0% found this document useful (0 votes)
48 views27 pages

Orbital Mechanics

Uploaded by

amansahu33980
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
48 views27 pages

Orbital Mechanics

Uploaded by

amansahu33980
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

Kepler’s Laws & Orbital Geometry

Logan Reich
April 2023

Table of Contents
1 Introduction 2

2 Kepler’s First Law 2


2.1 Law and Importance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2.1 Equation of an ellipse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2.2 Using gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2.3 Other types of Orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 Kepler’s Second Law 5


3.1 Law and Importance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

4 Kepler’s Third Law 8


4.1 Law and Importance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.2.1 Circular orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.2.2 Generalized . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

5 Orbital Geometry 11
5.1 Ellipses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5.2 Parabolas and Hyperbolas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5.3 Orbital Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

6 Conclusion 14

1
USAAAO Guide Logan Reich

1 Introduction
Kepler’s laws make up the basis of astrophysics, and remain tremendously important for under-
standing all of celestial mechanics. They allow us to understand the shapes of orbits, predict
hard-to-observe parameters like distance from easy-to-observe ones like period, and to set up key
equations to understand the rate of movement!

2 Kepler’s First Law


2.1 Law and Importance
Kepler’s first law says that all planets orbit in ellipses with the sun at one focus of the ellipse,
which can be generalized to the fact that all objects in closed orbits orbit their central bodies
in ellipses with the central body at a focus of the ellipse. Open orbits are either hyperbolas or
parabolas, and also have the central body at a focus.

Kepler’s first law is important for understanding the orbital behavior of objects, as well as provid-
ing the basis for the patched conics technique, which involves combining the paths of conic orbits
to create more advanced orbital maneuvers.

Figure 1: A diagram of the closed orbital behavior of a body (Source: BBC)

2
USAAAO Guide Logan Reich

Figure 2: A diagram of the open orbital behavior of a body (Source: Wikipedia)

2.2 Derivation
2.2.1 Equation of an ellipse
Before we can prove Kepler’s First Law, we first need to understand what an elliptical path looks
like. For this, we will find the equation of an ellipse in polar coordinates. If you are not familiar
with polar coordinates, all we are doing is defining points as (r,θ) instead of (x,y) where if we draw
a radius from our point to the origin, r is the length of this radius and θ is the angle it makes with
the positive x-axis.
For our ellipse, we can set one of our foci as the origin. In this example, if the major axis is
horizontal, then I’m setting the right focus as the origin to make the following math easier. By
definition, the distance from a point on the ellipse to this focus is r. Let’s call the distance to
the other focus r′ . This means our point is at (r,θ) in polar coordinates and (r cos θ,r sin θ) in
rectangular coordinates. Since the distance between the two foci in an ellipse is 2c = 2ae, we can
use the Pythagorean theorem to find r′ :
r′2 = r2 sin2 θ + (2ae + r cos θ)2
By expanding and using the identity sin2 θ + cos2 θ = 1, it can be rewritten as
r′2 = r2 + 4ae(ae + r cos θ)
By definition, we know that for an ellipse r + r′ = 2a, so we can substitute r′ = 2a − r to solve for
r. After simplifying, we get:
a(1 − e2 )
r=
1 + e cos θ

3
USAAAO Guide Logan Reich

2.2.2 Using gravity


Warning: this section involves excessive amounts of calculus. Feel free to skip this part.

To find the path of an orbit around a central body, we begin by using a polar coordinate system
with the Sun at the origin. The only force acting on our planet is gravity (− GM
r2
), directed towards
the origin. The acceleration from this force will contribute to two components:
1. Centripetal acceleration: −ω 2 r (negative since it acts inwards) where ω is the angular
velocity
d2 r
2. The change in radius: dt2

Thus, we have the equation


GM d2 r
− = − ω2r
r2 dt2
However, ω also varies with r (due to conservation of angular momentum), so we must express it
in terms of r by using the fact that angular momentum L = µωr2 where µ is the reduced mass.

GM d2 r L 2 d2 r L2
− = − ( ) r = −
r2 dt2 µr2 dt2 µ2 r 3
Here, we see a lot of r in the denominator, so we make the sub r = u1 . However, we need to be
2
able to convert ddt2r into a usable form by abusing ω = dθ
dt
.

dr 1 du 1 du dt 1 du Lu2 L du
=− 2 = − 2 ( )(ω) = 2 ( )=−
dt u dt u dt dθ u dθ µ µ dθ

−L d2 u d dr
dt d2 r dθ 1 µ d2 r
= = ( )( ) =
µ dθ2 dθ (dt)(dθ) dt ω Lu2 dt2
d2 r L2 u2 d2 u
= −
dt2 µ2 dθ2
Therefore, our equation becomes

L 2 u 2 d2 u L 2 u 3
−GM u2 = − − 2
µ2 dθ2 µ
This leaves us with the simple differential equation:

GM µ2 d2 u
= +u
L2 dθ2
1 GM µ2
=
u= + A cos θ
r L2
This still looks ugly, but we can combine most of the constants using contrived substitutions:

( Lµ )2 a(1 − e2 )
r= =
GM (1 + e cos θ) 1 + e cos θ
This is the polar equation for an ellipse! Therefore, we have proven Kepler’s First Law.

4
USAAAO Guide Logan Reich

2.2.3 Other types of Orbits


In an orbit, there are two types of energy we mainly have to worry about. Potential energy, which
we say is negative because it keeps the object bound in the orbit, and kinetic energy, which we say
is positive because it can help an object escape orbit. An orbit is bound (elliptical) if the energy
overall is negative, which means that the gravitational potential energy is greater than the kinetic
energy. A circular orbit is the very special case where gravitational potential energy is two times
kinetic energy.

For a parabola, where the overall energy is 0 (meaning the gravitational potential perfectly
balances the kinetic energy) and the orbital eccentricity is 1,
2p
r=
1 + cos θ
For a hyperbola, which have positive orbital energies (meaning the gravitational potential is less
than the kinetic energy) and eccentricities greater than 1,

a(e2 − 1)
r=
1 + e cos θ

3 Kepler’s Second Law


3.1 Law and Importance
Kepler’s second law says that planets sweep out equal areas in equal times. This is a restatement
of the conservation of momentum, only one that Kepler came up with before the conservation of
angular momentum!

The area of an entire ellipse is abπ, so with the period, the area that is swept out over a certain
period of time can be easily calculated. This is useful for deriving various important orbital
relations, especially when used together with the conservation of angular momentum.

Figure 3: An illustration of Kepler’s second law (Source: Labster Theory)

5
USAAAO Guide Logan Reich

Example 2.1: (USAAAO Round One 2022)

Solution: From Kepler’s second law, equal areas are swept out in equal times, regardless of where
in the orbit that area is, and using ∆t as the amount of time we are looking for:
∆t Amovement
=
T Atotal
using the total area formula and the area that is swept out as half the total area πab 2
plus a
triangular area 2 × abe
2
that is formed by the semiminor axis and the lines connecting the star to
points A and B,
∆t 2 × abe
2
+ πab
2
=
T πab
∆t abe + πab
2
=
T πab

6
USAAAO Guide Logan Reich

Dividing through,
e
∆t = ( + 0.5)T
π
which is answer choice B.

3.2 Derivation
Warning: This section uses significant amounts of calculus! If you want to attempt reading this
section and you do not have a solid calculus background, please check out our Calculus Primer
handouts!

Figure 4: Diagram of a small angular area (Source: Caroll and Ostlie)


From Figure 4,
dA = dr(rdθ) = rdrdθ
Integrating from the principal focus (the focus that the central body is at) to r,
1
dA = r2 dθ
2
Taking the time derivative
dA 1 dθ
= r2
dt 2 dt
The orbital velocity can be expressed in terms of the radial and tangential velocities,
dr dθ
v = vr + vt = r̂ + r θ̂
dt t
Solving for vt and using substituting in,
dA 1
= rvt
dt 2
As vt is the tangential velocity,
L
rvt = |r × v| =
µ
thus,
dA L
=
dt 2µ

7
USAAAO Guide Logan Reich

4 Kepler’s Third Law


4.1 Law and Importance
Kepler’s third law states that ONLY in the solar system

P 2 ∝ a3

where P is the period in years and a is the semi-major axis in AU. This expression is a specific
version of Kepler’s third law, known as Newton’s version of Kepler’s third law, that only applies
to our solar system, because the proportionality constant in the solar system evaluates to 1. This
is not a coincidence; this reduction comes from the fundamental definition of our units. Newton’s
version of Kepler’s third law is
4π 2 a3
P2 =
G(M1 + M2 )
where all values are in SI units. A common way of reducing this is using the approximation that
if M1 is the mass of a star and M2 is a planet, M1 ≈ M1 + M2 . Kepler’s third law is an incredibly
important basis for many relations and formulas in orbital mechanics, and all of astronomy in
general.

Example 3.1:(USAAAO Round One 2022)

8
USAAAO Guide Logan Reich

Solution: From the graph, there is 4.2 years between the primary and secondary eclipse, which
are when the stars eclipse each other (see the section on binaries in the Stellar Evolution handout
for more info). The deeper eclipse is the primary; though this is irrelevant for the problem. Since
both stars follow circular orbits, the orbital period is simply double this time difference or 8.4
years. Using Kepler’s third law,
4π 2 (14.8AU )3
(8.4yrs)2 =
G(M1 + M2 )
using the appropriate unit conversions to SI and solving for the mass, this results in 46 solar
masses, or answer choice E.

Example 3.2: (USAAAO Round One 2020) Planet Nine is a hypothetical planet in the
outer Solar System, with a semimajor axis between 400 and 800 AU. Which of the following
is a possible orbital period for Planet Nine?
A. 71.1 years
B. 600 years
C. 1,500 years
D. 15,000 years
E. 360,000 years

Solution: Use Kepler’s third law for the solar system, where

P 2 ∝ a3

and they are equal if P is in years and a is in AU. Plugging in for 400 AU and 800 AU, the only
choice greater than the period for 400 AU and less than the period for 800 AU is 15,000 years, or
answer choice D.

Example 2.3: Does Kepler’s third law in the solar system work for Earth?

Solution: Yes, as in the solar system,


P 2 ∝ a3
and they are equal if P is in years and a is in AU. Plugging in for one year and one AU, we see
that it works and the proportionality constant must by 1!

4.2 Derivation
There are two derivations in this section: a simpler version that does not involve calculus and is
independent of Kepler’s second law, and another more rigorous one based on the second law.

4.2.1 Circular orbits


First, set centripetal force equal to the gravitational force:
v2 mM
m =G 2
r r

9
USAAAO Guide Logan Reich

GM
v2 =
r
r
GM
v=
r
In physics, velocity × time = distance. In an orbit, the distance is the circumference (2πr) and
the time is the period:
vP = 2rπ
r
GM 2rπ
v= =
r P
Squaring both sides and rearranging:
4π 2 a3
P2 =
GM
which is the version of Kepler’s third law used when the central body mass is much larger than the
orbiting body mass. Note: we substituted a in place of r since the two are equivalent in a circle.

4.2.2 Generalized
For the second derivation, starting with the integrated version of Kepler’s second law,
L
A= P

Rearranging and substituting A = abπ:

2µabπ 2 µ2 4π 2 2 2
P2 = ( ) = ab
L L2
Then using Kepler’s first law forms

L2 /µ2
r=
GM (1 + ecosθ)

a(1 − e2 )
r=
1 + ecosθ
Setting them equal to each other and solving for L
p
L = µ GM a(1 − e2 )

From the geometry of an ellipse


b2 = a2 (1 − e2 )
Plugging both of those into the equation derived from Kepler’s second law

4π 2 3 4π 2
P2 = a = a3
GM G(m1 + m2 )

which is the full, more accurate version of Kepler’s third law.

10
USAAAO Guide Logan Reich

5 Orbital Geometry
5.1 Ellipses
All closed orbits are ellipses. An ellipse is defined as the locus of points where the distances from
two foci sum to a constant.

Figure 5: Ellipse diagram (Source: Hyperphysics)

From the diagram, we can define points F1 and F2 as the two foci, b as the semi-minor axis, and
a as the semimajor axis. Alternatively, it can be defined as the points a given summed distance
from a focus and a directrix, or a specific line. The eccentricity of an ellipse, a measure of how
“circular” it is, ranges from 0 to 1 in an ellipse and can be defined as

a2 − b 2
e=
a
or through
b2 = a2 (1 − e2 )
which is often more useful. A circle has an eccentricity of 0. Most planets have extremely low
eccentricities. Eccentricity can also be expressed in terms of the focus-directrix distance. In this
ellipse, we can call the point in the orbit closest to where the central body is (F1 in this case) the
perihelion, perigee, perilune or periapsis, and the farthest point the aphelion, apogee, apolune, or
apoapsis, depending upon what body it is orbiting. -helion is the Sun, -gee is the Earth, -lune is
the Moon, and -apsis is general. An elliptical orbit can be modeled using the following equation:

(x + ea)2 y2
+ 2 =1
a2 a (1 − e2 )

The periapsis of the orbit is at distance


a(1 − e)
and the apoapsis is at distance
a(1 + e)

11
USAAAO Guide Logan Reich

A circular orbit, a special form of an elliptical orbit, can be modeled using

L2
r=
GM m2
where L is the angular momentum, G is the gravitational constant, M is the mass of the central
body, and m is the mass of the orbiting body.

Example 4.1: (USAAAO Round One 2022) The orbit of some planet to its star has an
eccentricity of 0.086. What is the ratio of the planet’s closest distance to its star to the
farthest on its orbit?
(a) 0.842
(b) 0.188
(c) 1.188
(d) 0.158
(e) None of the above

Solution: Using the formulae for the apoapsis and periapsis of an orbit, this ratio is equal to

a(1 − e)
a(1 + e)

which equals
1−e
1+e
Plugging in for the given eccentricity,
1 − 0.086
1 + 0.086
This equals 0.842, or answer choice A.

5.2 Parabolas and Hyperbolas


Parabolas and hyperbolas are the two forms of open orbits. Both can also be defined using a
focus-directrix definition, and do possess an eccentricity. A parabola has an eccentricity of 1, and
a hyperbola has an eccentricity greater than 1. Ellipses, hyperbolas, and parabolas fall into a class
of functions known as conic sections. A parabolic orbit can be modeled using

y 2 = 2p(0.5p − x)

where p is the distance of closest approach, and x and y are coordinate distances relative to the
central star the difference from the center can also be expressed using the equation

RL2
β= √
(GM m2 ) 1 − e2

Where R is the orbital distance, L is the angular momentum, G is the gravitational constant, M
is the central body’s mass, m is the orbiting body’s mass, and e is the eccentricity.

12
USAAAO Guide Logan Reich

Figure 6: A parabolic orbit (Source: Physics Stack Exchange)

Example 4.2. If an orbit has an eccentricity of 2, what shape is it, and it it open or closed?

Solution: The orbit is a hyperbola, as the eccentricity is greater than 1. Hyperbolas are open
orbits (the orbiting body never returns to any part of the orbit), so this is an open orbit.

5.3 Orbital Elements


While all orbits are conic sections, more information is needed to identify them in space relative
to Earth, and many of those parameters can help better model orbits. Beyond the two most basic
orbital parameters (the eccentricity and the semimajor axis length) the inclination i, the longitude
of the ascending node Ω, the argument of periapsis ω, and the true anomaly θ are also important.

The inclination is a measure of the tilt of the orbital plane relative to the ecliptic or another
reference plane. The longitude of the ascending node refers to the angle from the vernal equinox
(the zero right ascension, zero declination point) of the point where the orbital plane intersects
with the ecliptic, moving from above the ecliptic to below the ecliptic.

The argument of periapsis is the angle from the line of the center of the orbital plane to the
ascending node to the line along the major axis towards the direction of the periapsis. The true
anomaly is the angle from the celestial body to the segment along the major axis to the periapsis,
and is a way of parameterizing where in the orbit the body is.

13
USAAAO Guide Logan Reich

There are a series of alternatives choice of angular parameters. One of the most common is the
mean anomaly M , which is another useful way to express the position of the body in its orbit. P
is the period of the orbit. With τ as the reference time where the object is at periapsis,

M= (t − τ )
P
This equation is useful as the mean anomaly directly represents the angle for Kepler’s second law.

Figure 7: A diagram showing the orbital elements (Source: Wikipedia)

6 Conclusion
Kepler’s laws and orbital geometry are tremendously important as they underpin all of orbital
mechanics, and now you can use them too! Almost all celestial mechanics questions, especially
those in USAAAO, rely on Kepler’s laws as a crucial foundation. Understanding this material will
set you up for success!

14
Advanced Orbital Mechanics
Harsh Ambardekar
June 2023

Table of Contents
1 Introduction 2

2 Energy 2
2.1 Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Vis-Viva Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3 Virial Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

3 Orbital Maneuvers 4
3.1 Hohmann Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

4 Further Reading 6

5 Practice Problems 6

6 Conclusion 6

1
USAAAO Guide Harsh Ambardekar

1 Introduction
At this point, you should have gone through the Basic Physics and Kepler’s Laws handouts (if not,
please do those first). In this handout, we’ll focus on advanced applications of orbital mechanics,
such as how to use conservation of energy and how orbital maneuvers work.

2 Energy
2.1 Basics
The total energy of an object in an elliptical orbit (at all points) is

GM m
E=− ,
2a
where a is the semi-major axis of the orbit. The proof is trivial and left as an exercise to the reader.

Just kidding, of course! To prove this, we start by equating the energies at the perigee (clos-
est point of orbit) and apogee (farthest point of orbit):

mv12 GM m mv22 GM m
E= − = −
2 r1 2 r2

Now, let’s multiply the first equation by r12 and the second equation by r22 :

mv12 r12
Er12 = − GM mr1
2
mv22 r22
Er22 = − GM mr2
2
Subtracting these equations and using the fact that v1 r1 = v2 r2 (by conservation of angular
momentum) yields
E(r12 − r22 ) = −GM m(r1 − r2 )
Isolating for E and using the difference of squares factorization and the fact that r1 + r2 = 2a
(properties of an ellipse), we finally get

r1 − r2 −GM m −GM m
E = −GM m 2 2
= =
r1 − r2 r1 + r2 2a

Note that the total energy of an elliptical (or circular) orbit is always negative, meaning the
object can’t escape. This is why elliptical and circular orbits are called closed orbits. For an
object in a closed orbit to escape its host star, it must be given enough speed such that it has a
positive energy.

A parabolic orbit is an orbit with an eccentricity of 1. The total energy of a parabolic


orbit is 0. This means that
1 GM m
E = mv 2 − =0
2 r

2
USAAAO Guide Harsh Ambardekar

at all points in the orbit, so r


2GM
v=
r
at all points in the orbit. In other words, in a parabolic orbit, the object’s velocity is always equal
to the escape velocity; it can “just barely” escape.

Example 2.1: (2021 USAAAO First Round #15) An interesting phenomena that happens
in the Solar System is the capture of comets in the interstellar medium. Assume that a
comet with a mass of 7.15 ∗ 1016 kg is captured by the solar system. The perihelion of this
comet’s orbit after it is captured is equal to 4.64 AU, and its velocity with respect to the
Sun before being captured by the Solar System was very small. Calculate the velocity of
the comet at the perihelion.

Solution: This problem seems very daunting at first. We don’t even know what type of orbit
the comet is following. Or do we? We have to read the problem carefully. We are told the initial
velocity is very small, which means that the initial kinetic energy is 0. We also know that the
comet was ”captured” from far away, so the initial gravitational potential energy is 0. Thus, the
total initial energy, and energy at all points in the orbit, is 0, meaning the orbit is parabolic! We
can thus use conservation of energy at the perihelion:
1 GM m
E = mv 2 − =0
2 r
r
2GM
v=
r
Plug in all the numbers (making sure to convert to SI base units) to get that v =19.6 km/s □

The total energy of a hyperbolic orbit is positive. This means that the object is on a trajectory
that will escape its host star.

2.2 Vis-Viva Equation


The vis-viva equation is very common in orbital mechanics. It gives the velocity of an object at
every point in a closed orbit. It’s essentially just a statement of conservation of energy. We start
with
GM m 1 GM m
Etot = − = mv 2 −
2a 2 r
2
Now, if we solve for v , we get
2 1
v 2 = GM ( − )
r a
This is the vis-viva equation.

Example 2.2: (2022 USAAAO First Round #16) In 2025, the Parker Solar Probe will pass
just 6.9 ∗ 106 km from the Sun, becoming the closest man-made object to the Sun in history.

3
USAAAO Guide Harsh Ambardekar

It will make five orbits, passing close to the Sun once every 89 days, before the planned
end of the mission in 2026. How fast will the Parker Solar Probe be traveling at its closest
approach to the Sun?

Solution: We want the velocity, so we know we’ll need to the vis-viva equation. We know the r
value, but we don’t know the semi-major axis a. But we do have the period of orbit, so we can
use Kepler’s third law and solve for a:
T2 4π 2
=
a3 GMsun
so
GMSun T 2 1
a=( 2
) 3 = 5.8 · 107 km

Now we use the vis-viva equation to find the velocity at the closest approach:
r
2 1
v = GMsun ( − ) = 190 km/s □
r a

2.3 Virial Theorem


In the Basic Physics handout, we introduced the Virial Theorem. The theorem states that in a
circular orbit, the kinetic energy K and potential energy U are related by
1
K = − U or 2K + U = 0
2
Because we know that Etot = K + U , we can also write that
K + K + U = 0 → K + Etot = 0 → Etot = −K
The Virial Theorem can be used for spherical bodies such as stars and galaxies, in addition to
orbits.

Example 2.3: (2020 USAAAO First Round #14) As a consequence of the virial theorem,
how does the stellar temperature (T) change if we add more arbitrary energy (E) to the
star?

Solution: We use the Viriral Theorem, specifically that Etot = −K. The key thing to realize is
that Etot is negative, so when we add energy, we are making Etot more positive and thus decreasing
the magnitude of the total energy. For example, the total energy could go from -500 J to -450 J,
which means the kinetic energy would decrease from 500 J to 450 J. Because the kinetic energy
decreases, the stellar temperature would also decrease. □

3 Orbital Maneuvers
In this section, we will study two types of transfers that rockets undergo when moving between
orbits. The math here is a bit complicated, but when you’re doing these types of problems, the
most important thing to keep in mind is that you don’t need any advanced formulas; you just need
conservation of energy, the velocity in a circular orbit, and occasionally Kepler’s laws.

4
USAAAO Guide Harsh Ambardekar

3.1 Hohmann Transfer


In a Hohmann transfer, a satellite moves from one circular orbit to another circular orbit using
an elliptical transfer orbit.

Figure 1: A Hohmann transfer orbit (Source: Wikipedia)


The Hohmann transfer requires two impulses, instances where the satellite changes its velocity.
The first impulse moves the satellite from the original circular orbit to an elliptical orbit (1 to 2
in the figure). The second impulse occurs when the satellite adjusts its orbit from the elliptical
transfer orbit to the second circular orbit (2 to 3 in the diagram).

Let’s calculate these impulses. Again, impulses are essentially changes in velocity; we will
denote them
q with ∆v. The main equations we will use are the velocity of an object in a circular
GM
orbit (v = R
) and the vis-viva equation previously derived.

Let’s say the satellite goes from a circular orbit of radius r1 to a circular orbit of radius r2 . The
initial velocity of the satellite is r
GM
v1 =
r1
Then, the satellite fires its thrusters and it instantaneously (at the same point) moves into an
elliptical orbit. The semi-major axis of this elliptical orbit (yellow in the diagram) is a = r1 +r 2
2
.
You can visualize this using the diagram; the major axis is just r1 + r2 so the semi-major axis is
half of that. The distance from the satellite to the central object (in the elliptical orbit) is r1 so
we can use the vis-viva equation to get the velocity of the satellite right after it fires its thrusters
and goes into an elliptical orbit:
r s
2 2 2r2
v2 = GM ( − ) = GM ( )
r1 r1 + r2 r1 (r1 + r2 )

We can then calculate the first impulse as


r r
GM 2r2
∆v1 = v2 − v1 = ( − 1)
r1 r1 + r2

5
USAAAO Guide Harsh Ambardekar

Now, the satellite completes half of the elliptical orbit. Its velocity afterwards at the “halfway
point” can be found using the vis-viva equation again, except this time, the satellite is r2 away
from the central object:
r s
2 2 2r1
v3 = GM ( − ) = GM ( )
r2 r1 + r2 r2 (r1 + r2 )
Now, the satellite will fire its thrusters again to move to a circular orbit. It’s velocity afterwards
will just be the velocity of a satellite in a circular orbit of radius r2 :
r
GM
v4 =
r2
The second impulse is thus
r r
GM 2r1
∆v2 = v4 − v3 = (1 − )
r2 r1 + r2
The total ∆v required for a Hohmann transfer is ∆v1 + ∆v2 .

4 Further Reading
If you want to learn more about the Hohmann transfer and other orbital maneuvers, section 14.8
of Roy and Clarke Astronomy: Principles and Practice is really good. You can learn more about
transfer time and how to “coordinate” the transfer such that the satellite enters the secondary
orbit right when the secondary object is there (e.g. Earth to Mars).
Another important but less well-known maneuver is the bi-elliptic transfer orbit. It’s very
similar to a Hohmann transfer except it consists of two half-elliptical transfer orbits rather than
1. The process of calculating the ∆v’s is the exact same. You can read more about it here.

5 Practice Problems
Below are some practice problems you can try. The ones marked with a * are challenging because
they are either quite complicated or require knowledge outside of this handout.
• 2023 First Round: 10*, 12, 21*, 26*, 28
• 2022 First Round: 2, 16, 21*
• 2021 First Round: 7
• 2020 First Round: 18, 23

6 Conclusion
Orbital mechanics show up a lot on the USAAAO (both First Round and USAAAO). Memorizing
the formulas in this handout isn’t enough; you need to know how to apply them, which is why
doing practice problems is so important. But once you thoroughly understand the material, you
will be very successful on USAAAO orbital mechanics problems.

6
Basic Physics Mechanics
Harsh Ambardekar
June 2023

Table of Contents
1 Introduction 2

2 Kinematics 2
2.1 Important Quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 The Kinematics Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

3 Dynamics and Energy 3


3.1 Newton’s Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.2 Energy and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.3 Circular Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

4 Momentum and Collisions 5


4.1 Linear Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.2 Impulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.3 Center of Mass and Conservation of Momentum . . . . . . . . . . . . . . . . . . . . 6

5 Conclusion 6

6 Practice Problems 6

1
USAAAO Guide Harsh Ambardekar

1 Introduction
In order to do well in astronomy and astrophysics olympiads, one must have a solid understanding
of physics. In this handout, we will focus on basic mechanics, though you will likely want to study
other areas of physics (such as electromagnetism and thermodynamics) in the future if you want
to perform at the highest level.

2 Kinematics
We will start by studying how objects move. This study is called kinematics. When we do
kinematics problems, we’re not concerned with how or why the object started moving; we simply
want to study its future motion.

2.1 Important Quantities


Let’s begin by exploring the basic quantities in kinematics.

• Position is, quite simply, the position of an object relative to some origin. It is usually
measured in meters and is denoted by → −x (the arrow is above the x because position is a
vector quantity).

• Displacement is the change in an object’s position. It is also measured in meters and is


denoted by ∆→

x.

• Velocity is the rate at which position changes and is denoted by → −v . In other words, →
− −

v = ddtx .
It is important to note that velocity is not just a number; velocity also has a direction because
it’s a vector (e.g. 5 m/s north). Speed is the magnitude of velocity.

• Acceleration is the rate at which velocity changes: →


− d−
→v
a = dt
.

Let’s do an example to understand all these quantities.

Example 1: An object’s position is given by the function x(t) = t2 − 5t + 4.


a) What are the object’s position, velocity, and acceleration at t = 3? b) At what times is
the object moving to the right? c) At what times is the object speeding up?

Solution: The position is simply x(3) = 32 − 5 ∗ 3 + 4 = -2. The velocity function is v(t) = x′ (t) =
2t − 5, so v(3) = 2 ∗ 3 − 5 = 1. The acceleration function is a(t) = v ′ (t) = 2. Therefore, a(3) = 2.
Now, for part b), we need to determine when the object is moving to the right. For an object to
move to the right, it’s velocity must be positive. So v(t) > 0, 2t − 5 > 0, t > 2.5. The object is
thus moving to the right at all times after t = 2.5. Now, take a second and think about what
it means for an object to be speeding up. It means that the magnitude of the velocity must be
increasing, not necessarily that the velocity is becoming more positive. What this means is that
for an object to be speeding up, it’s velocity and acceleration must have the same sign. Otherwise,
it will be slowing down. For example, if the velocity is initially negative and the acceleration is
initially positive, the velocity will become more positive but its magnitude will decrease (e.g. -4

2
USAAAO Guide Harsh Ambardekar

m/s to -3 m/s), meaning the object will be slowing down. In this problem, the acceleration is
always positive, so for the object to be speeding up, the velocity must be positive. But we just
determined that that happens for t > 2.5! Therefore, we can say that the object is moving to the
right and speeding up for all times after t = 2.5.

2.2 The Kinematics Equations


There are four essential equations you must know to solve kinematics problems. Note that you
can only use these equations if the acceleration is constant.

1. vf = v0 + at. This is fairly self-explanatory. The object’s velocity at a later time (vf ) will
be equal to its initial velocity (v0 ) plus its change in velocity (at). Remember that a = dv
dt
,
so at essentially represents the change in velocity.

2. ∆x = v0 t+ 12 at2 . This equationRcan be obtained by integrating the first equation with respect
dx
R
to time (recall that v = dt , so dx = v dt.
v +v
3. ∆x = 0 2 f ∗ t. This equation essentially says that the displacement (∆x) is equal to the
v +v
average velocity ( 0 2 f ) times time. This intuitively makes sense.

4. vf2 = v02 + 2a∆x. This is the least intuitive out of all these equations. Try deriving it on your
own using the other equations. Hint: start with the definition of acceleration and then use
the third equation.

3 Dynamics and Energy


Dynamics is the study of why things move. This can be explained using forces.

3.1 Newton’s Laws


Newton’s first law states that an object at rest will remain at rest unless acted on by a net
external force. You’ve likely heard of this being associated with inertia. Inertia is the tendency
of an object to remain unchanged; if it is moving, it will want to keep moving, and if it is at rest,
it will want to stay at rest.

Newton’s second law states that Fnet = ma. Note that this equation represents the net
force on an object, not just any single force. The net force must be determined by treating each
individual force as a vector and then adding those vectors together.

Newton’s third law states that every action has an equal and opposite reaction. As an
example, if you push against a wall with 10N, the wall also pushes against you with 10N.

3.2 Energy and Work


NOTE: Sections 3.2 through 4.3 of this handout are heavily based on Everaise’s Astronomy Book.

3
USAAAO Guide Harsh Ambardekar

Kinetic energy is the energy of an object due to its motion while potential energy is the
energy of an object due to its position. Energy is a scalar and given in units of joules (J). For a
point particle of mass m moving at velocity v, the kinetic energy K is defined as
1
K = mv 2
2
GM m
The kinetic energy of an object in orbit of radius r, of mass m around a mass M is K = 2r
.

For a constant force F exerted on an object as it is displaced a distance d, the work done
can be calculated with the equation W = F d cos(θ), where θ is the angle between the force and
displacement vectors.

The quantity work is useful because it is directly related to kinetic energy according to the
Work-Energy Theorem: The work done on an object is equal to its change in kinetic energy.
That is W = ∆K).

The gravitational potential energy is the energy some object has relative to the field formed
by the gravitational pull of another. The formula for the gravitational potential energy of a mass
m relative to a mass M a distance r away is: Ug = − GMr m , where G is the universal gravitational
constant equal to 6.67 ∗ 10−11 kgN
2 m2 .

The Virial Theorem states that for a body in a circular (ONLY CIRCULAR, NOT ELLIP-
TICAL) orbit, the kinetic and potential energies of the body are related as follows: 2K + U = 0,
and if we let E represent the total energy of the body, then E = K + U = −K. Thus, the total
energy of a body in a circular orbit is always negative. This theorem is very useful in solving
problems (see practice problem 4 for a simple application).

The reason energy is useful is because of the law of conservation of energy. We will explore
this more in the orbital mechanics handout.

3.3 Circular Motion


So far, we have dealt mainly with translational motion, with accelerations constant in both mag-
nitude and direction. In this section, however, we will deal with a very different kind of motion:
one in which the magnitude of the velocity and acceleration is constant, but the direction of both
the acceleration and velocity constantly changes. This situation is called circular motion and
includes many examples such as the earth rotating around the sun and a ball being whirled around
on a string.

Since the velocity (and hence kinetic energy) of the particle is constant, we can then conclude
that the net force and velocity vectors for the particle must be perpendicular. This leaves us with
two possible directions for the acceleration vector, towards the center of the circle, or away from
the center of the circle. Intuitively, since acceleration is the rate of change of velocity, we should
conclude the acceleration vector is directed towards the center of the circle. We will call this inward
direction the centripetal direction.

4
USAAAO Guide Harsh Ambardekar

For an object traveling in a circle of radius r at a constant speed v, the acceleration experienced
by the object has magnitude
v2
a=
r
The angular frequency ω is defined as ω = dθ dt
, or (because s = rθ) ω = vr . For an object
undergoing uniform circular motion, we also have that the period

T =
ω
From Newton’s second law, the centripetal force Fc is equal to

mv 2
mac =
r
It is very important to understand that the centripetal force is not an actual force; it is simply
the net force in the centripetal direction for an object undergoing uniform circular motion. What
this equation says is that if you add all the centripetal components of the forces on an object
2
undergoing uniform circular motion, that number will be equal to mvr .

4 Momentum and Collisions


When studying collisions, it is often difficult to analyze the motion of objects using only forces.
We thus introduce the concept of momentum.

4.1 Linear Momentum


The linear momentum of an object is defined as the product of its mass and velocity; →

p = m→

v.
dp
Using this definition, we can also express Newton’s second law as Fnet = dt .

4.2 Impulse
We now define a quantity known as impulse for a force acting on an object. We like to think of
one object applying an impulse on another object when it applies a force on it.

− →

The impulse J of a force F acting on a particle over a specific time interval is defined as
Z tf

− →

J = F dt = Favg ∆t
ti

For forces of constant magnitude, J = F t. The impulse of the net force acting on a particle during
a given time interval is equal to the change in momentum of the particle during that interval.
Jnet = ∆p. This can easily be proven using F = dpdt
and integrating.

5
USAAAO Guide Harsh Ambardekar

4.3 Center of Mass and Conservation of Momentum


The center of mass is a position defined relative to an object or system of objects. It is the average
position of all the parts of the system, weighted according to their masses.

When two objects collide, they exert equal and opposite impulses on each other. This is because
they exert equal and opposite forces on each other by Newton’s third law, and these forces are
exerted over the same time interval since both objects are experiencing the collision. Therefore,
the total momentum of the system will not change, nor will the velocity of the center of mass.
This principle is known as the law of conservation of momentum.

In an elastic collision, the kinetic energy of the system is conserved in addition to the linear
momentum. Furthermore, in an elastic collision, the velocities of the objects in the center of mass
reference frame are reversed. The reason that this trick works is that kinetic energy is conserved in
an elastic collision. In an inelastic collision, however, kinetic energy is not conserved, but linear
momentum is still conserved.

Example Problem 4.1: Ball A of mass 2 kg is moving at 3 m/s to the right and collides
with Ball B of mass 1 kg, which is at rest. If the collision is elastic, what are the final
velocities of the balls? If the collision in inelastic (the balls stick together), what is the final
velocity of the balls?

Solution: By conservation of momentum, the initial center of mass velocity in both cases is
vcom = mAmvAA+m
+mB
B vB
= 63 = 2 m/s. If the collision is inelastic, the two balls will be moving together
after the collision. Therefore, both balls will be travelling at the same speed as the center of mass.
Since the velocity of the center of mass doesn’t change, this means that both balls will have a final
velocity of 2 m/s.

If the collision is elastic, we want to go to the center of mass frame to simplify things. To do
this, we subtract 2 m/s from the velocities of the balls. This means that in the center of mass
frame, the initial velocity of ball A is 1 m/s and the initial velocity of ball B is -2 m/s. Now, we
just flip the signs to get the final velocities of the balls in the center of mass frame. To get back
to our initial frame, we must add back the center of mass velocity. Thus, the final velocity of ball
A is −1 + 2 = 1 m/s and the final velocity of ball B is 2 + 2 = 4 m/s.

5 Conclusion
In this handout, we learned many essential physics concepts. Of course, this handout was a very
quick overview of basic mechanics and was meant to be more a review than a lesson. If you want to
learn more physics in-depth, either take a physics class at your school or learn by yourself! There
are many, many resources (textbooks, videos, courses) online to help you get started.

6 Practice Problems
Practice Problem 1: Two trains initially 500 meters apart are moving towards each other, both
at a constant speed of 5 m/s. How long does it take the trains to collide?

6
USAAAO Guide Harsh Ambardekar

Practice Problem 2: You swing a ball attached to a string around in a horizontal circle. Suppose
the tension in the string is 50 N, the mass of the ball is 5 kg, and the radius of the circle is 2 m.
What is the speed of the ball? What is the ball’s period of revolution?

Practice Problem 3: A ball of mass 5 kg moving at 10 m/s enters a rough surface (friction
is present) and comes to rest. How much work did friction do on the ball?

Practice Problem 4: An asteroid in a circular orbit around a planet has a kinetic energy of
50 J. If the radius of the asteroid’s orbit is doubled, what is the new kinetic energy of the rocket?
By how much does the gravitational potential energy of the change? (hint: use the Virial Theorem
for the second part)

Practice Problem 5 (2020 USAAAO First Round): Calculate the speed of the sun around
the center of mass due to the presence of Jupiter. You may look up the mass of Jupiter, the mass
of the Sun, and the distance from Jupiter to the Sun.

Practice Problem 6: Search up gravitational slingshots and try to understand the concept.
Here is a good place to get started.

You might also like