtained from the orbital
in the plane orthogonal to H :
elements.
Two define the size and shape of the trajectory:
r̂ = cos(θ)x̂ + sin(θ)ŷ
θ̂ = − sin(θ)x̂ + cos(θ)ŷ
• Semimajor axis ( a )

• Eccentricity ( e ) We can now rewrite the vector function r and its deriva-
tives as:
Three define the orientation of the orbital plane:
r = r(cos θx̂ + sin θŷ) = rr̂
• Inclination ( i ) defines the angle between the orbital
plane and the reference plane. ṙ = ṙr̂ + rθ̇θ̂
• Longitude of the ascending node ( Ω ) defines the an-
gle between the reference direction and the upward r̈ = (r̈ − rθ̇2 )r̂ + (rθ̈ + 2ṙθ̇)θ̂
crossing of the orbit on the reference plane (the as-
cending node).
(see "Vector calculus"). Substituting these into (1), we
• Argument of periapsis ( ω ) defines the angle be- find:
tween the ascending node and the periapsis.
( )
(r̈ − rθ̇2 )r̂ + (rθ̈ + 2ṙθ̇)θ̂ = − rµ2 r̂ + (0)θ̂
And finally:
This gives the non-ordinary polar differential equation:
• True anomaly ( ν ) defines the position of the orbit-
ing body along the trajectory, measured from peri- 1.
apsis. Several alternate values can be used instead of
true anomaly, the most common being M the mean In order to solve this equation, we must first eliminate all
anomaly and T , the time since periapsis. time derivatives. We find that:
H = |r × ṙ| = |(r cos(θ), r sin(θ), 0) × (ṙ cos(θ) −
Because i , Ω and ω are simply angular measurements
r sin(θ)θ̇, ṙ sin(θ) + r cos(θ)θ̇, 0)| = |(0, 0, r2 θ̇)| = r2 θ̇
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 1.
been mentioned here for completeness, but are not re-
quired for the proofs below. Taking the time derivative of (3), we get

1.
3 Mathematical solution of the dif-
Equations (3) and (4) allow us to eliminate the time
ferential equation (1) above derivatives of θ . In order to eliminate the time deriva-
tives of r , we must use the chain rule to find appropriate
For movement under any central force, i.e. a force substitutions:
parallel to r, the specific relative angular momentum
H = r × ṙ stays constant: 1.
d
Ḣ = dt (r × ṙ) = ṙ × ṙ + r × r̈ = 0 + 0 = 0
1.
Since the cross product of the position vector and its
velocity stays constant, they must lie in the same plane, Using these four substitutions, all time derivatives in
orthogonal to H . This implies the vector function is a (2) can be eliminated, yielding an ordinary differential
plane curve. equation for r as function of θ .
Because the equation has symmetry around its origin,
3.2 Properties of trajectory equation 5

r̈ − rθ̇2 = − rµ2 (see Vector triple product). Notice that

d2 r
dθ2 · θ̇2 + dr
dθ · θ̈ − rθ̇2 = − rµ2 u · u = |u|2 = 1

d2 r
( H )2 ( ) ( )2
dθ2 · r2 + dr
dθ · − 2·H·
r3

− r rH2 = − rµ2 u · u̇ = 12 (u · u̇ + u̇ · u) = 1 d
2 dt (u · u) = 0

1. Substituting these values into the previous equation, one


gets:
The differential equation (7) can be solved analytically by
the variable substitution
r̈ × H = µu̇
1.
Integrating both sides:
Using the chain rule for differentiation one gets:
ṙ × H = µu + c
1.

1. Where c is a constant vector. Dotting this with r yields


an interesting result:
d2 r dr
Using the expressions (10) and (9) for dθ2 and dθ one
gets r · (ṙ × H) = r · (µu + c) = µr · u + r · c =
µr(u · u) + rc cos(θ) = r(µ + c cos(θ))
1.
Where θ is the angle between r̄ and c̄ . Solving for r:
with the general solution
r·(ṙ×H) (r×ṙ)·H |H|2
r= = =
1. µ+c cos(θ) µ+c cos(θ) µ+c cos(θ)

where e and θ0 are constants of integration depending on Notice that (r, θ) are effectively the polar coordinates of
the vector function. Making the substitutions p = |H|
2
ds
the initial values for s and dθ . µ
c
Instead of using the constant of integration θ explicitly and e = µ , we again arrive at the equation
0
one introduces the convention that the unit vectors x̂ , ŷ
defining the coordinate system in the orbital plane are se- 1.
lected such that θ0 takes the value zero and e is positive.
This then means that θ is zero at the point where s is This is the equation in polar coordinates for a conic
maximal and therefore r = 1s is minimal. Defining the section with origin in a focal point. The argument θ is
2
parameter p as Hµ one has that called “true anomaly”.
1 p
r = s = 1+e·cos θ

3.1 Alternate derivation 3.2 Properties of trajectory equation

Another way to solve this equation without the use of For e = 0 this is a circle with radius p.
polar differential equations is as follows: For 0 < e < 1 this is an ellipse with
Define a unit vector u such that r = ru and r̈ = − rµ2 u .
It follows that 1.

H = r × ṙ = ru × dt d
(ru) = ru × (ru̇ + ṙu) = 1.
r (u × u̇) + rṙ(u × u) = r2 u × u̇
2
p
For e = 1 this is a parabola with focal length 2
Now consider For e > 1 this is a hyperbola with

r̈ × H = − rµ2 u × (r2 u × u̇) = −µu × (u × u̇) = 1.


−µ[(u · u̇)u − (u · u)u̇]
1.
6 3 MATHEMATICAL SOLUTION OF THE DIFFERENTIAL EQUATION (1) ABOVE

The following image illustrates an ellipse (red), a parabola For an elliptic orbit one switches to the "eccentric
(green) and a hyperbola (blue) anomaly" E for which

1.

1.

and consequently

1.

Theta
1.
Focal point

and the angular momentum H is

1.

Integrating with respect to time t one gets

1.

under the assumption that time t = 0 is selected such that


the integration constant is zero.
An elliptic Kepler orbit with an eccentricity of 0.7, a parabolic
As by definition of p one has
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 (13) 1.

The point on the horizontal line going out to the right from this can be written
the focal point is the point with θ = 0 for which the
p
distance to the focus takes the minimal value 1+e , the 1.
pericentre. For the ellipse there is also an apocentre for
which the distance to the focus takes the maximal value
p For a hyperbolic orbit one uses the hyperbolic functions
1−e . For the hyperbola the range for θ is
for the parameterisation
[ ( ) ( )]
1 1 1.
− cos−1 − < θ < cos−1 −
e e
1.
and for a parabola the range is

for which one has


[−π < θ < π]
1.
Using the chain rule for differentiation (5), the equation
2
(2) and the definition of p as Hµ one gets that the radial 1.
velocity component is
and the angular momentum H is
1.
1.
and that the tangential component (velocity component
perpendicular to Vr ) is
Integrating with respect to time t one gets
1.
1.
The connection between the polar argument θ and time t
is slightly different for elliptic and hyperbolic orbits. i.e.
7

1. 1.

To find what time t that corresponds to a certain true and therefore that
anomaly θ one computes corresponding parameter E
connected to time with relation (27) for an elliptic and 1.
with relation (34) for a hyperbolic orbit.
Note that the relations (27) and (34) define a mapping As
between the ranges

θ 1 − cos θ 1− e−cosh E
e·cosh E−1 e · cosh E − e + cosh E e+1
[−∞ < t < ∞] ←→ [−∞ < E < ∞] tan2 = = = =
2 1 + cos θ 1+ e−cosh E
e·cosh E−1
e · cosh E + e − cosh E e−1

and as tan θ2 and tanh E2 have the same sign it follows that
4 Some additional formulae
1.
For an elliptic orbit one gets from (20) and (21) that
This relation is convenient for passing between “true
1. anomaly” and the parameter E, the latter being connected
to time through relation (34). Note that this is a mapping
and therefore that between the ranges

1. [ ( ) ( )]
−1 1 −1 1
− cos − < θ < cos − ←→ [−∞ < E < ∞]
From (36) then follows that e e
E
and that 2 can be computed using the relation
cos E−e
θ 1 − cos θ 1− 1−e·cos E 1 − e · cos E − cos E + e 1 + e 1 − cos E 1+e E
tan2 = = = = ( · ) = · tan2
2 1 + cos θ 1+ cos E−e
1−e·cos E
1 − e · cos E + cos E − e 11 − e 11++xcos E 1−e 2
tanh−1 x = ln
From the geometrical construction defining the eccentric 2 1−x
anomaly it is clear that the vectors ( cos E , sin E ) and
From relation (27) follows that the orbital period P for an
( cos θ , sin θ ) are on the same side( of the x-axis. )From
elliptic orbit is
this then follows that the vectors cos E2 , sin E2 and
( )
cos θ2 , sin θ2 are in the same quadrant. One therefore
1.
has that

1. As the potential energy corresponding to the force field


of relation (1) is
and that
µ
1. −
r
it follows from (13), (14), (18) and (19) that the sum of
1.
the kinetic and the potential energy
where " arg(x , y) " is the polar argument of the vector
( x , y ) and n is selected such that |E − θ| < π
Vr 2 + Vt 2 µ
For the numerical computation of arg(x , y) the −
2 r
standard function ATAN2(y,x) (or in double precision
DATAN2(y,x)) available in for example the program- for an elliptic orbit is
ming language FORTRAN can be used.
1.
Note that this is a mapping between the ranges
and from (13), (16), (18) and (19) that the sum of the
kinetic and the potential energy for a hyperbolic orbit is
[−∞ < θ < ∞] ←→ [−∞ < E < ∞]

For a hyperbolic orbit one gets from (28) and (29) that 1.
8 6 THE OSCULATING KEPLER ORBIT

Relative the inertial coordinate system one gets a Kepler orbit that for true anomaly θ has the
same r, Vr and Vt values as those defined by (50) and
(51).
x̂ , ŷ If this Kepler orbit then also has the same (r̂ , t̂) vectors
for this true anomaly θ as the ones defined by (50) and
in the orbital plane with x̂ towards pericentre one gets (51) the state vector (r̄ , v̄) of the Kepler orbit takes the
from (18) and (19) that the velocity components are desired values ( r¯0 , v¯0 ) for true anomaly θ .

1. The standard inertially fixed coordinate system (x̂ , ŷ) in


the orbital plane (with x̂ directed from the centre of the
homogeneous sphere to the pericentre) defining the ori-
1. entation of the conical section (ellipse, parabola or hy-
perbola) can then be determined with the relation
See also Equation of the center – Analytical expansions
The Equation of the center relates mean anomaly to true 1.
anomaly for elliptical orbits, for small numerical eccen-
tricity.
1.

5 Determination of the Kepler or- Note that the relations (53) and (54) has a singularity
when Vr = 0 and
bit that corresponds to a given
initial state √
µ

µ
Vt = V0 = = (r·Vt )2
p
This is the "initial value problem" for the differential µ
equation (1) which is a first order equation for the 6-
dimensional “state vector” ( r̄ , v̄ ) when written as i.e.

1. 1.

1. which is the case that it is a circular orbit that is fitting the


initial state ( r¯0 , v¯0 )
For any values for the initial “state vector” ( r¯0 , v¯0 ) the
Kepler orbit corresponding to the solution of this initial
value problem can be found with the following algorithm: 6 The osculating Kepler orbit
Define the orthogonal unit vectors (r̂ , t̂) through
Main article: Osculating orbit

1.
For any state vector (r̄, v̄) the Kepler orbit corresponding
to this state can be computed with the algorithm defined
1.
above. First the parameters p, e, θ are determined from
r, Vr , Vt and then the orthogonal unit vectors in the orbital
with r > 0 and Vt > 0 plane x̂, ŷ using the relations (56) and (57).
From (13), (18) and (19) follows that by setting If now the equation of motion is

1. 1.

and by defining e ≥ 0 and θ such that where

1.
˙ t)
F(r̄, r̄,
1.
is a function other than
where

1. −µ ·
r2
9

the resulting parameters 8 Citations


p, e, θ, x̂, ŷ
[1] Bate, Mueller, White. pp 177–181
defined by r̄, r̄˙ will all vary with time as opposed to the
case of a Kepler orbit for which only the parameter θ will [2] http://ssd.jpl.nasa.gov
vary
The Kepler orbit computed in this way having the same
“state vector” as the solution to the “equation of motion” 9 References
(59) at time t is said to be “osculating” at this time.
• El'Yasberg “Theory of flight of artificial earth satel-
This concept is for example useful in case
lites”, Israel program for Scientific Translations
(1967)

˙ t) = −µ ·
F(r̄, r̄,

˙ t)
+ f(r̄, r̄, • Bate, Roger; Mueller, Donald; White, Jerry (1971).
r2 Fundamentals of Astrodynamics. Dover Publica-
tions, Inc., New York. ISBN 0-486-60061-0.
where

˙ t)
f(r̄, r̄,
10 External links

is a small “perturbing force” due to for example a faint • JAVA applet animating the orbit of a satellite in an
gravitational pull from other celestial bodies. The param- elliptic Kepler orbit around the Earth with any value
eters of the osculating Kepler orbit will then only slowly for semi-major axis and eccentricity.
change and the osculating Kepler orbit is a good approx-
imation 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 pow-
ered 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 ē = e · x̂ is useful. From (53), (54)
and (56) follows that

1.

i.e. ē is a smooth differentiable function of the state vector


(r̄, v̄) also if this state corresponds to a circular orbit.

7 See also
• Two-body problem

• Gravitational two-body problem

• Kepler problem

• Kepler’s laws of planetary motion

• Elliptic orbit

• Hyperbolic trajectory

• Parabolic trajectory

• Radial trajectory

• Orbit modeling
10 11 TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES

11 Text and image sources, contributors, and licenses


11.1 Text
• Kepler orbit Source: https://en.wikipedia.org/wiki/Kepler_orbit?oldid=680463295 Contributors: Patrick, Michael Hardy, Giftlite, Beland,
Bender235, Woohookitty, Hairy Dude, Bo Jacoby, SmackBot, W!B:, Chris the speller, A. Pichler, CRGreathouse, Myasuda, Bob Stein -
VisiBone, Swpb, Hue White, Wassermann~enwiki, Paolo.dL, Superwj5, The Thing That Should Not Be, Pomona17, JMiscreant, UnCatBot,
Nettings, Terry0051, Addbot, Fgnievinski, Yobot, Stamcose, AnomieBOT, JackieBot, Obersachsebot, NOrbeck, FrescoBot, EmausBot,
Mmeijeri, ZéroBot, Iiar, Miroslav T, Teapeat, Ethg242, BG19bot, 2pem, ChrisGualtieri, JeffAEdmonds, Jaxcp3, MatthewJ00, Jeanhausser,
H.utkuunlu and Anonymous: 27

11.2 Images
• File:Kepler_orbits.svg Source: https://upload.wikimedia.org/wikipedia/commons/b/b7/Kepler_orbits.svg License: GFDL Contributors:
Own work Original artist: Stamcose
• File:NewtonsLawOfUniversalGravitation.svg Source: https://upload.wikimedia.org/wikipedia/commons/0/0e/
NewtonsLawOfUniversalGravitation.svg License: CC BY 3.0 Contributors: Self-made by User:Dna-Dennis Original artist:
User:Dna-Dennis
• File:Orbit1.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/eb/Orbit1.svg License: CC-BY-SA-3.0 Contributors:
Lasunncty (talk) Original artist: Lasunncty (talk).
• File:OrbitalEccentricityDemo.svg Source: https://upload.wikimedia.org/wikipedia/commons/8/89/OrbitalEccentricityDemo.svg Li-
cense: CC-BY-SA-3.0 Contributors: ScottAlanHill 600×480 (24,403 bytes) (Examples of orbital trajectories with various eccentricities.
Created by submitter.) English Wikipedia Original artist: ScottAlanHill
• File:Stylised_Lithium_Atom.svg Source: https://upload.wikimedia.org/wikipedia/commons/e/e1/Stylised_Lithium_Atom.svg License:
CC-BY-SA-3.0 Contributors: based off of Image:Stylised Lithium Atom.png by Halfdan. Original artist: SVG by Indolences. Recoloring
and ironing out some glitches done by Rainer Klute.

11.3 Content license


• Creative Commons Attribution-Share Alike 3.0

Você também pode gostar