<<

Orbital Dynamics: Formulary

Prof. Dr. D. Stoffer Department of Mathematics, ETH Zurich

1 Introduction

Newton’s law of motion: The net force on an object is equal to the mass of the object multiplied by its acceleration. (1) F(t) = ma(t) where F(t) : the net force acting on the object at time t. m : the mass of the object. x(t) : position of the object at time t, in an inertial frame. v(t) = ˙x(t) : velocity of the object at time t. a(t) = ¨x(t) : acceleration of the object at time t. Newton’s law of : The attractive force F between two bodies is proportional to the product of their masses m1 and m2, and inversely proportional to the square of the distance r between them: m m (2) F = G 1 2 . r2 The constant of proportionality, G, is the gravitational constant. G = (6.67428 ± 0.00067) × 10−11 m3 kg−1 s−2 = (6.67428 ± 0.00067) × 10−11 N m2 kg−2 . The n–body problem n point masses mi with positions Ri (with respect to an inertial frame) move under the influence of gravity. Let rij := Rj − Ri be the position of the point mass mj relative to the position of the point mass mi. The equations of motion are n X mimj (3) m R¨ = G e , (i = 1, 2, . . . , n) i i r2 ij j=1,(j6=i) ij 1 P P where eij := rij. Let R := miRi/ mi be the centre of mass. rij i i The 10 classical first integrals (conserved quantities)

(4) R(t) = C1t + C2 n X mi(Ri × R˙ i) = C3 ()(5) i=1 T + V = C4 (total energy)(6) 1 Pn ˙ 2 Pn P mimj T := miR is the kinetic energy and V := −G is a potential 2 i=1 i i=1 j6=i rij of the force field F. The components of C1, C2, C3 and C4 are the 10 classical first integrals.

1 2 The Two–Body Problem:

Consider two point masses m1 and m2 under the influence of gravity. Let R :(m1 + m2)R = m1R1 + m2R2 be the centre of mass, r := R2 − R1 be the relative position of m2 with respect to m1. Then µ µ (7) ¨r = − e = − r r2 r r3

with µ := G(m1 + m2). The angular momentum

(8) h := r × r˙ = constant

is the specific angular momentum (angular momentum per mass). It is a constant of motion. Let v = r˙ = vrer + v⊥e⊥ then h := rv⊥eh where eh is the unit vector in the direction of h. The equation

h2 1 p (9) r = = µ 1 + e cos ϑ 1 + e cos ϑ where ϑ is the , p is the semilatus rectum and e is the excentricity of the conic. e ∈ [0, 1) : (e = 0: circle) e = 1 : parabola e > 1 : hyperbola

The energy integral More precisely: the specific energy, energy per mass. 1 2 Kinetic energy: 2 v µ Potential energy: − r According to (6) the (specific) energy 1 µ (10) E = v2 − = constant 2 r is constant along orbits of the two–body problem. Equation (10) relates the velocity to the radius along an orbit. Computing the energy at the periapsis yields 1 µ µ (11) E = v2 − = − (1 − e2) 2 r 2p Formulae for the radial and the perpendicular components of the velocity h (12) v = e sin ϑ r p h v = (1 + e cos ϑ)(13) ⊥ p

2 Circular orbits (e = 0) µ E = − (energy)(14) 2r rµ v = (velocity)(15) r 2π T = √ r3/2 (period)(16) µ

Elliptic orbits (e ∈ (0, 1)) The following formulae hold. a(1 − e2) r = (radius)(17) 1 + e cos ϑ 1 µ µ E = v2 − = − (energy)(18) 2 r 2a

Kepler’s laws of planetary motion 1. The orbit of every planet is an ellipse with the sun at one of the foci. Thus, Kepler rejected the ancient Aristotelean, Ptolemaic and Copernican belief in circular motion. 2. A line joining a planet and the sun sweeps out equal areas during equal intervals of time as the planet travels along its orbit. This means that the planet travels faster while close to the sun and slows down when it is farther from the sun. With his law, Kepler destroyed the Aristotelean astronomical theory that planets have uniform velocity. 3. The squares of the orbital periods of planets are directly proportional to the cubes of the semi–major axes of their orbits. This means not only that larger orbits have longer periods, but also that the of a planet in a larger orbit is lower than in a smaller orbit. More precisely: 4π2 (19) T 2 = a3 µ

or, for the sun, planet1, planet2 of mass ms, m1 and m2: T 2 m + m a 3 (20) 1 s 1 = 1 T2 ms + m2 a2

Parabolic orbits (e = 1) 1 µ E = v2 − = 0 (energy)(21) 2 r r2µ v = ()(22) esc r The following holds: If v < vesc then the orbit is an ellipse (circles included).

3 If v = vesc then the orbit is a parabola. If v > vesc then the orbit is a hyperbola. Hyperbolic orbits (e > 1) The semi–major axis is negative! 1 µ µ E = v2 − = > 0 (energy)(23) 2 r −2a 2 1 v2 = µ −  (vis–viva equation)(24) r a 3 The Two–Body Problem: Position as a Function of Time

3.1 Elliptic orbits ϑ : True anomaly E : Excentric anomaly M : The Mean anomaly is the rescaled time; the period T is rescaled to 2π; passage through pericentre at time t0 corresponds to M = 0. Kepler’s equation: ( relationship between excentric and mean anomaly)

(25) E − e sin E = M.

Relationship between true and excentric anomaly

ϑ r1 + e E (26) tan = tan . 2 1 − e 2

3.2 Hyperbolic orbits

h Again, ϑ, E and M := ab (t − t0) are the true anomaly, the excentric anomaly and the mean anomaly. Kepler’s equation for hyperbolic orbits:

(27) e sinh E − E = M.

Relationship between true and excentric anomaly

ϑ re + 1 E (28) tan = tanh . 2 e − 1 2

3.3 The orbit in space Consider an inertial system with x, y, z–coordinates. For arbitrary initial conditions r(t0) = r0 6= 0 and r˙(t0) = v0 6= 0 there exists a unique solution of (7) µ (29) ¨r = − r r3

4 • The vectors r0 and v0 are not very descriptive. The orbit is easy to describe in a (ξ, η, ζ)–frame with periapsis on the ξ–axis and the ξ, η–plane containing the orbit. Three parameters are needed to describe the position of the (ξ, η, ζ)–frame with respect to the (x, y, z)–frame, for instance the tree Euler angles. Ω: longitude of the ascending node ( (e , e ) where e = 1 r ). ∠ x k k rasc asc i: inclination (∠(ez, eζ ) = ∠(ez, eh)). ω: argument of the pericentre (∠(ek, ep)). • To describe the (with periapsis on the positive ξ–axis) two param- eters are needed. e: describes the shape of the orbit. p: describes the size of the orbit. Alternatively, e and a could be used.

• To describe a point on the orbit one parameter is needed. ϑ: the true anomaly. Alternatively, the excentric anomaly E, the mean anomaly M or the elapsed time t − t0 since passage through periapsis may be used. Orbit elements The parameters Ω, i, ω, p, e, ϑ are called elements of the orbit. They may be determined as follows from given r and v.

1. Inclination i from he , hi h (30) cos i = z = z h h where h = r × v. If i < π/2 = 90o the orbit is prograde, if i > π/2 the orbit is retrograde.

2. Define k := ez × h. k points to the direction of the ascending node. If     hx −hy (31) h =  hy  , then k =  hx  . hz 0

Ω = ∠(ex, k) may be determined from he , ki −h (32) cos Ω = x = y p 2 2 k hx + hy

If hx > 0, i.e., ky > 0, then Ω ∈ (0, π). If hx < 0, i.e., ky < 0, then Ω ∈ (π, 2π). 3. From the vis–viva equation (24) r (33) a = . v2r 2 − µ

5 4. From v2 1 1 (34) e = − r − hr, viv . µ r µ determin e = |e|. If necessary p = a(1 − e2).

5. ω = ∠(e, k) may be determined from he, ki (35) cos ω = . e k

If e3 > 0 then ω ∈ (0, π)(e3 being the third component of e). If e3 < 0 then ω ∈ (π, 2π).

6. ϑ = ∠(e, r) may be determined from he, ri (36) cos ϑ = . e r If the distance to the pericentre is • increasing, i.e., if hr, vi > 0 then ϑ ∈ (0, π) • decreasing, i.e., if hr, vi < 0 then ϑ ∈ (−π, 0) or ϑ ∈ (π, 2π).

4 Rocket dynamics

There are a lot of different rocket engines, usually categorised as either high– or low– thrust engines. High–thrust engines can provide thrust accelerations significantly larger than the local gravitational acceleration, while for low–thrust engines the thrust accel- eration is much smaller than the local gravitational acceleration. To provide thrust, mass is expelled out of the rocket nozzle. Thus the rocket mass is decreasing.

4.1 The thrust The thrust of a rocket is (37) S = −m˙ c wherem ˙ describes the loss of mass (it is negative) and where (p − p )A (38) c = v + e a e −m˙ is the effective exhaust velocity. ve is the velocity of the expelled particles relative to the rocket. pe is the pressure of the exhaust at the nozzle exit, pa is the outside ambient pressure (atmospheric pressure, which has value 0 in vacuum), and A is the nozzle exit area. Assumptions: Rocket in force–free space, c is constant, one–dimensional motion. (39) mv˙ = −m˙ c or, integrating from t0 to t1 m0 m1 − ∆v (40) v1 − v0 = c log , = e c . m1 m0 This equation (in either form) is referred to as the rocket equation.

6 4.2 The equations of motion Let be γ : the flight path angle ϕ, r : the coordinates v : the tangential velocity of the rocket ak : the tangential acceleration of the rocket a⊥ : the normal acceleration of the rocket ρ : the radius of curvature u = ∠(S, v) : controll variable Then v2 (41) a = −v γ˙ + cos γ ⊥ r (42) ak =v ˙

From these equations one derives the equations of motion of a rocket in a central 1 2 gravitational field. Thrust S, atmospheric drag R = 2 σv Acw (σ: density, A: cross sectional area, cw: drag coefficient). S R µ v˙ = cos u − − g sin γ, (g = )(43) m m r2 v2 S (44) v γ˙ = −(g − ) cos γ + sin u r m v (45) ϕ˙ = cos γ r (46) r˙ = v sin γ

• There is no lifting force as for aircrafts • u is a controll variable in order to inject the rocket into the desired orbit. u = 0 corresponds to the motion when S k v.

4.3 Injection into orbit

Velocity in a circular LEO with altitude of 300km: vr = 7.728 ... km/s From (43) one gets Z τ S Z τ R Z τ (47) v = dt − dt − g sin γdt 0 m 0 m 0 | {z } | {z } | {z } idealer Antriebsbedarf Widerstandsverlust gravity loss

The total loss for injection into a LEO is about 14% (≈ 4% air drag, ≈ 10% gravity loss). The required ∆v is 9.2 – 9.3 km/s

4.4 Multistage rockets Let

(48) m0 = mt + ms + mL

7 be the total mass of a rocket at the start. mt is the mass of propellant, ms the structural mass and mL the payload mass. From the at burnout m m (49) Z = 0 = 0 ms + mL m0 − mt one immediately gets the characteristic velocity, cf. (40) (50) ∆v = c log Z.

Moreover, define the structural coefficient σ = ms and the payload ratio ν = mt+ms mL = mL . mt+ms m0−mL Optimal staging Let mti, msi, mLi, respectively, be the propellant mass, the structural mass, the payload mass of the i-th stage, respectively, i = 1, . . . , n. Note that the total mass m0i = mti + msi + mLi of the i stage is the payload mass of the (i − 1)-th stage. Define mi = mti + msi to be the sum of the propellant mass and the structural mass of the i-th stage.

Problem: For given effective exhaust velocity ci and stuctural coefficient σi of the P i-th stage, given payload mass mL and given ∆vtot := i ∆vi, i = 1, . . . , n, minimise Pn M := i=1 mi. Solution: Solve n X ci + 1/λ (51) ∆v − c log = 0 tot i c σ i=1 i i for λ, then determine

ci + 1/λ (52) Zi = ciσi and n M + mL Y (1 − σi)Zi (53) = m 1 − σ Z L i=1 i i 5 Orbital Manoeuvres Part A: Impulsive Orbit Transfer The limiting case of finite characteristic velocities ∆v during a short time ∆t → 0 is considered. At times tk, k = 1, 2,... the vehicle undergoes velocity changes ∆vk = + − ∆vk − ∆vk

5.1 Hohmann transfer The Hohman transfer is the minimum–fuel two–impulse transfer between circular or- bits. From the vis-viva equation (24) one gets for 0 < r1 < r2 √ r 2 2 r 1  r 1 r 2 2  (54) ∆vtot = µ − − + − − r1 r1 + r2 r1 r2 r2 r1 + r2

8 5.2 Bi–elliptic transfer

For 1 < α < β the bi–elliptic transfer from a circular orbit of radius r1 over a point B with rB = βr1 to a of radius r2 = αr1 the total characteristic velocity satisfies s ∆v 2 r  1 1  1 (55) bi = (β − 1) + 2 + − √ − 1 v1 β(β + 1) α β α

5.3 Change of the orbit plane One-impuls changes of the orbit plane of circular orbits are very costly. ∆v (56) = 2 sin(ϕ/2) vcircle For large changes of the orbit plane bi–elliptic transfers are more efficient.

5.4 Rendezvous Synodic period for two objects on coplanar circular orbits 2π 1 (57) S = = 1 1 n1 − n2 − T1 T2

with angular velocities n1, n2. Initial phase angle for rendezvous with Hohmann transfer:  1 + r /r 3/2 (58) β = π 1 − 1 2 2 Part B: Low Thrust Manoeuvres The equations of motion are d d2 S (59) v = x = + g dt dt2 m where S is the thrust and g is the gravitational acceleration. For almost circular orbits one gets by integrating r µ rµ (60) ∆vlow thrust = − = vcircle(r0) − vcircle(r) r0 r i.e., ∆v is equal to the difference of the orbital velocities on the circles.

6 Interplanetary Mission Analysis

6.1 Domain of influence of a planet Inspect the three–body problem spacecraft–sun–planet. Considering this problem as a perturbed vehicle–planet–two–body problem one gets G(m + m ) r r  ¨ p v sv sp (61) rpv = − 3 rpv − Gms 3 − 3 rpv rsv rsp | {z } | {z } Ap Ss

9 where Ap is the acceleration due to the planet and Ss is the due to the sun. For short,

(62) ¨rpv − Ap = Ss. Analogously,

(63) ¨rsv − As = Sp. with

G(ms + mv) rpv rsp  (64) As = − 3 rsv, Sp = −Gmp 3 + 3 rsv rpv rsp According to Laplace the domain of influence of a planet is defined as the set of all points for which S S (65) p ≥ s . As Ap 2/5 The domain of influence of a planet is approximately a ball of radius (mp/ms) rsp. According to this definition the moon is well inside the domain of influence of the .

6.2 Patched conics Within the domain of influence (sphere of influence) of a planet the two–body problem vehicle-planet is considered. The exit velocity is approximatively equal to v∞. Outside of the domain of influence of the planet the two–body problem vehicle-sun is considered. The initial velocity is equal to vv = vplanet + v∞

6.3 Flyby or

Entrance into the domain of influence with v−∞, exit with v+∞. In the sun–vehicle

system this leads to ∆v = v∞ − v−∞. For the magnitude of ∆v one has

2v∞ (66) ∆v = 2 1 + v∞  rp v0 r0

where r0 is the radius of the planet, rp is the radius of the periapsis and v0 is the 2 velocity on a circular orbit of radius r0 (note r0v0 = µ).

6.4 The restricted three–body problem

The two primaries with masses m1, m2 move on circular orbits around their centre of mass. In a rotating frame with scaled distances the primaries have fixed positions (−µ2, 0, 0), (µ1, 0, 0) where µ1 = m2/(m1 + m2), µ2 = m1/(m1 + m2). The equations of motion for a test particle (or a vehicle) ∂U (67) x¨ − 2y ˙ = ∂x ∂U (68) y¨ + 2x ˙ = ∂y ∂U (69) z¨ = ∂z 10 1 2 2 µ1 µ2 2 2 2 2 2 2 2 2 where U = (x + y ) + + with r1 = (x + µ2) + y + z , r2 = (x − µ1) + y + z . 2 r1 r2 More explicitely

µ1 µ2 x¨ − 2y ˙ − x = − 3 (x + µ2) − 3 (x − µ1)(70) r1 r2 µ1 µ2  (71) y¨ + 2x ˙ − y = − 3 + 3 y r1 r2 µ1 µ2  (72) z¨ = − 3 + 3 z r1 r2 The Jacoby integral µ µ  (73) C := x2 + y2 + 2 1 + 2 − x˙ 2 − y˙2 − z˙2 r1 r2

is a constant of motion. There are 5 equilibria: the three Euler points L1 between the two primaries, L2 and L3 on√ the positive and negative x–axis and the two Lagrange points L4,5 = ((µ1 − µ2)/2, ± 3/2)

7 Perturbations

The Keplerian motion of satellites is perturbed by

• the oblateness of the earth,

• the atmospheric drag,

• the influence of the sun and the moon,

• the radiation pressure, electomagnetic forces, etc.

General assumption: The perturbation is much smaller than gravitation.

7.1 The perturbation equations The perturbation is given as an acceleration (force/mass). The perturbation

(74) F = Fξeξ + Fηeη + Fζ eζ

is given in a satellite oriented frame (eξ, eη, eζ ) with eξ = (1/r)r, eη ⊥ eξ in the orbital plane and eζ = eξ × eη. For given r(t), v(t) = r˙(t) the osculating elements a(t), e(t), i(t), Ω(t), ω(t),M(t) are the elements of the unperturbed Kepler motion corresponding to r(t) and v(t) = r˙(t). For the unperturbed two–body problem the elements are constant, for the perturbed problem the osculating elements vary slowly as time evolves. The osculating elements

11 satisfy the following differential equations. s a3 (75) a˙ = 2 F e sin ϑ + F (1 + e cos ϑ) µ(1 − e2) ξ η s a(1 − e2) (76) e˙ = F sin ϑ + F (cos ϑ + cos E) µ ξ η s a(1 − e2) 1 i˙ = F cos(ϑ + ω)(77) µ 1 + e cos ϑ ζ s a(1 − e2) 1 sin(ϑ + ω) (78) Ω˙ = F µ 1 + e cos ϑ ζ sin i s 1 a(1 − e2)h (2 + e cos ϑ + cos E) sin ϑi (79) ω˙ = − F cos ϑ + F − e µ ξ η 1 + e cos ϑ − Ω˙ cos i . R Introduce the two variables ν := ndt and χ := nt0 where n is the angular velocity of the mean anomaly M and t0 is the time of passing through periapsis. The equation for χ is

q a 2 µ (1 − e ) (80) χ˙ = F (2e − cos ϑ − e cos2 ϑ) + F (2 + e cos ϑ) sin ϑ e(1 + e cos ϑ) ξ η To determine M = ν − χ one has to integrate the equation r µ (81) ν˙ = n = . a3 It is often advantageous to take the variable u := ω + ϑ a independent variable. A lengthy transformation leads to dp 2r3Γ (82) = F du µp η de r2Γ (83) = F sin ϑ + F (cos ϑ + cos E) du µe ξ η di r3Γ (84) = cos(u)F du µp ζ dΩ r3Γ sin u (85) = F du µp sin i ζ dω r2Γh (2 + e cos ϑ + cos E) sin ϑ (86) = − F cos ϑ + F − du µe ξ η 1 + e cos ϑ e sin ui − F ζ 1 + e cos ϑ tan i where 1 (87) Γ = . r3 sin u 1 − µp tan i Fζ

12 7.2 The method of averaging Consider the differential equation

(88)x ˙ = εf(t, x)

where ε is a small parameter and where f is T –periodic with respect to t. Then the solutions of the averaged equation

(89)y ˙ = εf(y)

R T with f(y) := (1/T ) 0 f(t, y)dt satisfy (90) |x(t) − y(t)| ≤ Cε for t ∈ [0, L/ε] .

7.3 Oblateness of the earth The gravitational potential of the earth is approximated by µ µr2 (91) U = U + U = − 0 J (3 sin2 ϕ − 1)/2 0 J2 r r3 2

where UJ2 describes the influence of the oblateness of the earth. The perturbation

F = grad UJ2 is 3µr2J 1 (92) F = − 0 2  (1 − 3 sin3 i sin2 u) e + sin2 i sin u cos u e + sin i cos i sin u e  r4 2 ξ η ζ Setting Γ = 1 in (82)–(86) one gets after rescaling to the variable t

3.5r 3.5 ˙ 3r0  µ cos i r0  cos i ◦ Ω = − 3 J2 2 2 = −9.964 2 2 /24h(93) 2 a r0 (1 − e ) a (1 − e ) 3r 3.5r µ 5 cos2 i − 1 r 3.5 5 cos2 i − 1 ˙ 0 0 ◦ (94) ω = 3 J2 2 2 = 4.982 2 2 /24h 4 a r0 (1 − e ) a (1 − e ) a˙ = 0(95) i˙ = 0(96) e˙ = 0(97)

If i < 90◦ then Ω˙ < 0, i.e., the node line drifts westward. If i > 90◦ then Ω˙ > 0, i.e., the node line advances eastward. If i < 63.4◦ or i > 116.6◦ then ω˙ > 0, meaning that the perigee advances in the direction of the satellite. If 63.4◦ < i < 116.6◦ then the perigee regresses, it moves oposite to the direction of motion.

7.4 Atmospheric drag For nearly circular LEO–orbits one has approximately 1 (98) F = 0,F = − ρc Av2/m, F = 0. ξ η 2 w ζ 13 From (75) one gets with a = r and v2 = µ/r √ (99)r ˙ = − µrρcwA/m < 0.

From (84)–(86) one gets i˙ = 0, Ω˙ = 0, ω˙ = 0, meaning that the orbit plane and the direction of the perigee remain constant. For noncircular orbits one gets

p˙ < 0(100) e˙ < 0(101) Ω˙ = 0(102) i˙ = 0(103) ω˙ = 0(104)

The orbit becomes smaller and closer to a circular orbit. The orbit plane and the direction of the perigee remain constant.

8 Attitude dynamics

An Example: the dumbbell satellite Two masses m are connected with a massless rod of length l. Positions of the masses:

x1,2 = r cos ϕ ± l cos(ϕ + ϑ)(105)

y1,2 = r sin ϕ ± l sin(ϕ + ϑ)(106) With the kinetic and the potential energy

T = m[r ˙2 + r2ϕ˙ 2 + l2(ϕ ˙ + ϑ˙)2](107)  µ µ  (108) U = −m + r1 r2 and the Lagrange function L = T − U one derives the equations of motion from the Lagrange equation d ∂L ∂L (109) − = 0 dt ∂q˙ ∂q for q = r, ϕ, ϑ. Taking the limit l → 0 one gets µ (110) r¨ − rϕ˙ 2 = − r2 d (r2ϕ˙) = 0(111) dt 3µ sin(2ϑ) (112) ϑ¨ + = −ϕ¨ 2 r3 Equations (110) and (111) are the equations for the Kepler problem in polar coordi- nates. They are decoupled from (112). Equation (112) describes the attitude dynamics of the dumbbell satellite.

14 For circular orbits (112) degenerates to the pendulum equation

¨ 3µ (113) ϑ + 3 sin(2ϑ) = 0 . 2r0 The radial equilibrium solution ϑ = 0 is stable, the tangential equilibrium solution ϑ = π/2 is unstable. To investigate (112) it is convenient to replace the time t by the excentric anomaly E. One gets √ e sin E 3 sin(2ϑ) 2e 1 − e2 sin E (114) ϑ00 − ϑ0 + = 1 − e cos ϑ 2 1 − e cos ϑ (1 − e cos ϑ)2 where 0 denotes the derivative d/dE with respect to E.

Appendix

Vector identities

a × (b × c) = ha, cib − ha, bic ha, b × ci = hc, a × bi

Astronomical constants The Sun mass = 1.989 · 1030 kg radius = 6.9599 · 105 km 11 3 2 µsun = G msun = 1.327 · 10 km /s

The Earth mass = 5.974 · 1024 kg radius = 6.37812 · 103 km 5 3 2 µearth = G mearth = 3.986 · 10 km /s mean distance from sun = 1 au = 1.495978 · 108 km

The Moon mass = 7.3483 · 1022 kg radius = 1.738 · 103 km 3 3 2 µmoon = G mmoon = 4.903 · 10 km /s mean distance from earth = 3.844 · 105 km orbit eccentricity = 0.0549 orbit inclination (to ecliptic) = 5◦ 090

15 Physical characteristics of the planets Planet Equatorial Mass Siderial Inclination of radius rotation equator to (units of Rearth) (units of Mearth) period orbit plane 0.382 0.0553 58d 16h ≈ 2◦ 0.949 0.8149 243d(retro) 177◦ 180 Earth 1, 000 1.000 23h 56m 04s 23◦ 270 0.532 0.1074 24h 37m 23s 25◦ 110 11.209 317.938 9h 50m 3◦ 070 9.49 95.181 10h 14m 26◦ 440 4.007 14.531 17h 54m 97◦ 520 3.83 17.135 19h 12m 29◦ 360 0.18 0.0022 6d 9h 18m 122◦ 460

Elements of the planetary orbits Planet semimajor axis eccentricity siderial inclination to (in au) period ecliptic plane Mercury 0.3871 0.2056 87.969d ≈ 7◦ 000 Venus 0.7233 0.0068 224.701d 3◦ 240 Earth 1.0000 0.0167 365.256d 0◦ 000 Mars 1.5237 0.0934 1y 321.73d 1◦ 510 Jupiter 5.2028 0.0483 11y 314.84d 1◦ 190 Saturn 9.5388 0.0560 29y 167d 2◦ 300 Uranus 19.1914 0.0461 84y 7.4d 0◦ 460 Neptune 30.0611 0.0097 164y 280.3d 1◦ 470 Pluto 39.5294 0.2482 247y 249d 17◦ 090

16