You are on page 1of 450

1 | P a g e

Things you should have learned in high school, but may have
forgotten....
Scientific Notation
When numbers are very large or very small, you should write them in a short-hand
form. For example, let's say you have a number like 560,000,000 or 0.0000003.
Would you want to write those numbers over and over again? How can you simplify
them? First of all, there is the way that the power of 10 works - that you have the
following relations -

10
1
= 10 (anything to the power of 1 equals itself)
10
2
= 100
10
3
= 1000
10
4
= 10000
and so forth.

You'll notice how the power on the 10 equals the number of zeros after the 1. How do
you write 560,000,000? You might notice that this number is also equal to 5.6 x
100,000,000, which can be written as 5.6 x 10
8
. This format is known as scientific
notation. What about numbers smaller than 10? How do you write those? Here's how
-
10
0
= 1 (any number to the power of 0 = 1)
10
-1
= 0.1
10
-2
= 0.01
10
-3
= 0.001
and so forth.
These aren't as easy as the others. In this case, the power on the 10 could represent
how many places the decimal point is moved to the left of the 1. Now back to the
original question: how do you write 0.0000003 using this? You might note that
0.0000003 is equal to 3 x 0.0000001, which is 3. x 10
-7
. You'll notice that in both
cases, the decimal point is placed after the first non-zero number. This is the normal
way that these numbers are written and it is also useful to use this method when
writing similar numbers.
Let's try some more. How would you write the following numbers?
1. 0.00045
2. 345000
2 | P a g e

3. 0.066
4. -0.000102
5. -53000
Check your answers here.
The next thing that you might have to do with numbers in scientific notation is to
multiply or divide them. Generally this is pretty easy with a calculator. One of the
things people don't know about their calculators is that there is usually a built in key
that allows you to represent numbers in scientific notation in your calculator. The key
is usually labeled with EE or Exp - what it actually is called will depend upon the
type of calculator you have.
Let's go through an example of how you put a number into your calculator. Let's say
you want to do the following problem -
4.5 x 10
-5
x 3.3 x 10
6
.
Here's what you do -
1. Put in the front part of the first number, 4.5 in this case.
2. Press the EE or Exp key. This usually causes a "00" or a "x10" to show up.
3. Enter the power on the 10, in this case -5. You don't want to use the subtraction
key; use a key labeled or +/-. Don't put a "10" in since your calculator already has
taken care of that.
4. Now press the multiply key.
5. Enter the second number, first the 3.3.
6. Press the EE or Exp key.
7. Put the power that is operating on the 10 in, which is 6 in this case.
8. Now press the equals or enter key to finish off the calculation.

If everything went all right, you should have gotten 148.5 (or 1.485 x 10
2
, 1.485E2, or
1.485e2). Any way, that's the answer you should have gotten. If you didn't you should
review the steps.
Here are some examples to try out - try to get the correct answer in each case
1. 9.9 x 10
-5
x 4.5 x 10
-8

2. 1.02 x 10
4
/ 3.3 x 10
-9

3. 4.0 x 10
-2
x 1.0 x 10
8

4. 8.2 x 10
9
x 5.3 x 10
8

5. 9.8 x 10
19
+ 4.2 x 10
-9

3 | P a g e

Check your answers here.
In case you don't have a calculator, you can still do the math, mainly the
multiplication and division, by following these rules -
Multiplication: multiply the two front numbers, and then add the powers.
Division: divide the two front numbers, and then subtract the powers.

If you don't believe this, try the practice problems above without a calculator or
without using the EE/EXP key. You should get the same answers. Actually, you will
only want to do this with problems 1-4 since the last one is an addition.
WORD OF WARNING: If you use a calculator to work with numbers in scientific
notation, your calculator may write them in a way that they are not normally
written. For example, the number 3.4 x 10
22
could appear in your calculator
like 3.4 22 or 3.4
22
. The "x 10" part is often excluded to save space. If you
were to do a calculation and your calculator gives an answer similar to that shown
above (without the "x 10" part), make sure you write the number out correctly - don't
forget to write out the "x 10" part.
Why? There is a big difference between those numbers. If your calculator displays
3.4
22
, and you write on your answer sheet 3.4
22
, you'll lose points, since 3.4
22
means
3.4 taken to the power of 22, not what it is supposed to mean (3.4 x 10
22
). That's a big
difference - don't be lazy; write out the number properly.

When do you use scientific notation? This is one of those questions that doesn't
have a solid answer. If you have a number like -0.2, or 123, you really don't need to
write them in scientific notation, since that would be a bit silly and make the number
harder to read. In general it is best to use scientific notation if the number is in the
millions or greater, or if it is smaller than 0.001. While these are only guidelines you
should do whatever you are comfortable with.
Accuracy - Significant Figures
One thing that you'll run across if you use a calculator is that it is very literal in doing
calculations. For example, if you divide 10 by 13 you'll get 0.076923076923... Do you
really have to write all those numbers down? No, of course not. You should round the
numbers off so that there are the correct number of significant figures (SF). These
are the number of digits that are needed to give an accuracy that is appropriate for the
problem. Here're a few rules to follow -
4 | P a g e


1. Digits other than zero are always significant.
2. Zeros between other SF are significant. For example, 4003 has four SF.
3. Zeros to the left of the first non-zero digit in a number are not significant; they
merely indicate the position of the decimal point. For example, 0.033 has two SF,
0.000401 has three SF.
4. When a number ends in zeros and the zeros are to the right of the decimal point,
they are significant. For example, 0.00330 has three SF - 3, 3, and the rightmost 0 in
this case.
5. When a number ends in zeros that are to the left of the decimal point location, the
zeros are not necessarily significant. For example, 440 certainly has at least two SF,
but the 0 may or may not be significant. You'd probably have to know more about the
number, particularly how it was determined. One way to remove the ambiguity is to
include the decimal point in the number. If I wrote "440." then I would want to
include the zero as a SF, so there are now three SF in the number. Writing just "440"
doesn't make it clear as to whether the zero is significant or not. Including a decimal
point will help. So "329000" has three clearly SF, but the rest are uncertain. The
number "329000." has six SF.

Sometimes a zero is included after a decimal to show that there is a need for greater
accuracy, so that 3.20 has three SF, while 3.2 only has two. For some reason the
person who wrote 3.20 wanted greater accuracy when the number is used in further
calculations, and the rounding rules for significant figures takes effect (we'll get to
those later).
Here are some for you to try - determine the number of SF in each number
1. 6.751
2. 0.157
3. 28.0
4. 2500
5. 0.070
6. 30.07
7. 0.0067
8. 6.02 x 10
23

Check your answers here.

Now to actually use SF. If the numbers you are using are measurements of some sort,
odds are they are not exactly accurate, and the level of accuracy is given by the
number of SF. In the case of addition and subtraction, your final answer should have
the same number of decimal places as the value with the least number of decimal
5 | P a g e

places. If you were to take 9.221 - 7.01 your answer should be 2.21, not 2.211. It is
best to determine the number of SF when you get to the end of all of your
mathematical steps - so when you are ready to write down the final answer, double
check to see how many SF you should write down.

For multiplication and division, the answer cant be more accurate than the least
accurate part. If you were to multiply 3.209 by 2.2 your answer should have only two
SF in it since that is the least number of SF in the values you were given. You should
write the answer as 7.1, not as 7.0598, which is the number your calculator spits out.
A word of warning - if you are using a number that is not a measured quantity like a
constant that has an exact value, you should not use it to decide the number of SF in
your final answer. Exact values or constants are considered to have an infinite
number of SF. For example to calculate the circumference of a circle, you use the
formula 2 r, where =3.14 (or more SF) and r is the radius of the circle. How many
SF will your answer have? Should there be only 1 SF, since that is how many are in
the 2? No. The 2 is not a measurement but a constant, a part of the formula. The
same rule applies to the The number of SF would depend upon the number of SF in
the r, so depending upon the value of r, the number of SF in your answer can vary.
Here are some examples to try - determine the answer for each using the proper
number of SF.
1. 3.4 + 0.00344
2. 4.50 x 3.3005
3. 9.01/7.88
4. 4.510 x 10
12
x 3.401 x 10
-11

5. 607.1 x 4.4
Check your answers here. You should ALWAYS follow the rules of SF when you
do math problems, especially when you calculator spits out numbers like
7.38029347234. If an answer should have only three SF and you write out an answer
like 7.38029347234, you will get points taken off. If you aren't sure how many SF to
include, ask.
Units of measure
For the most part, metric units are used in astronomy. Here's a quick re-cap of some of
the common ones you'll run into -
Mass
6 | P a g e

kilograms - kg
grams - gm
Distance/Length
meters - m
kilometers - km
centimeters - cm
millimeters - mm
ngstroms -
The trickiest thing is trying to remember what each of these things are in case you
have to convert from one unit to another. Here is a listing if you need to convert from
one value to another.
1 kg = 1000 gm
1 gm = 0.001 kg
1 meter = 100 cm = 1000 mm = 0.001 km = 10
10

1 cm = 10 mm = 0.00001 km (10
-5
km) = 0.01 m = 10
8

1 mm = 0.000001 km (10
-6
km) = 0.001 m = 0.1 cm = 10
7

1 km = 1000 m = 100,000 cm = 1,000,000 mm = 10
13

1 = 10
-10
m = 10
-8
cm = 10
-7
mm = 10
-13
km

There are also some non-metric distances that are set up just for convenience. For
example, there is the distance between the Earth and the Sun, which is defined as 1 A.
U. (Astronomical Unit). This is useful for measuring distances within the solar
system. There is also the distance of a light-year - the distance light travels in one
year. Parsec is another distance which is often used interchangeably with light-year.
These are often used to indicate the distances between stars. For even greater
distances there are kiloparsecs and kilo light-years (1 kiloparsec = 1000 parsecs, 1 kilo
light-year = 1000 light-years) and for very great distances there is the mega parsec (a
million parsecs) and mega light-years (a million light-years). The actual values for
these distances and other common units of measure can be found in the table
of constants.
Introduction to Astronomy and Motions of
the Sky

7 | P a g e

What's covered here:

What is astronomy all about?
How do objects, like the Moon, Sun, Planets and Stars, appear to move in the
sky?
How do we model the sky, in particular, the positions of the stars?
Why is the North Star so important?
How is the sky mapped out and how are locations defined?
How high above the horizon do objects get?

What's this all about
What exactly is astronomy? What do astronomers study? They study pretty much
everything from the smallest atoms to the entire Universe. That's a pretty wide field of
study, to say the least. They can do some rather boring things such as using math and
physics to figure out how the Universe and everything in the Universe works - how
things formed, how they move, how they live their lives and possibly even how they
will ultimately be destroyed. They look at how things move, and unlike
meteorologists, they are able to predict things, such as when the Sun will rise, when
the Moon will set or a whole slew of other things. Some astronomers spend all of their
time gathering images of objects in the sky using a wide range of instruments. Of
course, there is not a lot that they can do on cloudy days - though you'll see there are
exceptions to this rule. Astronomers study such things as planets, stars, galaxies,
moons, asteroids, comets, space dust, gas between the stars, groups of galaxies, black
holes, the shape of the Universe and ultimately the fate of the Universe.
If you want to really get an astronomer mad at you, just mention astrology. This will
not win you many friends in the astronomical community. However, like an old
skeleton in the closet, we (astronomers) can't ignore some of the concepts that are
usually associated with astrology. Of course, if you want to make a lot of money from
a bunch of gullible people, become an astrologer - I wouldn't recommend it, but it
can't be denied that astrologers can make some big bucks, mainly because people are
just too easily boondoggled. Just remember, if you ever use the word astrology in
class when you actually should say astronomy, I just might have to make you stand in
the corner.
The history of astronomy is pretty long; in fact, you can think of astronomy as the
oldest of all of the sciences - not to be confused with the oldest profession! In spite of
this long and noble lineage, we can't ignore the linkage between astronomy and
astrology. In fact, in the old days, astrology was even considered a serious (!) topic of
8 | P a g e

study, along with astronomy. Both were important, at that time, in trying to figure out
how the sky operates. Why was it important that we knew how the sky operated?
In the old days, and I don't mean 1974 by "old," before the invention of such annoying
things as digital watches, cell phones and other modern bits of technology, people's
lives depended upon the sky. Remember, most people in the old days were not
formally educated, so they might not have been able to understand a clock or a
calendar (reading was generally only for the rich and wealthy, and those who worked
for them). Most people lived in an agriculturally based society - not too different from
Iowa, but even more rustic. What does that have to with astronomy? In those days, it
was important to know when to do things - remember, most of the people wouldn't
know a calendar from a hole in the ground - well, maybe they would if they fell in the
hole, but you know what I mean. Any ways, it was important for the people to know
when to do the things that were vital for their existence. Since most people were
farmers, they needed to know when to do such important things as plant the crops,
harvest the crops, expect the winter to come to an end, and check to see if you still had
enough food left to survive. These were pretty basic things, and there was no CNN to
tell people what to do and when to do it.
What does planting, harvesting and such have to do with astronomy? By knowing that
a certain constellation was visible in the evening sky (or the morning sky), people
knew it was time to harvest; they knew that at this time of year, the rivers would tend
to flood; or by noting the position of the Sun at noon in the sky, they knew that it was
a month until the end of the cold weather - killing frosts. Of course, people, being
rather uneducated, were also under the belief that the stars and planets in the sky
would also effect their lives just by their locations in the sky. Yes, this is what that
other "a" word (astrology) is all about. Since people were so dependent upon the
motions of objects and the predictable patterns in the sky (astronomy) to stay alive, if
they could use another aspect of the sky (astrology) to hedge their bets, then it perhaps
helped them to survive. Generally, astrology never helped anyone, except for the
astrologers who got rich off of all of the fools that believed them.
Also, other aspects of the sky were important, and are still important today. Do you
know what the phase of the Moon is today? Do you think that is important? You
probably don't think it is, but several major cultures still use the Moon as a basis for
their calendars - these include people who use the Islamic, Judaic, or Chinese
calendar. That's a pretty large number of people. You still think that you don't care
about what the phase of the Moon is? You don't think it's important for the typical
Iowan? Here's a question - how is the date of Easter determined? You know it is on a
Sunday, but the date changes from year to year. Why? I'll tell you - it depends upon
both the Sun and the Moon. The day that Easter is celebrated on is the first Sunday
after the first Full Moon after the first day of Spring (also called the Vernal Equinox).
9 | P a g e

You can go look it up - you'll see that this is how it is set up! The first day of Spring is
based upon the location of the Sun in the sky, so that and the phase of the Moon
determine the date of Easter.
In the old days, it was important for people to know how the objects in the sky moved
and when they were in certain locations. What are you going to do if it is cloudy all of
the time? How would you know when to plant your crops if you can't see the stars, the
Sun, the Moon, or whatever object you use as a calendar? What then? This is sort of
where the science (and math) comes in. Astronomers not only can tell where things
are, but are also able to predict where they should be. In this way, they can tell people
when to do their agricultural work at the right time, regardless of the weather. Here's
something you might not have known - in western Europe, most of the people who did
astronomy (and astrology) were employed by the Christian church, since they needed
to know when various festivals or celebrations had to be held. People working for the
church, such as the priest and monks, tended to be literate and were able to do
complex calculations that showed how the objects in the sky moved. If you wanted to
know when you were supposed to do various things, you needed to know when things
were going to be at certain locations and to do that you had to figure out how
everything worked - and this isn't easy.
Of course, if you were to make your fortune suckering people with astrology, then you
also needed to know math and the workings of the sky, since you had to know where
things were or will be in the future. If you can figure out the motions of the planets
along with the Moon and the Sun, then you can really make some serious money in
the astrology trade, so even the non-scientific astrologers did have to know enough
science to cast their horoscopes. Of course, nowadays, astrologers can be total idiots
since a computer can figure out everything for them.

Now that I've convinced you about the importance of astronomy in the old days and
how it used to be linked with astrology, let's just see how really tricky it is. What were
those complicated motions that your friendly neighborhood astronomer had to predict
and understand?
Motions of the Sky
First a word of explanation - when we talk about stuff "rising" and "setting" it is the
appearance of something rising or setting. In reality objects like the Sun or Moon
aren't physically rising from the horizon. Most of the motions we see in the sky are
caused by our (the Earth's) motion, both its rotation and orbital motion. It is just
common to say "the Sun rose in the East today" rather than to say "the Earth rotated
toward the direction of the Sun, which caused it to appear to rise in the east". So all of
10 | P a g e

the "motions" described below are really apparent motions - not really an indication
of how things are actually moving, but how they appear to move.
Motions/Events of the Sun
The Sun is pretty bright, and without it, all life on Earth would be nonexistent, so let's
see how complicated its motions are. We'll just consider the motions that you would
see in Iowa -
It rises along the eastern horizon and sets along the western horizon - but it
doesn't rise exactly due east and it doesn't set exactly due west. The exact
location on the horizon where it rises and sets changes over the months.
Another thing that changes is the time of day it rises or sets. Both of these
aspects of the Sun's motion change with the seasons.
Actually, there are two times during the year when the Sun does rise exactly
due east and set exactly due west. These times are used to determine the starts
of some seasons.
The Sun appears to travel along a different path of the sky during the day over
the course of a year. What does that mean? It means that sometimes the Sun
appears very high in the sky and sometimes it appears closer to the ground -
even at noon, it looks rather low. This changes over the course of the year as
well.
The Sun appears to move pretty fast - it gets back to the same position in the
sky in about 24 hours, but the number of hours that it is visible varies with the
seasons (fewer hours of daylight in winter and more hours of daylight in the
summer, which of course is when it is visible for either shorter or longer
intervals of time. This doesn't mean that the Sun is actually "moving" faster in
the different seasons, it just means that the path it appears to travel is different
in the different seasons).
The motion of the Sun across the sky appears to be pretty fast, but is it really
moving? Does it ever move? How can you tell?
Occasionally there are times when amazing things happen to the Sun - these are
events known as eclipses and these will cause the Sun to disappear from view
for a few minutes.
If I were to ask you to come up with a method that would allow you to determine the
exact rise or set time of the Sun for, say, May 17th, could you do it? This is actually a
pretty difficult problem and we won't tackle it (yes, you can breath a sigh of relief
now), but there are some other things you'll have to predict - just stay tuned.

Motions/Events of the Moon
11 | P a g e

After the Sun, the Moon is the most obvious thing in the sky. Even on hazy or semi-
cloudy nights, it is often possible to spot the Moon, so what are its motions that people
have to worry about?
It rises somewhere along the eastern horizon and sets somewhere along the
western horizon, but again not generally in the same place from one day to the
next. Unlike the Sun, there are no special days when the Moon will rise due
East or set due West.
The time that the Moon rises or sets is not always the same - it varies; typically
it will rise about one hour later each night.
The path it appears to travel is not the same as the path of the Sun; sometimes it
is on a more southernly path (so it is below the Sun's path as we view it from
Iowa), and sometimes it is on a more northernly path (above the Sun's path).
The most obvious aspect of the Moon is its appearance. These are the phases
that the Moon goes through. It takes about one month for the entire phase cycle
to be completed. Why does it do that? Can you predict the phases?
The motion of the Moon across the sky is noticeable when you compare it to
the stars - it can be in front of stars of one constellation during the course of
one night, and the next night it will be amongst the stars of another
constellation. It takes about one month for the Moon to traverse the sky and end
up back where it started (back by the stars that it started out near).
(Click here to see an animation of the motion of the Moon relative to the stars
at one day intervals)
Like the Sun, there are times when the view of the Moon is hindered. While in
the case of the Sun getting eclipsed the entire Sun may be hidden, during a
Moon eclipse (actually, we call it a Lunar Eclipse) it still may be possible to see
the Moon, but it may be colored dark brown, red or orange - sort of freaky.
Lunar eclipses are visible more often than solar eclipses by folks in Iowa.
In a way the motions of the Moon are even more complicated than the Sun's motions.
There are some aspects of the motions that are pretty easy to understand, as you'll see.
Motion of the Stars
The other main players in the evening sky are the stars, and I don't mean Brad Pitt or
Angelina Jolie, but those twinkly little things in the sky. What do we know about
them just by looking at them with our eyes?
Stars seem to remain in fixed patterns in the sky, which we call constellations.
They stay in these patterns and don't appear to change over your life time or the
life times of generations. Of course, people give them names, sometimes silly
12 | P a g e

names, but they do help us find our way in the sky since they appear to remain
in fixed patterns.
If the stars are in "fixed" patterns that means they don't move, right? WRONG!
They appear to move (while staying in their constellation patterns) over the
course of the evening, as anyone who has slept under the stars can tell you. If
you watch them over the hours, you will notice that most stars appear to rise
somewhere in the East and set somewhere in the West over the course of an
evening.
You may have noticed that I said that "most stars" appear to rise and set. This is
because not all stars rise or set. Some are visible in the sky the entire night.
Even if you could see them in the daytime, you would still never see them set.
These are the stars that are near the pole star, the star you may know as Polaris,
the North Star (it has nothing to do with snowmobiles, or a hockey team). Stars
that never rise or set do move, though. If you could watch them move over the
course of an evening, you'd see them move in a circle around Polaris, which
doesn't appear to move at all. These stars, those near Polaris that never set, are
called circumpolar stars.
Stars rise and set and move over the course of the evening, but do you always
see the same stars (constellations) over the course of the year? No; you will see
different constellations over the course of the year. The constellations in the
early evening sky in January and the constellations visible in the early evening
sky in July are different - except for those circumpolar stars that never seem to
go away.
If you were to watch one star every night for about a week, you'd notice that
this star would rise earlier and earlier each night. After many weeks different
constellations will become visible in the early evening sky because of this
gradual shifting of the rise time of stars.
Of course, if you have looked carefully at the stars, you have noticed that they have
different appearances - some stars are brighter and some are fainter. If you have good
eyes and look at stars carefully, you may also notice that the colors of stars vary - this
is particularly easy to notice with the really bright stars. Some stars are bluish-white,
others are yellow, and others are red.
Motions of the Planets
With your unaided eyes (that means without using binoculars or a telescope) you can
see six planets (can you name them all?). How can you tell that an object in the sky is
a planet and not a star? I'll tell you after I go over the motions of the planets (yes, I do
like to keep my students in suspense!).
13 | P a g e

Like pretty much everything else, planets rise somewhere in the East and set
somewhere in the West. The path they appear to follow is pretty close to the
path that the Sun and Moon appear to follow.
Not all planets move in the same way. Some are only seen near the Sun, either
around the time of sunrise or sunset (Mercury and Venus are the two that do
this - in the animation, the Sun is the yellow dot, Mercury is the red dot, and
Venus is the bluish dot. The stars have been removed to show the motion
clearly). This causes people to sometimes call them the Morning or Evening
Stars. This is particularly true of Venus, which is the brightest thing up in the
sky after the Sun and the Moon.
The other planets have some rather interesting motions. Planets like Mars,
Jupiter and Saturn move generally toward the East compared to the stars over
the course of weeks or months. The animation shows how Mars moves over
several weeks' time relative to the stars.
The planets don't have a uniform motion. By this I mean they don't travel
amongst the stars with the same speed. Sometimes they move faster and
sometimes they move more slowly. There are even times when they will stop
and then appear to go backwards for a while. This retrograde motion was one
of the really complex problems concerning the motion of the planets. Why
would a planet want to go backwards?
Here's the really tricky bit - how can you tell the difference between a star and a
planet? You might remember a song that goes "twinkle, twinkle, little star..."
Have you ever heard of a song that goes "twinkle, twinkle, little planet"? I
thought not. That's because planets don't twinkle. In case you aren't sure,
twinkling is the shimmering or blinking appearance that stars, especially bright
stars, have. Planets don't twinkle, and since the planets that are visible to the
naked eye are pretty bright, you can tell they are planets by not seeing them
twinkle. As long as it isn't an airplane, if it's really bright and really steady in
brightness, then it is probably a planet. As to which planet it is, you might have
to guess or just pick one out of the air - just make sure you don't tell anyone
that it is "Vulcan."
Have you figured out the six planets that you can see with your naked eye? Did you
remember to count the one that you live on?
Models of the Sky - Celestial Sphere
All of the motions that I went through are pretty complex, and they all need to be
explained, so that you can predict the motions, so that you can set up some sort of
calendar system, so that you'll know when to celebrate your various festivals and
celebrations or when to plant or harvest your crops, or, well, whatever you need to
14 | P a g e

know. To explain all of these motions, early astronomers/astrologers needed to figure
out the mechanics of the sky - they needed a model (something that could explain the
motion; not necessarily a physical model, but at least something that they could write
down). Since the motions of the stars are fairly uniform, the model for their motions is
easiest to do and we'll tackle it first. One of the concepts that people believed in the
old days was that the stars (and everything else in the sky) were moving and the Earth
wasn't. This sort of made sense, mainly because people did not feel any motions from
the Earth - they had no sense that it was moving, like you would feel motion on a ship,
a horse or a wagon. Obviously to these people, the Earth didn't move. Now we know
that this idea is totally bogus, but in the old days it made sense, and it also helped
astronomers to figure out how to make a
working model of the sky.
The early, simple model of the sky is known
as the Celestial Sphere. What is it?
Basically a big see-through imaginary globe
around the Earth. The stars are stuck on this
sphere, and as the sphere spins around, the
stars would move with it. This really
explains how the stars move pretty clearly,
since they are on this spinning globe, those
near the axis of the rotation, like near
the North Star, would make little circles
around the pole star, while those further
from the pole would move along larger paths
across the sky. If you don't get this idea, think about how an umbrella looks as you
spin it around above you. If you have a bunch of spots on the inside of the umbrella,
the spots far from the handle will make large circles around as you rotate the handle,
while those near the handle will make little circles. If you were to take the ends of the
umbrella and stretch them around you would have a sphere, just like a Celestial
Sphere surrounding the Earth. People thought that the stars' patterns never changed
(they stayed in the same constellation patterns) since they were all "stuck" on this big
sphere. Stars also never appeared to change their appearances, their colors or their
brightnesses - at least not in a way that was noticeable. By putting them on a fixed,
big sphere, people sort of made them eternal. The worst part of it all is that this model
worked really well! BUT IT IS TOTALLY COMPLETELY WRONG!!! One of
the obvious problems is that it implies that all of the stars are at the same distance
from the Earth, and this is not correct. Even though this model of the sky is not correct
in terms of the real Universe, it is useful in determining positions of objects and
defining a coordinate system of the sky. Just remember - IT IS WAY WRONG!!!!
15 | P a g e

The constellations that the stars are associated with used to just be traditional
groupings. Different cultures had different constellations. When astronomy got more
scientific and organized, the little stick figures you see for the constellations sort of
changed over time. Today there are 88 constellations in the sky and they are actually
sort of like counties in a state. There are 99 counties in the state of Iowa and each
piece of land in the state is part of some county, based upon where the boundaries are
drawn. This is also how we now divide up the sky - the constellations don't just define
a stick figure of a dog, horse or person, but a region of the sky. Something in that
region belongs to a certain
constellation. This is shown in
Figure 1.
Figure 1. The territory that
belongs to a certain constellation
is shown by the green lines, while
the red lines show a traditional
way of drawing the "shape" of the
constellation. The constellation
boundaries are not subject to
change, but the lines that
represent the stick figure for the
constellation can be drawn in any
sort of pattern.
Even though the Celestial Sphere
is not a real model of the sky, it is
useful for mapping out locations of objects in the sky. The Celestial Sphere was set
up so that it used some of the coordinates that are used on the Earth. Remember, the
Earth was thought to be at the center of the Celestial Sphere, so it wasn't too difficult
to extend the Earth's coordinates to the sky. Putting the Earth in the middle of
everything may seem egotistical, but there were other reasons for doing it. Figure 2
shows the basic set up for a Celestial Sphere. There are a couple of special locations
that need to be pointed out.
Figure 2. The Celestial Sphere, an imaginary ball
around the Earth upon which the stars were
thought to be located. This is not true - the stars
aren't on this sphere, but it does provide an easy
way to map out the sky.
Celestial Poles - these are points on the Celestial
Sphere that are directly above the Earth's Poles, so
16 | P a g e

there is aCelestial North Pole and a Celestial South Pole. You can also say North
Celestial Pole and South Celestial Pole; it doesn't really matter how you say it.
Celestial Equator - this is just a line that is directly above the Earth's Equator. Like
the Earth's equator, the Celestial Equator goes all the way around the Celestial Sphere.
We need to simplify some locations in the sky now. We can talk about stuff that is
directly over head, or we can say stuff is at your zenith - this just means the location
right over your head. You have a personal zenith and what is there depends upon
where you are on the Earth, what time of day or night it is and what time of the year it
is. There is also a term for the exact opposite of zenith, but we don't really care about
that since it would be in the direction of the ground and there are no stars visible down
there.
Another special direction is the horizon, though that is not really one particular
direction, but sort of, well, your horizon. To be kind of technical, the ground meets the
sky at the horizon, and generally, there are 90 between the horizon and the zenith
(especially in Iowa). If you do want to get specific about the horizon, you could say
the southern horizon, the northern horizon, etc.
Here's a rather surprising concept that you might not have known about - the ancient
astronomers did not, I repeat, NOT, believe that the Earth was flat. In fact, it was
pretty much agreed that it was spherical, though the size of the sphere was so big that
to us puny humans it looked like a flat surface. Actually, as you'll see, ancient
astronomers were sort of fixated on everything being spheres. We'll get to that later.
As previously mentioned, the Celestial Sphere was useful for finding your way about
the sky, since like on a globe of the Earth you can designate various coordinates to
measure locations or positions of objects. First, we'll tackle the north-south coordinate
system. On the Earth, the amount that you are north or south of the Earth's equator is
measured in degrees of latitude. There are certain special locations on the Earth. If
you are at the Equator, you are at a latitude of 0; if you are at the North Pole, you are
at a latitude of +90 or 90 North (this is the highest possible value) while at the South
Pole your latitude would be -90 or 90 S.
We're not at one of those locations - we're in Cedar Falls (at least I am). How do we
determine our latitude? Simply draw a line from our location to the center of the Earth
and a line from the Equator to the center of the Earth (see Figure 3). Now measure the
angle between these two lines - this is our latitude! For Cedar Falls, our latitude is
42.5 N. The latitude of any location on the Earth can be found by measuring its
angular distance from the Earth's equator. Draw a line from the North pole to the
17 | P a g e

center of the Earth and you'll end up with an angle of 90 - which is the latitude of the
North pole!
Figure 3. The definition of latitude. The angle
between your location and the equator (where
latitude =0) determines the value of your
latitude.
Now, if we were to take this coordinate system
and stretch it up to the sky, we would be
cooking! Why can we do this? First of all, the
Celestial Sphere is already sort of set up the same
way as the latitude system - there is a Celestial
Equator and there are Celestial poles. Just extend
the latitude system to the Celestial Sphere to get
the angles that objects are North or South of, not the Earth's equator, but from the
Celestial Equator. We can't call this system latitude, since that name is already taken.
Instead, this coordinate system is known as declination (abbreviated as dec).
Declination is just the angle an object on the Celestial Sphere has as measured from
the Celestial Equator, just like latitude is the measure of the angle between an object
on the Earth's surface and the Earth's equator.
One really neat thing about this system is how declination and latitude are linked. An
object at your zenith (remember, that is right over your head) will have a declination
value equal to your latitude! If you are located on the Earth's equator (0 latitude), at
your zenith would be a declination of 0 (which is the declination of the Celestial
Equator). If you are at the North pole, you are really cold, and your latitude is 90 N.
If you don't freeze to death, you might notice that at your zenith is a declination of 90
N. If you are in Cedar Falls, 42.5 N latitude, then at your zenith is a declination of,
you guessed it, 42.5 N. Check out Figure 4 to see this sort of arrangement.
Figure 4. How declination and latitude are
related. An object at your zenith has a
declination value that equals the value of
your latitude.
Just like latitude, declination is measured in
units of degrees. The two extremes are at the
North and South Celestial poles: +90 to -90
respectively (or you could say 90 N and 90
18 | P a g e

S). You can't have a declination greater than +90 or less than -90!
Sometimes a degree is pretty big and you need to measure an angle that is much
smaller than a degree, so you need to use a smaller unit of measure (sort of like the
way an inch is a smaller unit of a foot). To make life easier, we can divide one degree
into smaller units known as minutes. To be precise, 1 = 60' (the dash stands for
minutes). Sometimes using minutes is not enough; even smaller units are needed. We
can divide each minute up into (you guessed it) seconds. Of course, the division is 1'
= 60 " (two dashes signify seconds). For those with nothing better to do, you might
note that 1 = 3600". An object's declination can be given very precisely as, for
example, -34 27' 41''.
A lot of times astronomers have to keep track of time (not that we ever have any dates
or something like that, but just to keep track of events in the sky). If astronomers need
to talk about minutes and seconds, like as in units of time, how can we know that they
are talking about time units and not angle units? To distinguish between angular
minutes and seconds and time minutes and seconds, the word arc second or arc
minute is often used. You could say that there are 3600 seconds in a degree, or you
could say that there are 3600 arc seconds in a degree. By using arc seconds people
would know that you are talking about angles and not time.
I should mention that angles are used for not just positions but also relative positions
and angular sizes. What's that mean? You could say that one object is 10 from
another, or you could say that an object is 10 wide. These are two different things.
One is how far apart two objects appear in the sky (their actual separation may be
quite different). The other is how big an object appears. This is basically the amount
of sky that an object covers. Two big astronomical objects that you can see with your
eye are the Sun and the Moon. How many degrees across do you think they are? 5?
10? Well, you might be surprised to learn that both are about 1/2 in size! That's
pretty tiny! You don't believe me? Well, take your thumb (you have a thumb don't
you?), hold it at arm's length and place it over the Moon (this is less painful than
holding it over the Sun). Can it cover the Moon? Your thumb is about 2 across when
held at arm's length. Also try it using a pencil held at arm's length. You might be
surprised.
We're going to divert for a moment from the discussion of the celestial sphere and all
that to talk about angles, sizes and distances. I mentioned that your thumb held at
arm's length is about 2 wide. Why did I have to say "held at arm's length"? What
happens if you move your thumb really close to your eyes? If you are pretty normal,
you'll notice that your thumb looks really big. Does that mean it has a larger angular
size? Yes; your thumb's angular size is now larger. What does that actually mean? It
means that it covers a larger region of your field of view - it appears bigger. If you
19 | P a g e

could move your thumb further away, it would look smaller - it would have a smaller
angular size (basically be less than 2 in size). Is your thumb actually changing its
size? No, of course not - not unless you hit it with a hammer during this process.
This is all rather cute and silly, but what does this have to do with science and all?
Actually, there is a relation between the size of an object, the distance the object is
from you and its angular size. The relation between these things is
S = R 0.0175
where S is the actual size of the object (how wide it is), R is the distance of the object
from your eyes and is the angular size (how many degrees wide it is). This nifty
little formula is known as the Small Angle Formula, since it really only works well
for small angles (less than 10). One of the neat things about this formula is how these
quantities are measured. R and S are in the same units - by this I mean that if R is
measured in inches, then so is S; if S is measured in kilometers, then so is R; if S is
measured in pickles, then so is R; and so on. is measured in degrees - you should
know what those are; at least, I hope you do. The number in the formula is there for
making sure everything comes out properly in the end (so that you end up with the
correct units).
If you want to actually use this formula, you need to know two out of three of the
things in the formula. For example, if you know that the Moon is 1/2 in size (= ),
and it is 3476 km wide (=S) can you then determine how far away is it? Just take the
formula and determine the value of R (the distance). The formula is
R = S /( 0.0175) = 3476 km / (0.5 x 0.0175) = 397,257 km
The Moon is 397,257 km away. Basically, if you know two of the things, you can
always get the third. This is one way to determine the actual size of an object so long
as you know the angular size and the distance. Conversely, you can get the distance if
you know the actual size and the angular size, or you can get the angular size if you
know the distance and the actual size.
We'll now go back to the coordinates discussion.
How about the East-West designations? We used latitude for the North-South system,
so let's see if we can do the same with longitude for the East-West coordinates.
Unfortunately, we can't use the longitude system since the objects in the sky move too
quickly - they aren't located constantly over the same location on the Earth. Instead,
astronomers use a system based on a clock. This is logical since it appears that the sky
gets back in the same configuration after 24 hours (again, you should remember that
20 | P a g e

this is only how it appears; the sky isn't really rotating!). Astronomers divide the sky
into 24 units known as hours, which go all the way around the Celestial Sphere.
To make your life complicated the name for the East-West coordinate system is
known as Right Ascension (R.A. for short). Values of R.A. increase as you go further
to the east. There is a location labeled 0
h
; further east is 1
h
, then 2
h
, ... and 22
h
, then
23
h
, finally 24
h
which is actually the same as 0
h
. This set up is similar to how military
time is given - numbers increase as you go east until you are back to where you
started from. If you need more precise units, 1
h
of R.A. can be divided up into 60
minutes (1
h
= 60
m
), and one minute of R.A. can be divided into 60 seconds (1
m
=60
s
).
Note the units of R.A. and dec are different, R.A. has h m s, while dec has ' " . An
object's full coordinates could be something like 2 33' 17'', 14
h
7
m
33
s
- not that you
need things so precise all of the time, but if you need to be that precise this will help
you not get lost. Every object in the sky, not just the stars, but also the planets, the
Moon, the Sun and everything else - can be located at a distinct value of declination
and Right Ascension, just like every location on the Earth can be designated according
to a set of latitude and longitude coordinates. Here is a handy little java program that
allows you to click on a map of the sky to see the coordinates - I've only included the
degrees and minutes values for declination and the hours and minutes for RA since it
is sort of a rough map. One of the things you'll see is how the RA increases in value as
you go toward the left, which is East in this view.
If you were to sit outside on a clear night, you would notice that over the course of the
evening different objects pass over your head, and other objects will appear to rise and
others will appear to set. If you were to go outside a month later you would see stars
still doing these things, but the stars wouldn't be in the same places doing the same
things at the same times as you saw them a month ago. A star that was located near
the eastern horizon last month might now be very high above the southeast horizon at
the same time of night. A star you saw in the west might no longer be visible. When
and where stars are located in the sky varies from night to night. Remember, different
constellations are visible at different times of the year - you'll see why this is in the
next set of notes.
Perhaps you want to see an object when it is highest above the horizon, as far from the
ground as possible. This is the best time to see an object, since there is less chance
that it will be obscured by trees or buildings, and the atmospheric haze is generally
less as you get further above the horizon. When is an object located furthest from the
horizon? We don't mean just when an object is at your zenith, since not all objects will
pass directly over your head.
21 | P a g e

The best time to view an object is when it is on your meridian. Like the zenith, the
meridian is a special location, though it is not just a single point. It is actually a line
that runs north-south and passes through your zenith. When objects are on your
meridian, they are at their "highest" location in the sky. You are most familiar with
this when you see the Sun high up in the sky. It is highest above the southern horizon
at about noon. At this time, it is directly due south of you. If you were to watch any
star or object in the sky and measure its height above the horizon, you'd notice the
largest angle it has above the horizon is when it is on this north-south line. Check out
Figure 5 for an illustration of this. Another feature of the meridian is that all objects
on your meridian have the same value of R.A. (this is because it is a line that runs
exactly North and South - no part of it is further to the East than any other part of it).
This is the same way that a line of longitude (which, like the meridian, runs north to
south) acts - all objects on that line have the
same value of longitude.
Figure 5. Meridian line goes from north to
south and passes through your zenith.
Objects on the meridian are at their highest
elevation above the ground. If there was a
star on the Celestial Equator, you would see
it rise in the East, be highest above the
southern horizon (when it is on your
meridian) then get lower to the ground
before it sets in the West.
Now we're going to do some fun things with objects on your meridian - and believe
me, you'll want to pay close attention to this stuff since you can be guaranteed that it
will be on the test. Let's get cracking!
Now remember that at the North Celestial pole you would find the Pole star (Polaris).
Since it is at the Celestial North Pole, it is at a dec=90 N. Okay, that's pretty easy. If
you were to stand on the Earth's North Pole, Polaris is at your zenith (or you could say
that it is 90 above the horizon). See Figure 6. Being at the North Pole is not much
fun, since it is so dang cold. Let's go somewhere warmer, like the Equator. If you
were to stand on the Earth's Equator, Polaris is no longer at your zenith, it is on the
horizon (or you could say 0 above the horizon). This is illustrated in Figure 7.
22 | P a g e


Figure 6. When you are at the North Pole,
Polaris is over your head.

Figure 7. When you are the equator, Polaris is
on the horizon.
Do you notice a pattern to this? What was your latitude at those locations? At the
North Pole your latitude was 90 N, and Polaris was 90 above the horizon. At the
Equator you latitude was 0, and Polaris was located 0 above the horizon (actually it
was right on the horizon). Wow, this is amazing! Your latitude = angle that Polaris
is above the horizon. As long as you can see Polaris, you can measure its height
above the horizon, and that angle will equal your latitude. Perhaps you can use this
information to save your life someday when you are lost in the woods. All you would
have to do would be to find Polaris, then measure its height above the horizon, and
then you'd know what your latitude was. Perhaps that wouldn't really help much to
save you if you were lost in the woods, but at least you'd be able to apply something
you learned in this class to a real life experience. Of course in the old days, when
navigators used the stars to sail the seas, the Polaris trick was really helpful.
Unfortunately it only works for Polaris, since there is no star at the Celestial South
Pole.
Let's try the Polaris thing for you at UNI. Remember the latitude of UNI = 42.5 N,
and that means that an object at our zenith is at a declination of 42.5 N. This means
that the Celestial Equator is located 42.5 south of our zenith. This is shown in Figure
8, which shows the angles and the declinations toward the northern and southern
horizons. If the angle between the zenith and the Celestial Equator is 42.5, what is the
angle between the Celestial Equator and the southern horizon? 90 - 42.5 is 47.5, so
the Celestial Equator is 47.5 above the southern horizon. You may want to make note
of this angle in Figure 8. Let's see what is happening on the northern side of Figure 8.
The difference in declination between what's at your zenith (dec=42.5 N) and the
North Celestial Pole is 90-42.5=47.5 (hey didn't we see that number somewhere
23 | P a g e

else?). This is the angle between your zenith and Polaris - see where that one is in
Figure 8. So what is the angle between the northern horizon and Polaris? Well, it's 90
from the zenith to the ground, and we already know one angle on the northern side, so
the height of Polaris is just the difference, 90-47.5 = 42.5. Polaris is 42.5 above the
horizon, which is what I said in the first place! If you are completely confused by this,
then you better read over this section again before we go to the next one, where you
will figure out some other angle problems.

Figure 8. The angles between different
points on your meridian.
How to work out those annoying angle
problems
Let's look at another problem, similar to the
one gone over in the previous section
dealing with how to use the declination
values of objects. All problems like these
have 3 parts -

A latitude
The declination of an object (how far it is from the Celestial Equator)
The height above the northern or southern horizon
The diagram that helps you figure out the problem, like that shown in Figure 8, shows
only the northern and southern horizon. Why not the eastern or western horizon? You
might want to think of these problems as meridian problems, since that is where we
are looking at the objects in the sky, when they are on your meridian (I hope you
didn't forget what the meridian is, because if you did, you may want to look it up
again). Anyways, the diagram just shows the angles along the meridian going from
the ground, the horizon, and extending upwards. You really need to get your brain
around these problems, since they will definitely be on the test.
Let's go back to the set up in Figure 8, the way that the sky is oriented along your
meridian as viewed from Cedar Falls. How high above the horizon would an object be
if it were on the Celestial equator? That's an interesting question. First, you should
remember that the Celestial equator is where the declination = 0, and that location is
shown in Figure 8. Something in that direction would be located 47.5 above the
ground when it is on your meridian (also the greatest height above the horizon).
24 | P a g e

That was pretty easy; what about an object that is located at a declination of 33 N?
How high above the ground is it? You have to first determine where that location is in
the diagram. Here's one of the rules you'll need to remember (don't worry I'll
summarize these all later): Declination is measured from the Celestial equator. So
if an object has a declination of 33 N, there must be an angle equal to 33 between the
object and the Celestial equator. But on which side of the Celestial Equator is the
object? Is it to the left or right of the Celestial equator? Which side has the label
"northern horizon"? That tells you which way is north, so you want to put the object
to the north (or, in this case left) of the Celestial equator. This is shown in Figure 9.
What is the height? 33 + 47.5 (the angle between the object and the Celestial equator
added to the angle between the Celestial
equator and the ground) = 80.5!
Figure 9. The location of an object along
your meridian which has a declination of
33 North as viewed from Cedar Falls.
Here's one that is a bit more confusing -
what would the height of an object be if it
has a declination of 80 N? If you follow the
preceding logic, you'd find that it is 52.5
above the northern horizon. Wait a second,
isn't 80 N just 10 away from the North
Celestial pole? Yes, 90-10=80. Okay, so the object is 10 from the North Star, Polaris.
That's fine. Now here's the tricky bit - how can you be sure of which side of Polaris
the object is at? I hate to admit it, but this is a trick question, since there are actually
two correct answers. How is that possible? You remember those stars that never set,
circumpolar stars? That means at one time they are found above the North Celestial
Pole and at another time they can be found on the other side, or below the North
Celestial Pole, since they have to go in a circular path around Polaris. This is shown in
Figure 10. I'll try to avoid these kinds of questions, since they are confusing.
25 | P a g e

Figure 10. Circumpolar objects can be
found at two locations along your meridian.
Let's try another one. What would be the declination of an object that is located 30
above the Northern horizon when you are located at a latitude of 15 South. Work on
this for a while and once you have an answer, or if you get completely stumped, just
keep reading.
What did you get for an answer? Did you even get an answer? Let's see how you solve
this - and here is where I'll summarize the way that you tackle these problems.
1. Make a drawing showing your horizon and
your position.
2. Your latitude = the declination of your zenith -
so you know right away one of the declination
values of the sky (in this case you would put 15
S right over your head) - Figure 10a

3. Figure out where the Celestial equator is - is it
to your north or south? You are in the southern
hemisphere (you have a southern latitude), so the
equator is north of you. An easy way to
remember this is to see what the direction of your
latitude is; the equator is in the opposite
direction. The Celestial equator is to the left
(north) of your zenith location - Figure 10b
shows this.

26 | P a g e

4. Now put in any of the angles that you can. You
know that the zenith-Celestial equator angle is
15, since that is how it is defined (remember,
declination is measured from the Celestial
equator). What would be the angle between the
Celestial equator line and the horizon? 90-15=75
left over - remember subtract from 90, not 100!
These angles are shown in Figure 10c.

5. Now you can do anything. If you want to put
in the South Celestial pole and those angles you
can, but here you don't need them. Let's put the
object in the picture. What was its location? In
case you forgot, it is 30 above the northern
horizon. Is it above the line marking the Celestial
equator? No, since the Celestial equator is 75 up
and 30 is much lower than that. Now you can
draw it in, and you better put it below the
Celestial equator. You know the angle between
the object and the horizon, since that is what
you're given, so put that in as well. This is shown
in Figure 10d.

6. What is its declination? How is declination
measured or determined? If you haven't figured
that out, I'll say it again - declination is
measured from the Celestial equator. How far
from the Celestial equator is the object? Hmm,
the Celestial equator is 75 up, the object is 30
up, so the difference is 75-30=45. So the
declination is 45. Is that correct? No it isn't! I
forgot something. Declination is direction
dependent - you have to say North or South. How
do you determine if it is North or South? You
need to determine if it is North or South of the
Celestial equator. It's to the left of it, so that
means it's North of the Celestial equator. The
final answer is that the object is at a declination
of 45 N. Figure 10e shows the final setup.

If this is what you got, great! If not, well, you should work on it some more. There are
more examples and practice problems available at this link.


Now that you've read this section, you should be able to answer these questions....
27 | P a g e

How do the stars move over the course of an evening?
What is a celestial sphere used for? Is it realistic?
What does declination define?
What does Right Ascension define?
What can the small angle formula tell you? What do you need to know before
you can use it?
What does the zenith indicate?
What does the meridian indicate?
How do you determine the height of an object on your meridian above the
horizon?
How does the declination and your latitude impact the height of objects in the
sky?
Motion of the Sun

What's covered here:
How does the Sun appear to move over the course of a year?
What defines the seasons?
What causes the seasons?
What is really going on - does the Earth move or does the Sun move?
What is the zodiac, and does it help you predict the future? (NO!!!!)
What are the phases of the Moon?
What causes tides?
What are eclipses?

Once people figured out how the stars moved - or thought they did - they could turn
their attention to the next object - the Sun. Unfortunately, its motion isn't easy to
understand. The Sun's path varies over the course of the year. Sometimes it rises in the
northeast, and sometimes it rises in the southeast. Only on two days does it rise
directly in the East and set directly in the West. These special dates are known as
the Equinoxes. To give you their full names, they are the Vernal Equinox, which is
around March 21, and the Autumnal Equinox, which is around September 21. You
may recognize these dates as the beginnings of the seasons of Spring and Autumn.
These dates - the Equinoxes - have nothing to do with the weather; they have to do
with the location of the Sun relative to the Celestial Equator.
Now for the rest of the year, the Sun's path and its rising and setting locations vary. As
seen from Iowa, during the winter the Sun rises in the southeast and sets in the
southwest. In the summer it rises in the northeast and sets in the northwest. There are
28 | P a g e

two days when the rising and setting locations are at their most extreme (furthest north
or furthest south). These days are also the dates that the Sun travels a path that is also
an extreme - very long and high above the horizon or very short and low to the
horizon. These are the days known as the Winter Solstice, which occurs around
December 21 (shortest day), while the other is called, oddly enough, theSummer
Solstice, and it occurs around June 21. Of course, you know these days as the
beginning of Winter and Summer. Like the dates of Equinoxes, they have really
nothing to do with the weather, but with the position of the Sun relative to the
Celestial Equator.
One thing you have to remember about the Sun is that it makes it very difficult to see
anything else in the sky when it is out - even though the stars and planets are out
there, the brightness of the Sun is so overwhelming that you don't have much of a
chance of see them until the Sun sets. Which stars would be visible? Which
constellations would be visible if we could turn off the Sun? That depends upon what
time of the year you look. If you were to turn the Sun down so that you could see the
stars at the same time that you could see the Sun, you would notice that the Sun
appears to move slowly toward the East from one day to the next - it moves about 1
each day. In about a month it has moved 30 to the East relative to the stars; in four
months, it will be about 120 east of where it started; and after one year, it will have
gone about 360. That means it is back where it started from, since there are 360 in a
circle! This also explains why we see different constellations in different seasons. As
the Sun moves slowly in front of various constellations, those constellations are no
longer visible since they are too close to the Sun, but constellations far from them are
visible, since they will be visible when the Sun has set or before it rises. Since the Sun
appears to move relative to these stars, it will gradually cover up other stars, and other
stars that were previously not visible will again be viewable as the Sun gets further
away from them.
Actually, the folks in the old days could figure out where the Sun would be relative to
the stars by looking at the stars which were visible when the Sun set. They knew
which constellations the Sun covered up and when they were covered up (which time
of the year they were or were not visible). If you were to map out the path of the Sun
relative to the stars, you would see it as a curved line on the Celestial Sphere. Take a
look at Figure 1 to see the path relative to the Celestial Equator. This image is of a
flattened out Celestial Sphere, and the dates mark the locations of the Sun relative to
the stars over the course of the year.
29 | P a g e


Figure 1. The path of the Sun, the ecliptic, shown relative to the background stars and
the Celestial Equator (dec=0). The location of the Sun on the equinoxes and solstices
is indicated. Some declination values are also indicated.
As is apparent, the path of the Sun is curved relative to the Celestial Equator. There
are times during the year when it is north of the Celestial Equator and other times
when it is south of it. The declination of the Sun varies throughout the year. (Of
course, its R. A. changes as well, becoming slowly larger each day as the Sun moves
eastward relative to the stars, but we'll pay more attention to the declination). On the
days of the Equinoxes, the Sun is right on the Celestial Equator, so it has a declination
of 0, and on the Solstices, it has the most extreme value for its declination, 23.5 N on
the date of the Summer Solstice and 23.5 S on the Winter Solstice. The Solstice dates
mark when the Sun is at its greatest distance from the Celestial Equator.
The path the Sun appears to make amongst the stars is known as the ecliptic. Just like
the Celestial Equator, it would make a large circle on the Celestial Sphere. In fact the
ecliptic is a big circle that is tilted 23.5 relative to the circle made by the Celestial
equator. This is shown in Figure 2.
Figure 2. The location of the Ecliptic on the
Celestial Sphere.
Why is it like this? Why does the Sun travel on its
own unusual path? Here is where I get to shatter all
of your delusions - there is no Santa Claus! Oh,
wait, that wasn't the delusion I was supposed to
shatter. No, the concept I get to confuse you with is
concerning all these motions I have described so
far.
30 | P a g e

The stars do not move across the sky in approximately 24 hours.
The Sun does not move across the sky in approximately 24 hours.
The Sun does not travel amongst the stars and moves slowly eastward each
day.
If they aren't moving, what is? WE ARE! Almost all the motions of the sky are due to
motions of the Earth. The main motion is the rotation of the Earth. We spin around
once in approximately 24 hours - that is why we see the stars and Sun appear to move
in about 24 hours. What about the Sun moving eastward relative to the stars over the
course of the year? Again, we are doing it - we are moving around the Sun, and it
takes one year for us to get back to where we started. This motion results in our seeing
the Sun in front of stars of different constellations over the course of the year. Figure
3 illustrates this concept. I'm not saying that nothing in the Universe moves except for
the Earth - it's just that the Earth's motion is so large, so close, and so obvious to our
senses that it has the greatest influence on how we see the sky. As you'll see,
practically everything in
the Universe moves.
Figure 3. The apparent
motion of the Sun amongst
the stars is due to the
motion of the Earth
around the Sun and our
changing viewpoint. The
stars that we would see
behind the Sun in January
would be different from the stars we would see behind the Sun in February, March,
and every other month, since we are changing the location from which we view the
Sun.

If our motion about the Sun makes it look like the Sun is in front of different stars
over the course of the year, why is the apparent path of the Sun, the ecliptic, tilted
relative to the Celestial Equator? Again, our bad - we're the ones that are tilted. If you
hold your head to the side and walk around all day like that, and if you don't know
you have your head tilted, you might think that the entire world is at an angle.
Since the Earth is tilted, there are times when the tilt has the Sun located north of the
Celestial Equator and other times when the Sun is located south of the Celestial
Equator. If the Earth were not tilted then the Sun would be always located on the
Celestial Equator - which would be pretty boring. The angle of the tilt, 23.5, is an
important number (remember seeing it in values for the Sun's declination?). Just stay
tuned, you'll see it again.
31 | P a g e

The Earth is tilted over; is that such a big deal? You're darn right it is, because without
this tilt, there would be no seasons. As the Earth goes around the Sun, the tilt of the
Earth causes different parts of the Earth to receive different amounts of sunlight.
During the months of May, June and July, the northern hemisphere of the Earth is
tilted more toward the Earth than the southern hemisphere. That gives the northern
hemisphere a greater amount of heat and results in higher temperatures and more
sunburns. The opposite is true during November, December and January, when the
Northern hemisphere is tilted away from the Sun. Check out Figure 4 to see the
situation.
Figure 4. The tilt of the Earth and its motion
around the Sun make it appear as if the Sun
is going further to the north (north of the
Celestial Equator) or south (south of the
Celestial Equator) over the course of a year.
Here is an animated image showing how the surface of the Earth gets different
amounts of sunlight depending upon the time of year and the latitude. Each image is
taken about one week apart at the same time of day, and since the curved surface of
the Earth is flattened down in the image the lighting pattern is rather strangely shaped.
You should pay careful attention to the date of each image and how some parts of the
Earth are in total darkness some times during the course of the year (the polar
regions). If you want to see how the sunlight falls on the surface of the Earth over the
course of a single day, just click here. In this case, the images are about one hour
apart.
On the Summer and the Winter Solstice (around June 21 and December 21
respectively), the Sun reaches its most northern and southern declinations. People who
live at a latitude of 23.5 north and south of the equator will have the Sun at their
zenith at noon only on that day of the year (June 21 or December 21 depending upon
whether they live at 23.5 north or south). You may have noticed these latitudes
marked on maps because of their special relation to the Sun - these are the Tropics.
They are the Tropic of Capricorn, located at 23.5 S, when the Sun is at the zenith on
about December 21, and the Tropic of Cancer, found at a latitude of 23.5 N, where
the Sun is found at the zenith on about June 21.
These two lines also limit the locations where the Sun is visible at the zenith. Only
between the latitudes of 23.5 N and 23.5 S would you ever have the Sun directly
overhead. Since the declination system is an extension of the latitude system, the
Sun's declination can only have values within that range as well, between 23.5 N and
23.5 S.
32 | P a g e

Here is an animation of the Sun relative to the stars. Each image is seven days apart so
that you are seeing how far the Sun moves in a week's time relative to the stars. You'll
see that it moves toward the left (East), and sometimes it goes further to the south and
sometimes it goes further to the north. These are the stars and constellations that the
Sun would appear to be in front of at some time during the year, if we could see the
stars located behind the Sun in the daytime. You may notice that many of the
constellation names are familiar to you; gee, I wonder from where?
We have the Sun appearing to move amongst the stars (but you know that it is really
due to the motion of the Earth around the Sun) along the ecliptic. This path, the
ecliptic, goes through various constellations in the sky. In fact, it goes roughly through
12 rather special constellations. These constellations are sort of set apart from the rest
of the constellations because of this aspect and we refer to them as the zodiac. This
brings up a rather interesting aspect of astrology - determining what your "sign" is. If
you were to look in a newspaper at a horoscope column, you can determine what your
"sign" is by the date of your birth. What does the date have to do with it? The date is
supposed to correspond to the location of the Sun relative to the stars on the day you
were born. Since the Sun appears to move, the stars it is in front of will change
gradually over the course of a year. Whatever constellation (of the zodiac) the Sun
was in front of on the day you were born will tell you what your "sign" is. This may
seem like a straightforward explanation, but it is not very correct.
If you were born today, what would your sign be? This should be the constellation
that the sun is located in today. If the date is August 30, then a person born today
would be of the "sign" Virgo, at least according to what it says in the paper. That
means the Sun is in front of the stars of the constellation of Virgo on August 30. Is it?
If you were to go to the star charts and check, you'd find that the Sun is in the
constellation of Leo, not Virgo. In fact, it won't be in Virgo for some time. By the
time it is in Virgo, the astrology columns in the newspaper say that the Sun is
supposed to be in Libra. The Sun is not in the constellation that corresponds to the
dates of the horoscope signs. The Sun is usually located in the previous "sign's"
constellation. All of the signs are off by one. All that time you thought you were a
Scorpio, you were actually born under the sign of Libra - of course, if you believe any
of that astrology crap to begin with, you're in more trouble than I can believe.
Why is the system all screwed up? Is this any way to do business? Of course it is not,
and to be honest, the system was originally set up correctly. When this system was
initially set up, the Sun was in the correct "sign," so that on August 30, the Sun was in
the constellation of Virgo (unlike how it is now). Things have changed since the
astrological signs were first set up by the Babylonians in about 2000 BC. The entire
system has been gradually shifting due to the wobbling of the Earth. The Earth is
wobbling? Yes, the Earth sort of acts like a spinning top - and like a spinning top, it
33 | P a g e

wobbles. We shouldn't say wobbles, since that doesn't sound too scientific. The
term precession is used to describe the wobbling - and it does sound more scientific.
The Earth is slowly precessing, and the pole of the Earth, or the axis of rotation, will
point one way, then another, then back again and so on, just like a wobbling toy top.
This is mainly due to the Moon's gravitational influence.
Originally, on the Vernal Equinox in 2000 BC, the Sun was located in between Aries
and Pisces, so for the next month everyone should be an Aries (March 21 - April 21).
Due to the Earth's precession, the Earth has wobbled so much that the celestial poles
and equator are aligned with different parts of the sky. Things got screwed up, since
the location of the celestial equator defines the location in the sky where the
equinoxes occur. Now, on the Vernal Equinox, the Sun is not at the same location
relative to the stars it was in when the system was set up. It's actually in Pisces and
getting closer to Aquarius - which, by the way, is sort of the basis for that song "Age
of Aquarius," but you might be too young to remember that golden oldie. The
beginnings of the seasons have also slowly changed relative to the stars, since the
equinoxes mark these. It used to be that the first day of spring occurred when the Sun
was in front of the stars of Aries; now it occurs when the Sun is in front of the stars of
Pisces.
I've put together two little movies that show how the changing orientation of the
Earth's poles changes the coordinates. The first movie shows how the location of the
celestial pole changes over time, from about 5000 BC to about 10,000 AD. The
celestial pole would be at the center of the circle for the coordinate grid. This movie
only shows part of the entire precession cycle (about half), but it is still enough to see
how the "North Star" changes over time. We're actually kind of lucky to have the
current star at that location now, since for most of the time there isn't a very bright star
near the north Celestial pole. The second movie shows the effect of the changing
alignment on the declination and Right Ascension system. The program is set up to
show where the Sun is on the first day of spring - which is how we set the Right
Ascension value, since that is where it equals 0, and on this date the Sun's declination
is also 0. The program causes the grid to tilt a bit, but you'll see that the Sun remains
on the 0 RA and 0 Dec location for each time step, which is 100 years. Our grid of RA
and dec must also change relative to the stars, since we keep changing our alignment
to the Sun due to precession. While it takes a long time to make the shift noticeable, it
was known to ancient astronomers, though of course they didn't know the cause.
Nowadays if we want to point our telescopes with very precise coordinates we have to
calculate the effects of precession on the coordinates - which is pretty small from year
to year, but is important if you want very precise coordinates.
The tilt of the Earth doesn't change much (currently at 23.5 ), but the direction that
the pole points changes. The North pole star (Polaris) will not be there all the time - in
34 | P a g e

a few hundred years the current north star will be just another star in the northern sky,
since it will not be located at the North Celestial Pole. In the past, other stars would
have been called the Pole star, since they were closer to the North Celestial Pole than
Polaris was. During your life time, Polaris will be the North Star, since the wobbling
is pretty slow. One precession takes about 26,000 years. Figure 5 shows a simplified
view of the precession of the Earth.
Figure 5. The direction that the Earth's pole points
changes slowly so that in the far future it will be
pointing to stars such as Alderamin, Vega, Thuban, and
eventually again Polaris.
Length of a Day - Solar versus Sidereal
How long does it take the Earth to spin around exactly
once? We could figure that out by timing how long it takes something in the sky to get
back to its original position from one day to the next. If we time the motion of the
Sun, we see that it takes almost exactly 24 hours for the Sun to get back to where it
started from one day to the next. I guess that answers it, right? Before we jump the
gun, let's time another object - a bright star, for example. How long does it take a star
to get back to the same place in the sky from one day to the next? Does it take 24
hours for one complete rotation? No it doesn't. It takes 23 hours and 56 minutes. Big
deal; that's almost 24 hours; there is only a four minute difference; does it really
matter? You bet your banana skin it matters!The basic upshot is stars rise or set four
minutes earlier each day. If a star rises tonight at 8 P.M., it will rise at 7:56 the next
night, then 7:52 the night after, and then 7:48 the next night. A week after the first rise
time, it will rise 4 x 7 = 28 minutes earlier (7:32). In one week, a star will be rising
about half an hour earlier - that's a pretty big difference, so don't ignore those four
minutes.

Why is there a four minute difference? Which of these values tells us what the
rotation period of the Earth is? Remember, it is the spinning of the Earth that causes
the observed motions of the Sun and the stars over the course of the day (or night) -
but there are two different time spans here - which one corresponds to the rotation
period of the Earth?

Believe it or not, it is the stars, not the Sun, that determine the amount of time for one
rotation of the Earth. While all clocks on the Earth are based on the 24 hour time scale
of the Solar Day, it is the more subtle Sidereal Day (or "star" day) that tells us how
fast the Earth is spinning. It takes the Earth 23 hours and 56 minutes to complete one
rotation.

35 | P a g e

Why is there a difference between a Solar day and a Sidereal day? The cause is the
motion of the Earth, in this case our orbital motion around the Sun. To illustrate
what's going on, follow the stick in Figure 6. It starts out one day pointing directly at
the Sun (at noon) and at a very distant star (a star way off to the right).
Figure 6. The time it takes for a position on the Earth to line up
with a distant star (way off to the right) is 23 hours and 56
minutes. However, the Sun will not be lined up with the position
on the Earth, and an additional four minutes are needed.
After 23 hours and 56 minutes, the Earth will have not only made one complete
rotation, but will have also moved in its orbit. The stick will no longer be pointing
toward the Sun, but it will again be pointing toward the star - this tells us that the
Earth has made exactly (no more, no less) one rotation. The time on our watches is
11:56 AM - NOT NOON! - since it is 23 hours and 56 minutes from the previous day.
If you want to see the stick pointing again at the Sun, you must wait four more
minutes for the Earth to spin a little bit further around. When that happens, the time
will again be noon.

You might be a bit amazed at how I was able to easily draw up the motions of the
Earth and such so quickly, but how did I know which way it was going? There is a
rule about how things in the solar system move and you can use it to draw similar
diagrams. All major motions in the solar system are in a COUNTER-CLOCK
WISE direction when observed from above the North Pole. This includes the motions
associated with the Earth, the Moon, the orbital motions of the planets, and most of
their rotation motions as well. There are of course some exceptions, but they aren't
that common. If you have to quickly draw any solar system motions, you'll know
which way the stuff is moving - again, there are a few exceptions and I'll tell you what
they are if necessary.
The Moon
After figuring out how the stars and the Sun move, it is time to tackle the next object -
the Moon. The motion of the Moon is more complex; it doesn't follow exactly the
ecliptic or the celestial equator, but does make a path around the Earth that is similar
to each of those paths. What sets the Moon apart from the other objects is the fact that
its appearance changes - it goes through phases. The phases of the Moon take 29.5
days to go through an entire cycle.
The phases occur in a very predictable sequence. Here is the order of the
phases - New (when you can't see the Moon - it's all dark), Waxing
Crescent, First Quarter (when you see the right half lit), Waxing
36 | P a g e

Gibbous, Full (when you see the entire lit surface), Waning Gibbous, Third Quarter -
also called Last Quarter (when you see the left side lit up), Waning Crescent, and back
to where we started, New. A picture of the phases is shown in Figure 7. It takes about
one week to go from one major phase to the next - by major phase I mean New, the
Quarters and Full. If the Moon is New today, it will be a First Quarter Moon in about
one week, and a full Moon two weeks from today. The fact that it takes about 29.5
days to go through the cycle is the reason there are about 30 days in a month, since
many ancient societies used the Moon as a time keeper.

Figure 7. The phases of the Moon as seen from Iowa when the Moon is high in the sky
(on your meridian).
What causes the phases? To figure that out, you need to look at the interaction of the
light source (Sun) and the alignment of the Moon with the Earth. The Moon will have
a certain phase depending upon two things -
1. The location of the Moon in its orbit about the Earth
2. The location of the Sun relative to the Earth and the Moon at that time
Don't forget about these two things.
The Quarter Moons occur when the Sun and the Moon are 90 degrees apart in the sky
as viewed from the Earth. The New and Full phases occur during times when the
Earth, Moon and Sun are in a straight line. Figure 8 is a composite of the various
phases and the location of the Moon in the sky. Remember, it takes about a week for
the Moon to go from one major phase to the next, so that the view you see during one
evening isn't too much different from the view you see the next night. You may have
seen the Moon when it is close to the Full phase and it may appear to you to be Full
for several days, while technically it is only Full at the time it is in a line with the
Earth and the Sun. Also, the way that the Moon is illuminated gives us the view we
see - when most of the lit surface is turned away from the Earth, we see only a small
crescent; when most of the surface is turned toward the Earth, we see the gibbous
phase Moon.
37 | P a g e

Here is a little java program showing just one phase at a time. You can see how the
Moon looks to you in the sky depending upon where it is located in its orbit about the
Earth.
Figure 8. The phases of the
Moon shown at their
locations relative to the
position of the Earth and
the Sun (off to the right).
The phase of the Moon is
determined by its location
relative to the Earth and
Sun. The right side of the
Moon is the only part that
is illuminated, since the
Sun is off to the right. The
phase that we see depends
upon how much of that side
of the Moon is visible from
the Earth, which depends upon where the Moon is in its orbit about the Earth.
At what time of day does the Moon rise and set? When does it cross your meridian
(this is another way of asking "When is it highest in the sky?")? Well, that will
depend on the phase. The phase will determine the location of the Moon relative to the
Sun. Your location relative to the Sun will determine what time of day it is.
Remember, we base our time system upon the location of the Sun - so if the Sun is on
your meridian (high in the sky), then it is Noon at the part of the Earth you are located
at. A person at a location on the opposite side of the Earth from you would be looking
at their watch and noting that it is midnight. Of course, daylight savings time, and the
rather irregular way that time zones are set up, may actually mean that it is not exactly
noon or midnight, but we'll make it simple and assume that it is.
Take a look at the set up shown in Figure 9. If you were at the point labeled "noon,"
the Sun would be high in the sky, but the Moon would be on the horizon - it would be
rising. If you were located at the location labeled 6 PM, the Moon would be high in
the sky, and the Sun would be on the horizon; in this case, this is also referred to as
the time of sunset. A person at the midnight position would see the Moon on the
horizon, setting. Remember, as viewed from above the North pole, all motions are
counter-clockwise. The Earth will be spinning around while the Moon remains in
about the same position (the first quarter phase location). All day, anyone seeing the
Moon would see a First Quarter moon. To see the Moon you must be located on the
side of the Earth that is toward the Moon, so the person located at the 6 AM spot
38 | P a g e

wouldn't see the First Quarter Moon, and neither would anyone located at a position
corresponding to 1 AM, 2 AM, 3 AM... all the way until Noon. It would take a day or
two for there to be a noticeable change in the Moon's rising and setting times.
Follow this link to see a little java program showing how the different phases of the
Moon would appear to an observer, when they would be seeing them and what the
Moon would be doing. You can also change the location of the observer to get
different times of day.
Figure 9. The location of the
First Quarter Moon allows you
to determine when it rises
(noon), sets (midnight) and
when it is high overhead (6
P.M.).
In diagrams like Figure 9, you
would first have to put the
Moon in the appropriate
location relative to the Sun and
the Earth for its current phase,
then what you see depends upon where on the Earth you are, and your time depends
upon where you are relative to the Sun. A few basic rules to follow for the Moon-
rising-setting problems -
1. New Moon is in line with the Sun, so it does everything exactly when the Sun does
its stuff - rises at 6 AM, sets at 6 PM, and on the meridian at Noon.
2. Full Moon does everything at opposite times relative to the Sun - rises at 6 PM, sets
at 6 AM, and is highest (on the meridian) at Midnight.
3. The Quarter phases are located at a 90 degree angle relative to the Sun.
Once you figure these things out, it is pretty easy to figure out what the Moon is doing
and when it is doing it.
Follow this link to see a little java program that quizzes you about the what the Moon
is doing depending upon the set up. You can determine the answer or just keep
guessing until the program tells you the answer. Either way, you'll eventually get the
correct answer and a diagram showing you the set up.
How long does it take the Moon to orbit once around the Earth? It takes about 27.3
days. Why not 29.5 days (the time for the phase cycle)?
39 | P a g e

Again, it has to do with the fact that the Earth is moving around the Sun. Take a peak
at Figure 10. It shows the variation from one Full moon to the next. Remember, the
Moon has to be in a straight line with the Earth and Sun for it to be Full. It starts out
lined up with the Sun, but after 27.3 days, the Moon will have made one complete
orbit of the Earth (again be located to the left of the Earth). At this time is it Full? No,
because it is not in a perfect line with the Sun. You have to wait about 2 more days for
it to again be aligned with the Sun and for it to be Full again.
If you spend a couple of days watching the Moon relative to the stars you'd see that it
moves about 12 degrees each night (since it has to go 360 degrees in about 30 days).
You may not notice the motion of the Moon relative to the stars over the course of an
evening since it is sometimes difficult to see the stars close to the Moon, but you can
certainly note the fact that each night it rises about 50 minutes than the previous night,
so the Moon's position relative to the stars
has changed.
Figure 10. The Moon makes one complete
orbit of the Earth in 27.3 days, but it will not
be again Full until a total of 29.5 days has
passed.
One orbit of the Moon takes 27.3 Days. This
would be the Moon's Sidereal Period since
it is the time for the Moon to be back in the
same location relative to the stars, and this is also the time for one orbit. How long
does it take for one rotation (spin) on its axis? Does the Moon actually spin on its
axis? If you said "no," then you're wrong. The Moon does spin on its axis, but it does
it in 27.3 days. That's the same amount of time for one orbit - what does that mean? It
means that one side of the Moon always faces the Earth - that the Moon has one
side tidally locked with the Earth.
If you still think that the Moon doesn't spin around think about this - if it didn't, we'd
see different sides of the Moon, not just always the same view. This is shown in
Figure 11. If the Moon did not rotate, it would always have one side pointing in the
same direction, as is shown by a line on it. That line would always be pointing in the
same direction in space, but on the Earth we would see different sides of the Moon as
it goes around the Earth. That's not what we see - we can only see one side of the
Moon from the Earth, since that one side is always pointing toward us.
40 | P a g e

Figure 11. The Moon makes one
complete orbit in the same
amount of time it takes to make
one rotation. If it didn't rotate,
we'd have the situation on the
left, where different sides of the
Moon would be visible from the
Earth as it goes around the
Earth. One side of the Moon is
always facing the Earth, since the
orbital period is the same as the rotation period.
The time for it to complete a cycle of phases is 29.5 days. I don't think we can just call
it that, can we? No, of course not; we'll have to give it a more "scientific name." We
refer to this time as the Synodic Period. This is just the time it takes for the same
Earth-Sun-Moon alignments to occur again, so it is the time for the Moon to go from
one Full Moon to the next Full Moon, or the time it takes to go from one first quarter
moon to the next first quarter moon - either way, it is 29.5 days.
Tides
You may have noticed that I used the phrase "tidally locked" above. What's that all
about? Living in Iowa, you are probably not too familiar with the phenomena of tides,
and just to be clear, I'm not referring to laundry detergent - what I'm talking about is
how the Moon and the Earth pull upon one another and how the consequences of
those "pulls" are predictable. Tides usually refer to the water levels of large bodies of
water like the oceans which change near the shore due to the pull of the Moon (and
the Sun) on the water. While we haven't gotten to gravity yet, you can at least
appreciate the fact that a large nearby object like the Moon has a pretty good
gravitational influence on the Earth - it is pulling on the Earth all the time, but the
only thing that we see responding to the pull is the water. This gravitational pull
causes a bulge in the water in the oceans. This bulge follows the motion of the Moon.
The bulges actually occur on both sides of the Earth. The side that's closest to the
Moon feels the strongest pull and bulges out, and the side furthest from the Moon
feels a lesser pull and is sort of left behind (since the center of the Earth also feels the
pull, more than the water on the furthest side of the Earth). This is shown in Figure 12.
Figure 12. The Moon's pull on the water and
the Earth produces bulges on the two sides
of the Earth. The degree of the pull is shown
by the arrows, with the side nearer the Moon
41 | P a g e

having the largest pull, and the side further having the smallest pull.
Since it takes only 23 hours and 56 minutes for the Earth to rotate, coastal locations
will pass through the high water bulges (have high tides) two times each day.
Remember, the motion of the Moon is much slower, so the Earth actually rotates
through the tides.
The Sun also has an influence on the tides, but since it is further away, it doesn't pull
as strongly. However, when both the Moon and the Sun are pulling along the same
axis, the tides are highest. These tides are known as spring tides, and they have
nothing to do with Spring, but they occur when the Moon is either Full or New. When
the Moon and Sun are 90 degrees apart from one another (during the Quarter phases)
the tides are flattened out - these are the neap tides. Tides are very important for
coastal regions, since in some places, you can't get a ship out of the harbor if the water
level is too low. A storm that occurs during high tide could cause coastal flooding
(this is often seen during hurricanes). Knowing when the high tides occur is very
important for many places for a variety of reasons.

Tides also have an important side effect on a planet (and its moon). The rotation rate
of the Earth is changing - it is slowing down - since all of this pulling and tugging is
going on. Now you know why it seems like some days last forever (well, not really).
The slow down has already happened with the Moon; that's why only one side of the
Moon faces the Earth. In the past the Moon spun around a lot faster. The Earth's
strong pull on it has locked one side of the Moon to always face the Earth. Another
side effect of this slow down of the rotation of the Earth and the Moon is that the
Moon has been getting further and further away from the Earth very slowly. After a
few million years the length of a day on the Earth will be very long and the Moon will
be visible from only one side of the Earth (and be very far away). We have a long way
to go before that happens.
Eclipses
Eclipses are events that occur when one object blocks another, usually resulting in
something getting darker or appearing fainter than before. The path of the Sun, the
ecliptic, is so named because that is where eclipses are seen to occur. The Sun, the
Earth and the Moon all participate in two different main eclipse types - Lunar
Eclipses and Solar Eclipses.

Lunar eclipses occur only when the Moon is Full and it is located on or very close to
the ecliptic. In this case the shadow of the Earth falls upon the Moon, making it dark.
Why don't we have lunar eclipses during each Full Moon? The orbit of the Moon is
slightly tilted with respect to the Earth's orbit about the Sun, about 5 degrees. For
42 | P a g e

there to be an eclipse the Moon has to be at the point where the planes of the orbit of
the Moon and Earth intersect - the nodes. This also explains why we don't have
eclipses every month - the orbit of the Moon is not aligned exactly with the ecliptic;
sometimes it is above it, sometimes below it. To further complicate things, the orbit of
the Moon "wobbles" with a period of 18.6 years. This aspect means that eclipse paths
repeat with a period of 18.6 years. When you view the solar eclipse maps at the end of
this set of notes, you will see the repeating eclipse paths - those that have the same
shape.
Figure 13. The tilt of the
Moon's orbit means that most
of the time it isn't lined up for
an eclipse to occur.
Is the Moon entirely dark
during a lunar eclipse? No, you can still see it, though it is dimmed and often colored.
Why? The Earth has an atmosphere which tends to bend light around the Earth, and
this falls upon the Moon. The light does get discolored, though; often a red, orange or
brown color is seen. This really scared the heck out of folks in the old days when they
didn't know what was going on, which was most of the time.
Figure 14. A multiple exposure
image of a lunar eclipse. When the
eclipse first starts out the
brightness of the Moon is so great
that only short exposures are used.
The umbra, which is the darkest
part of the shadow, doesn't appear
to have any color in these images.
Only at mid-eclipse is the Moon
darkened enough so that the color
of the umbra is visible. Eclipse
photograph copyright 2000 by Fred
Espenak courtesy of www.MrEclipse.com.
There are two regions of the shadow, the umbra and penumbra. These correspond to
the darkest part of the shadow and the not completely dark part respectively. These
are shown in Figure 15. If you were on the Moon, and looked back toward the Earth
(and the Sun), you would see the Sun blocked out if you were in the umbra, while if
you were in the penumbra, some part of the Sun's surface would still be visible to you.
43 | P a g e

Figure 15. The shadows, umbra
and penumbra, cast by the Earth
during a lunar eclipse. The Moon
is experiencing a total lunar
eclipse here since it is in the
umbra - the darkest shadow.
Lunar eclipses can be full - the Moon passes completely through the Earth's umbral
shadow,partial - it passes only through part of the umbral shadow, or penumbral - it
only passes through the penumbra. The best are of course the full eclipses, which can
last for hours as the Moon traverses the entire length of the umbral shadow, while the
least exciting are the penumbral eclipses, which are really difficult to see since the
sunlight that falls on the Moon's surface is so bright to begin with. Here is a table of
lunar eclipses that will be occurring over the next few years - some of which are
visible from Iowa. Here is an animation showing how the December 2010 lunar
eclipse will look from Iowa. The locations of the umbra and penumbra shadows are
indicated. The locations of the Earth's shadows appear to vary slightly in the
animation, since the Earth is
moving as well as the Moon.
Figure 16. The shadows, umbra
and penumbra, cast by the
Earth during a lunar eclipse are
seen as circles here. The
shadows that the Earth casts
look like circles, since it is a
sphere. The Moon's path is
shown for the three different
types of lunar eclipses. The top
path is for a penumbral eclipse,
where the Moon never passes
into the umbra, but is only in
the penumbral shadow. The
central path shows a total lunar eclipse, where the Moon passes entirely through the
umbral shadow. The bottom path is for a partial eclipse, where only a part of the
Moon's shadow goes into the umbra. It is possible to have partial penumbral eclipses,
but those are so lame they aren't worth mentioning.
While not as spectacular as solar eclipses, lunar eclipses are still rather neat to see -
though they are best viewed when there are no clouds in the sky. There are usually
about 2 lunar eclipses each year, though they are not always visible from the US. To
see the eclipse you have to see the Moon at the time of the eclipse, so many people
44 | P a g e

can see one. Of course it is possible to predict the dates and times of eclipses. The
next "good" lunar eclipses visible from the US will occur on the night of December
21, 2010.
Much more spectacular are solar eclipses. These occur only when the Moon is New
and located on the ecliptic (it is on the node). Just by chance the Moon and the Sun
have about the same angular size - they are both about 1/2 degree in size. The size of
the shadow - the umbra - is very small, because the Moon can only just barely cover
up the Sun. This is illustrated in Figure 17.
Figure 17. The shadows cast by
the Moon during a Solar eclipse.
Only at the point on the Earth
where the umbra reaches the
surface would you experience a
total solar eclipse.
To experience a total solar eclipse, you must be located in the narrow umbral path of
darkness. This is often referred to as the path of totality. It is so narrow (at most only
about 300 km wide), and the motions of the Sun, Moon and Earth are rather fast, that
you will only experience at most about seven minutes of totality. During the moments
before the total eclipse, various features can be seen; amongst them is the diamond
ring effect, where it appears as if the last bit of sunshine makes a diamond ring in the
sky. Once the main surface of the Sun is covered up, the outer layers of the Sun, such
as the chromosphere and
corona, are visible.
Figure 18. The diamond ring
effect. Only a fraction of the
surface of the Sun is visible
here, but it is enough to cast a
bright light through the
valleys along the edge of the
Moon. Eclipse photograph
copyright 2001 by Fred
Espenak courtesy
of www.MrEclipse.com.
You were probably told in
elementary school "Don't
ever look at a solar eclipse -
you'll go blind!" or something like that. That's not entirely true, though you have a
45 | P a g e

pretty good opportunity to really screw up your eyes if you are really stupid. As a
general rule you should never stare at the Sun, and most of the time people don't do
that. When people hear that there is going to be an eclipse, they get pretty stupid and
think "Gosh, I'll have to use my binoculars to see the eclipse!" or "I better get my
telescope out to see the eclipse!" No, no, no, no, no! If you look at the sun with
binoculars or a telescope at any time, not just when an eclipse is occurring, you will
probably permanently damage your eyes. You should never look at the Sun with any
sort of instrument that magnifies it. Cameras are also a no-no, since their lenses can
magnify the Sun and this is just as bad as using a telescope. Looking at the Sun with
binoculars or a telescope is equivalent to frying ants on the sidewalk with a
magnifying glass, only in this case, the ants are your eyeballs. Don't do it!
Even if you don't use anything to magnify the Sun, and decide to watch it with your
sunglasses, that's pretty stupid as well. Even polarized glasses won't protect you while
staring at the Sun. You've probably glanced at the Sun for short periods of time, and
you're left with an "afterglow" of its image. If you do that too much, that "afterglow"
won't go away - ever. During eclipses people can get proper eye protection to allow
them to safely view the Sun. These filtered glasses are available at low cost and
provide a safe method of viewing the Sun. You may have heard of other things that
can be used, but not all of those things are reliable.
Figure 19. Eclipse glasses, with special filters, are
the best way to view an eclipse safely.
Now after all of this stuff about not staring at the
Sun, I have to tell you that there is a time when it is
okay to look at the Sun with a regular telescope -
during totality, when the Moon completely covers up
the Sun. At no time before total coverage and at no
time after is it safe, only when the surface of the Sun
is completely blocked from view. While the surface of the Sun is blocked, it is
possible to see the other layers of the Sun's atmosphere, the big fluffy corona and the
pinkish chromosphere. Whether you actually see these layers depends upon what the
Sun is doing - how "active" it is. People who observe and photograph eclipses are
very careful to time their viewing. Usually they find the exact time for the start and
finish of totality and use stop watches or alarms to prevent themselves from damaging
their eyes. Before totality, they have to use special cameras or telescope filters to view
and photograph the Sun safely, but when totality hits, they take these filters off and
start snapping pictures like crazy.
46 | P a g e

Figure 20. A series of images
showing a total solar eclipse.
The images are combined
together to show the eclipse in
different stages, with totality
occurring in the middle image.
Eclipse photograph copyright
2001 by Fred Espenak courtesy
of www.MrEclipse.com.
It is also possible that the
Moon's umbral shadow may not
even reach the Earth. This will
occur when the Moon is slightly further away and its angular size is slightly too small
to completely cover up the Sun. In this case, only the penumbra of the eclipse reaches
the Earth. This produces an Annular eclipse. A viewer of the eclipse would see the
Sun as a ring about the Moon. Click here to see animation of an annular eclipse. There
are also eclipses that change from one type to another. These are hybrid eclipses
which can start out as an annular eclipse and then become a total eclipse or the other
way around. You have to remember that the circumstances for a total eclipse to occur
are very precise, and often these conditions can't always be met.
An annular eclipse is a type of partial eclipse, where only part of the Sun is covered
up. During a total solar eclipse, you need to be in the path of totality to get the full,
dark eclipse. Otherwise, you will only experience a partial eclipse. Only those people
who are in exactly the right place at exactly the right time will get to see anything
exciting (weather permitting), because of the rather special arrangement for a solar
eclipse, especially a total solar eclipse. For this reason, many people will go to a lot of
trouble and money to be at exactly the right place to see a total solar eclipse.
Eclipses are amongst the most popular astronomical events for non-astronomers to
view and you can spend a lot of money traveling around the world to view various
eclipses. There are about two solar eclipses each year, but they are sometimes only
partial or annular eclipses. Often the path of totality is over oceans, so it isn't always
easy to view them. Figure 21 shows the path of some recent past eclipses and those
that are coming up over the next few decades. As can be expected, many of the eclipse
paths are located over oceans since the Earth is covered with so much water. A table
of upcoming solar eclipses can be found here. Unfortunately, you'll have to wait until
2012 for the next solar eclipse to be visible in Iowa, and even then it will only be a
partial eclipse.
47 | P a g e

If you want to wait for one to come to you, then you have a bit of a wait. The next
total solar eclipse visible from the US will be Aug. 21, 2017. Don't forget that date! It
should be a seriously fun time! Here is a computer simulation showing how it will
look from UNI on that date. To see the total eclipse, you'll have to head south.
Eclipses of any type are fairly rare since they require a specific set of conditions. And
there are quite a few things that factor into eclipses (both solar and lunar eclipses).
You have the variation in the distance of the Moon and the Sun from the Earth, and
the changing alignment of the Moon's orbit relative to the ecliptic. Eclipses listed in
the tables noted above are not very common, with only a handful of eclipses each
year. So the next time one is visible from your location, make sure you take the time
to view it - you will have to wait a long time for the next one!

Figure 21. Paths of total solar eclipses are shown (on left) and annular eclipses (on the right). If
you click on the image you'll see a larger version of the map. For the total eclipse maps, an
observer would have to be located in the dark path to experience a total solar eclipse. To see the
Sun "surround" the Moon during an annular eclipse, and observer would have to be located
along the paths shown in the map on the righ. Eclipse maps courtesy of Fred Espenak -
NASA/Goddard Space Flight Center. For more information on solar and lunar eclipses, see Fred
Espenak's Eclipse Home Page:http://sunearth.gsfc.nasa.gov/eclipse/eclipse.html


Now that you've read this section, you should be able to answer these questions....
On the Vernal Equinox (or any other season start), what is the declination of
the Sun? Where on the Earth would you have to be located to have it at your
zenith?
How does the Sun move relative to the stars?
48 | P a g e

What is the ecliptic?
Why does the Sun appear to move relative to the stars?
What causes the seasons?
Why doesn't astrology work?
What is the sequence of phases, when would they be visible, rising, setting, and
how does it change from one night to the next?
What causes tides? What is the difference between a neap and spring tide?
What conditions are needed for a lunar eclipse? A solar eclipse? Why don't
they happen all of the time?
Motions of Planets - History of Science and Astronomy

What's covered here:
What are the complex motions of the planets that need to be explained?
What assumptions were used to explain these motions?
What ideas were put forth by the ancient Greeks?
What did Copernicus, Brahe, Kepler, Galileo and Newton do (astronomically
speaking)?
What are the basic laws for planetary motion, and motion in general?
What is the law of gravity about?

How is science actually done? Do people just suddenly come up with ideas out of the
blue, and we take those ideas to be truths? Generally no. Most often you hear about
the Scientific Method which is the general method that describes how science is
done, how discoveries are made, and how we can expand our current knowledge
about how things work, all within a framework of "quality control". The scientific
method can have several steps which are comprised of the following
A hypothesis is proposed - this is the "new" idea that you have to explain
something that may or may not have been previously known or observed
The hypothesis should have specific expectations - there should be predictable,
reproduceable results.
Experiments, observations or tests of your hypothesis needs to be conducted by
yourself and others (to see if the predictions of your hypothesis give the results
you expected)
Results of your experiment, observations or tests need to be examined by others
to determine if your hypothesis is on the right track, or if it needs to be
altered/amended or even trashed
49 | P a g e

You'll notice that I didn't say "determine if your hypothesis is true" in the last item,
since you can't prove a hypothesis to be true. Why not? A hypothesis is supposed to
explain something, give a reason as to why something happens - so it is a first attempt
to answer "why". You can only support a hypothesis with experiments, and strengthen
it, but you can't prove it to be totally, completely correct. Unfortunately a lot of times
in the news they'll announce a new scientific observation or discovery that "proves" a
hypothesis (or a theory) to be correct. Such statements are badly written - you can't
prove a hypothesis to be true.
Eventually a scientific hypothesis will pass the test of time, and have support through
many observations or experiments such that it can be considered a theory. And just
like a hypothesis, a theory can NOT be proven to be true, but it can be tested,
supported, or refuted. Theories should be considered a "work in progress". Many
theories exist in science, and those that you tend to learn about today are those that
have passed the test of time, as well as many other tests and are accepted by scientists
and society.
So if a theory lasts long enough and passes enough tests does it ever get placed into
another category, such as "this is the truth"? No. Theories are always theories (unless
they are shown to be wrong, in which case they tend to be ignored).
Along with theories, scientists have laws. The best way to think of laws is as a way of
describing something. Laws can tell you how something will act or behave, but laws
can not explain why something acts or behaves in a certain way. That's actually what
a theory should do. Theories give us the "why", while laws give us the "how". One
explains, the other describes.
While the Scientific Method may seem like a rather formal way of describing how
science works, you have to remember that this isn't the way things were always done.
In the past there may not have been testing or even an attempt to test a hypothesis.
Sometimes ideas were just accepted without argument or discussion. Is that any way
to do science? Today a scientist has to do several things before a new idea or concept
is developed or becomes accepted
Gain a familiarity with the work of others (what has previously been discovered
in this area, what are others currently investigating - no need to repeat what
others have done)
Observations or experiments that can either support your ideas or require you to
rethink your ideas (possibly even trash them entirely)
Discussion or presentation to others of your ideas or experiments, often to
determine if you are on the right path or if you've missed something
50 | P a g e

Generally it can take years to work out a new or complex idea, and it requires careful
consideration and scrutiny from the scientific community before radically new ideas
are accepted. Yes, it is a long process, but it helps to maintain quality. It also shows
that science isn't perfect - it keeps looking to improve itself! With that in mind, let's
see how scientists in the past tried to determine how planets move
Perhaps the most complex motions that early astronomers had to deal with were those
involving the planets. There are some rather complex motions that have to be
explained and understood, since there was a great deal of importance placed upon
planets for astrological reasons, though some cultures had some interest in planets for
other reasons. For example, two planets, Venus and Mercury,were always seen near
the Sun. Sometimes they were seen in the morning sky before the Sun rose,
sometimes in the evening sky after the Sun set. Another annoying aspect of planetary
motions was the way that the other planets (Mars, Jupiter and Saturn) would
sometimes appear to stop in their motions relative to the stars, go backwards for a
while, and then again go in their regular paths (eastward amongst the stars). These
motions had to be not only explained but also predicted.
Figure 1. The path of Mars
over several months has it
moving slowly eastward
relative to the stars except
for a few weeks when it
moves westward (to the
right)
In order for early
astronomers to predict the
motions of the planets,
they needed to make
models of the sky which would explain its workings. Unfortunately, most of these
models made several bad assumptions -
1. Geocentric - Earth is in the middle of everything. While this may seem egotistical,
there was also the idea that the Earth was not moving, since you don't feel any motion.
It was just easiest to put the Earth in the middle. Later there were some philosophical
and religious reasons for putting the Earth in the middle.
2. The Sun, Moon, and planets have their own orbits about the Earth. Obviously if the
Earth isn't moving, everything else must be moving.
3. The stars are located on the Celestial Sphere. We already went over this, and how
the rotation of the sphere explained the motion of the stars (though of course it's
wrong, like everything else on this list).
4. All motions are along circular paths and at a steady rate. Another way of saying this
51 | P a g e

is that there were uniform circular motions. Ancient astronomers were really hung
up on the idea that stuff in the sky had to move in paths that were perfect circles, or
associated with circular objects.
Several of these aspects were accepted due to tradition; others were based upon
theological beliefs or philosophies - for example, the concept of everything moving in
a circle. Why should everything move along a circular path? In part, because the
circle is a perfect geometric shape and there was the belief that the sky should be the
place of eternal perfection - only perfect, unchanging things are up there, while the
imperfect, changing things are found on the Earth. Anything that changed, like
anything that grew or evolved, had to be found on the Earth. The lights in the sky
never change, or so they thought.

Any ways, this is what most model makers and theorists had to work with, and in
some cases, their lives depended upon it. The belief or the promotion of a system that
went contrary to these concepts could mean persecution or death in some cases. No
wonder no one wanted to go against these concepts! Here we'll look at some of the
main architects in the history of astronomy. Some of their ideas were wrong, and
some were right, but regardless of what they did, they contributed to the story in some
way or another.
Babylonian Astronomy - 17th Century BC
Figure 2. An ancient tablet with astronomical observations. The
text of the tablet is a copy, made at Nineveh (currently Iraq, but
long ago Babylon) in the seventh century BC, of observations of
the planet Venus made in the reign of Ammisaduqa, king of
Babylon, about 1000 years earlier. Image is from the British
Museum.
The whole story starts with the Babylonians. They were amongst
the earliest astronomers who recorded the motion of the planets
in the sky and produced detailed star charts. They are the ones
we have to thank for the astrological signs and that whole
business - though the Egyptians did also contribute some of the
ideas. By watching the motions of the Sun and the Moon, the
Babylonians were also able to predict events such as eclipses. Their main contribution
was their accurate and long term recordings of the motions of objects in the sky.
These detailed records outlasted the Babylonians and were later picked up by other
civilizations. I guess they get the award for thorough record keeping.
Aristotle - 4th Century BC
52 | P a g e

Image of Plato and Aristotle (384 - 322 BC) from the
Stanza della Segnatura in the Vatican, in the painting
"School of Athens" by Raphael.

You may have heard of Aristotle, and he was pretty much
a dabbler in all sorts of areas of study, not just astronomy.
In part because of his reputation, his model of the cosmos
had a very strong influence on later astronomers. His
influence was so great that this model, or at least the basic
concepts behind it, lasted about 2000 years. Aristotle did
not actually make a model, but rather described one.
Aristotle proposed that the motions in the sky could be
explained by having 56 concentric spheres located about
the Earth. A simplified version of Aristotle's model is
shown in Figure 3. The motions of the spheres would produce all of the motions of the
Sun, Moon and planets. Since a rotating sphere would make a circular motion, that
was okay with the assumption of having circular motion and the model was
geocentric, so that was okay as well. The complex motions were explained by having
the spheres moving one another. One sphere would rub against another and move it in
a certain direction, which would then move another sphere in another way, and so on.
Not too precise, but it did explain the motions in a rough sort of way.
Figure 3. Aristotle's model of the sky, made
up of many concentric spheres, with the
Earth in the middle
Another aspect of his model that was later
used by other astronomers was that by
having the Earth sitting fixed in the middle
of the solar system there would be
no stellar parallax. Parallax is the apparent
shifting of objects caused by your changing
direction of view. That explanation was as
clear as mud, so let's try a little experiment.
Take your thumb and hold it out at arm's
length. Line it up with some distant object,
viewing it with just one eye. Now switch your eyes (by this I mean close one and open
the other). What happens? Did your thumb move? Didn't you hold it steady? You
probably did, but you also did change the direction from which you were viewing
your thumb. By viewing something from a different direction (or location), the
alignment of things changes. This shifting is the parallax. See Figure 4 for what's
53 | P a g e

happening. What does this have to do with the idea that the Earth is sitting fixed in the
center of the solar system? What would happen if the Earth did move? Ancient
astronomers thought that if the Earth was moving, it would be like shifting your eyes -
at one time you would see a nearby star in front of one group of distant stars, and
when the Earth moved to a different point in its orbit, you would see it in front of a
different group of stars. This is just like the way that your thumb appeared to move
when you changed your view. Ancient astronomers could not see any parallax motion
of the stars (stellar parallax). They thought it was due to the fact that the Earth wasn't
moving - no motion, no shift in perspective, no observed stellar parallax.
Unfortunately, there is another explanation for the lack of observed parallax; can you
think of it? The ancient guys didn't but
we'll get to it later.
Figure 4. Parallax demonstration. Each
observer (A, B, C) sees a different
alignment of the wolf and the distant
clump of trees. This type of change in
how the world looks based upon where
you are located was one of the reasons
people did not believe the Earth moved
(since they could see no change in the
relative positions of stars).
Hipparchus - 2nd
Century BC
Hipparchus (c. 190
BC - c. 120 BC) measuring the positions of the stars
This is one guy most people forget about, but he should not be
ignored. People were trying to figure out the motions of the planets
and they were having a really hard time getting the planets' motions
figured out by only using those dang circles - they just weren't
flexible enough to reproduce all the weird planetary motions.
Astronomers had been observing the motions of the planets for
centuries and knew that they had really tricky motions to figure out,
like the way that they sometimes go backwards (retrograde) or just
move at a different rate in the sky relative to the stars. Earlier
astronomers came up with a multi-circle system known as
the epicycle and deferent that caused the objects to move in loops, however models
based upon those produced motions that were very steady and uniform, unlike the
observed motions in the sky. Hipparchus improved upon the basic epicycle and
deferent model by introducing what is called the eccentric. This is basically saying
54 | P a g e

that the Earth is not at the center of the deferent, but is a bit off of the center. Think
about how motions would look if you are in the middle of a circular race track. If the
cars are moving around at a steady pace, then all of the motions would look smooth.
Now if you were to stand near the edge of the track, the motions would not be steady
as you see them - cars would zoom past you and then appear to move very slowly
when they are the other side of the track. This sort of "speed up" - "slow down"
motion is what we see in planets and by placing the Earth a bit out of the center,
Hipparchus was able to reproduce some of this irregular motion. Remember, the
actual speeds of the planets can not change, but our view of their motion can be
influenced by our location. This is shown in Figure 5.
Figure 5. The orbit of the planet is considered eccentric,
since the center of the orbit and the Earth's location are not
the same place. Since the object is not moving at a uniform
pace as seen from the Earth, we would see it changing speed
as it went along.
The epicycle and deferent are a way to explain the retrograde
motions of the outer planets. The deferent is the large circle that is also eccentric
(Earth not in the middle), and upon this large circle, the epicycle is locatd. The planet
is found on the epicycle. The planet generally moves in one direction (eastward) most
of the time mainly due to the motion of the deferent, except when the epicycle is
turned so the planet moves in the opposite direction (westward or retrograde). The
entire arrangement is shown in Figure 6.
Figure 6. Hipparchus's early model, with the three
main aspects all shown - eccentric, deferent and
epicycle. Most of the time the planet moves in an
eastward direction, but on occasion it moves in the
opposite direction (westward).
All of Hipparchuss fixes worked pretty good at
explaining the general motions of the planets, though
they were still not as precise at predicting the motions
of the planets - more work needed to be done.
Claudius Ptolemy - 2nd Century AD
Claudius Ptolemy (c. 87 - c. 165) in a woodcut made long after
his death.

The next person in the story was probably better able to test his
55 | P a g e

model than others who came before him. Since Ptolemy lived in Alexandria, Egypt
and he was a citizen of the Roman Empire, he was able to do much more than those
before him. One thing that the Romans did when they conquered other people was to
take all of the knowledge (scrolls, records, philosophy, art, etc.) back with them. A
large amount of this information was placed into the great library of Alexandria.
Ptolemy had access to hundreds of years of astronomical data, previous astronomers'
theories, and their philosophies, with which he could work. Ptolemy, unlike some
earlier astronomers, was interested in making a very accurate mathematical, working
model of the sky. He could work on his model by testing its results against the years
of observational data that were available in the library.
One of the tricky points that Ptolemy tackled was the retrograde motion of the outer
planets. He was able to combine several previously proposed devices with some of his
own into a system that actually worked. Ptolemy's model was thought to be the best
model out there because it produced numbers that were much more accurate than
anyone else's. His model used devices such as the the deferent and epicycle of
Hipparchus and his own idea, the equant. An equant is sort of like an eccentric, but a
bit more off center (this is shown in Figure 7). The planet doesn't move at a uniform
rate as measured from the center, but from a point off center. If you were standing at
the equant, the motion of the planet would be smooth and uniform - no speed ups or
slow downs. If you viewed the motion from any other location, the motion wasn't
uniform. The Earth was also off center, so all of this off-centerness really helped
Ptolemy in getting the rate of motion correct. Of course, retrograde motion was pretty
easy to get into his model by using the epicycles and deferents that Hipparchus
introduced.
Figure 7. Ptolemy's model of the sky including the
equant. If you watched the planet's motion from the
location of the equant, it would move at a uniform
(steady) rate. Viewing from other locations, like the
off centered Earth, would result in a non uniform
motion.
Of course, the worst thing about Ptolemys model was
that it worked very good at predicting the motions of
planets - and he could test his model against all those
years of observations in the library, so he could fine tune the parameters in the model.
People accepted his model because it worked so well and used all of the basic
assumptions (though he sort of fiddled with some stuff - the Earth wasn't exactly in
the middle). The details of this model and other observations of Ptolemy are included
in his classic work, Almagest. This was pretty much the main textbook of astronomy
for many centuries.
56 | P a g e

Arabic astronomers using a variety of instruments,
including astrolabes at the observatory of Taqf ad-Din at
Istanbul in 1577. Painting is from the book Shahinshah-
nama (History of the King of Kings), an epic poem by
'Ala ad-Din Mansur-Shirazi, written in honor of Sultan
Murad III (reigned 1574-95).

It should be noted that the work of Ptolemy, Aristotle
and all others from ancient times would have been lost if
not for Arabian and Islamic astronomers. Actually, the
name of Ptolemy's book is based upon its Arabic name
(which means "the greatest"). While Europe was
wallowing in the Dark Ages, astronomers in what is now
the Middle East preserved, translated and adapted many of the works of the ancient
Greeks and Romans. When trade with the Arab world opened up to Europe, these
manuscripts were again available to Europeans. A rather significant influence on
modern astronomy due to Arab astronomers are the star names - virtually all star
names were originally Arabic. Arabic astrnomers and scientists also passed on and
improved several things, particularly instruments used in measuring star positions (the
astrolabe, an invention from ancient times), and the mathematical methods to do
calculations. The number system we use today was passed on to us from Arab
scientists (who picked it up from Hindi scientists), and of course every high school
student's favorite subject, algebra, was greatly influenced by the Arabs. One of the
greatest astronomers/scientists you never heard of is al-Hassan Ibn al-Haytham (965 -
c. 1040). This guy was doing some rather amazing things with optics and the ideas of
planetary motions well before others who are generally given all of the credit for these
discoveries.
Not only were there astronomers in the Arab world, but elsewhere around the world.
Astronomers are seen in pretty much every ancient civilization around the world,
including the Incas, Aztecs, Native Americans, Pacific Islanders, China, Japan, Korea
and so forth. You name it, there were astronomers there. Unfortunately, much of the
ancient work of many astronomers has been lost over time - there weren't always
convenient places like the library of Alexandria in every country to store this
information. Of course, the library at Alexandria wasn't able to survive to the modern
era, but was pretty much destroyed in later wars.
As the Europeans started rediscovering the work of Ptolemy and other ancient
astronomers through more contact with the Arabic world, they noted that there were
problems appearing in Ptolemy's methods. These were in the form of errors in
predicted positions which slowly increased in size over the centuries until they were
57 | P a g e

too large to ignore. People didn't want to abandon Ptolemy's model, since it had
worked in the past. To make it keep working, they just altered it slightly - usually by
adding more epicycles. After a while there were so many epicycles attached it became
rather complicated to determine the motion of planets. Perhaps something was wrong
with Ptolemy's model?
Nicolaus Copernicus - 16th Century
Polish astronomer Nicolaus Copernicus (1473-1543)
Copernicus is the next one on the scene. He was aware
of Ptolemy's model, but thought that the increased
number of epicycles and the things like the equant
were not realistic. He basically scraped the system and
produced a heliocentric (Sun centered) model for the
solar system. This was a pretty radical concept at the
time, but he wasn't completely radical in his model -
he also had perfect circles in it and motions that were
uniform. Copernicus had to include epicycles to get
accurate predictions for planetary motions because of
this.
Here's what Copernicus had in his model and how it worked -
1. Put the Sun in the middle.
2. Have everything go around the Sun including not only the Earth but also all the
other planets in the solar system and the stars.
3. By having the other planets go around the Sun, you can easily explain Mercury's
and Venus's motion - they are always near the Sun in their smaller orbits, so they
never get far from it in the sky - they are visible only in the neighborhood of the
Sun. Here is a movie showing the motion of Mercury and Venus as seen from the
Earth and here is an overhead view. In each case, whenever you wanted to see
Mercury or Venus, they were always seen in the general direction of the Sun.
4. The motions that we see over the course of the day are due to the rotation of the
Earth on its axis.
5. By having the Earth go around the Sun, we would see the Sun in front of different
constellations over the course of the year. This means the Sun doesn't have to move!
6. The retrograde motion is easy to explain - this is just what happens when a faster
planet passes up a slower planet. Think of it like this - if you are on the highway going
65 mph, and there are a bunch of cars going 60 mph, you would be going past them.
Pretend that you are not moving (you're still going 65 mph, but just ignore the effects
of your motion). If you look at the other cars now, they're going backwards from your
58 | P a g e

viewpoint. Planets are doing a similar thing so that the Earth (which is moving faster
than Mars) will at times zip ahead of Mars, making it appear as if Mars is going
backwards. (See Figure 8.)
Figure 8. The retrograde motion of Mars can
be explained if we have the Earth travel around
the Sun at a greater speed than Mars. During
the time when the Earth is overtaking Mars, it
appears as if Mars is moving backwards
(westward) relative to the background stars.

7. By having the Earth go around the Sun, you
can have the Moon go around the Earth and the
phases come about because of this motion. This
also easily explains why the sidereal and
synodic periods for the Moon are different.
To be perfectly honest, Copernicus wasn't the
first person to come up with the idea of having
a heliocentric system. A fellow from ancient
Greece previously had the idea, but most of his
model's characteristics were lost so we don't know exactly what was in his model.
You also have to consider the rather curious aspect that Copernicus's model wasn't
any better in certain respects than Ptolemy's model. Copernicus did not use an equant
but he was still sticking with circles and uniform motions in those circles. This forced
Copernicus to have quite a few epicycles to account for all of the motions. This made
his model about as complicated as Ptolemy's - different, perhaps even more logical,
but not any simpler. The quality of the model wasn't really improved either - you
would get about the same accuracy for planetary motions with either Ptolemy's model
or Copernicus's so that wasn't much of an improvement.
Copernicus also had to be careful with his model - at least in how he advertised it or
who he spoke to about it. He knew that talking about a heliocentric model was not
only bad philosophically (since it went against Aristotle) and religiously (since the
church preferred a geocentric model), but promoting a heliocentric model could also
be fatal. You have to remember the time that Copernicus lived in was pretty volatile.
During his life
1517 - Martin Luther posted his theses on the church in Wittenberg, critical of the
Catholic church.
1535 - Thomas More was beheaded for supporting the Pope over Henry VIII of
England.
59 | P a g e

1536 - John Calvin published "Institutes of the Christian Religion" in response to
persecutions of Protestants in France.
and so on and so on. By the middle of the 16th century there was enough religious
unrest caused by the Reformers, the Counter-Reformers, the Calvinists, the
Protestants, the Lutherans, the Anglicans, the Papists and everyone else that saying
anything contrary to whomever was in charge was pretty much a guarantee of
imprisonment, persecution, banishment or even a death sentence.
Copernicus didn't really want to promote his theory in part because he worked for the
Catholic church and was aware of their position on the Geocentric solar system (they
liked it and were against a heliocentric system). He finally did have his theory
published (in the book title De Revolutionibus Orbium Coelestium), but he was nearly
dead by the time it was printed up. It is interesting to note that in the introduction of
the book, which was probably not written by Copernicus, the heliocentric model is
described as being only a useful tool in calculating the motion of the planets, and was
in no way the real universe - someone was being very careful not to get Copernicus
into really deep trouble. Copernicus's book was printed in 1543, the same year he died
and by 1616, the book was placed on the Prohibited Books list by the Catholic
Church.
Sort of to press home the point as to how serious some people were about these
things, just see what happened to the philosopher Giordano Bruno (1548-1600). He
supported Copernicus's model and also went so far as to say that the stars were infinite
and there were infinite worlds beyond our own. Though it isn't entirely clear why he
was arrested and imprisoned by the Inquisition, it is a fact that he was burned at the
stake in 1600. He must have really made someone very, very mad!
Tycho Brahe - Late 16th Century
Mural showing Tycho Brahe's (1546-1601) workshops
and instruments that he used to accurately measure
positions in the sky.

Up to this point, most of the people mentioned had
worked on figuring out how the Universe worked, where
the Sun and the Earth were and how they moved or
didn't move. The next fellow on this strange journey did
also study that, but that is not what he is known
for. Tycho Brahe didn't really add much directly to the
story of determining the motions of the planets, but his
observational contributions can't be ignored. He was a
Danish nobleman who got very lucky. In 1572 a bright
60 | P a g e

star appeared out of nowhere. Now this kind of thing isn't supposed to happen - the
stars are supposed to be eternal and never changing. Only things near the Earth
change, so the new star should be close to the Earth. Tycho made very precise
observations of this star (which was called a nova, because it was a new star - this is
an incorrect term, since it should have been called a supernova, but we'll get to that
later) as well as gathered up observations of other astronomers. From the data he was
able to determine that the "nova" did not have a stellar parallax - it never changed its
position in the sky. If you remember the discussion of stellar parallax earlier, this
would have been impossible if the object was nearby. Nearby objects show a parallax
shift - the "nova" did not. Obviously the "nova" had to be really far away.
Tycho's observations and discoveries about the "nova" got him on the very good side
of the King of Denmark; so good, in fact, that the King gave him an island to build an
observatory, a workshop and labs on. The observatory and its instruments were built
for precise work and that is what Tycho is known for - his accurate observations of the
night sky. He had made excellent observations of everything including not only the
stars but also a comet and the planets' positions. His observations of the comet gave
him the same result as he got for the "nova," that the comet was so far away it did not
show a parallax. It looks like this is an example of how the sky was not as the ancient
Greeks thought - things did change. His observations of the comet were important
since people at that time thought comets were close to the Earth - they usually
assumed them to be objects in the atmosphere. This idea was due to the fact that
comets moved relatively quickly compared to other things in the sky and they were
something that changed in appearance. Since Tycho's observations did not show a
parallax, he proved that the comet must be really far away, not an object in the
atmosphere but well beyond the distance to the Moon.
Tycho had years of very precise observations of the motions of the planets. How was
he able to do this? He didn't use a telescope; all observations were made with his
unaided eye. He just did the same observations over and over and over and over... By
repeating the observation, you reduce the amount of inaccuracy. This is what Tycho
did.
Eventually Tycho's luck ran out. The King that favored him died, and the next King
didn't really like Tycho, so he got out of Denmark and went to work for the Holy
Roman Emperor in Prague (of course the Holy Roman Emperor was neither holy, nor
a Roman nor an emperor, but I digress). While there he hired a number cruncher to do
some calculations - Tycho didn't really like to do the math; he preferred to do the
observations - sort of like how some astronomers are today; some stay at the telescope
all the time, and some sit in front of a computer all the time. Tycho was trying to
come up with a model of the solar system that was actually a bit of a hybrid between
Copernicus's and Ptolemy's, but he didn't want to do the math. Tycho's career in
61 | P a g e

Prague was sort of cut short when Tycho partied just a little too much one night and
died (1601).
Johannes Kepler - Early 17th Century
Johannes Kepler (1571-1630)
Kepler was the poor fellow who Tycho hired to do
all of the messy calculations. But when Tycho died
Kepler had access to all of his data. This is really
what Kepler wanted all along. Kepler was fairly
obsessed with figuring out the motions of the
planets. And he knew that to do that, and to do it
right, he needed the best data. Tychos death gave
him that data.
Kepler liked the model that Copernicus came up
with, but he couldn't get it to work using only those
dang circles. Over time he finally did hit upon the
correct solution and gradually came up with
the Three Laws of Planetary Motion. It should be noted that these laws do not
explain WHY the planets move, but they just show HOW the planets move. He wasn't
able to figure out what the force was that drove the planets in their paths, but at least
he had a way of figuring out how to accurately determine their locations - much more
accurately than Ptolemy's or Copernicus's models ever could. It took many years for
Kepler to derive each law, and of course he had different names for them (he didn't
call them Law #1, #2 and #3 like we do now). Without much further ado - here are the
three laws of planetary motion...
1. Planets move in ellipses with the Sun at one focus. (See Figure 9) There are several
aspects that can be explained by using an ellipse and not a circle to explain the
planetary motions. By using an ellipse, the planet can be closer or further from the
Sun since the Sun is not in the middle, and the ellipse stretches out the orbit. Where is
the Sun? The Sun is close to the center but it is not in the center. It is actually found in
the location known as the focus (plural: foci). This is a position in an ellipse that has
mathematical meaning which we'll get to. For the most part, the planets in the solar
system have orbits that look to be very circular, but are actually elliptical. The
elliptical aspect is not easy to see for the most part but it is there. By using an ellipse,
Kepler could get rid of the things that Ptolemy and Copernicus used (stuff like
deferents and epicycles) to make the numbers come out right.
62 | P a g e

Figure 9. Basic aspects of an ellipse showing
the Sun's location (at a focus), and the two
points where the Earth is closest and furthest
from the Sun.
There are times when a planet is closest and
furthest from the Sun, because the planets
orbit in ellipses. Aphelion is when it is furthest from the Sun and perihelion is when
it is closest to the Sun. The distance between the Earth and the Sun is always
changing. If that is the case, how can you answer a question like "How far is the Sun
from the Earth"? We do that by using a very convenient average of the distance
between the Earth and Sun. The average is 150 million km. That's not convenient!
How's this - the average distance between the Earth and the Sun is 1. Yes, it is
1 Astronomical Unit (or abbreviated as 1 A. U.). This is sort of like avoiding the use
of small units of distance to measure large distances. You could say that you were 66
inches tall. Is that tall? If you remember that there are 12 inches in a foot, and use feet
instead of inches as your unit of measure, 66 inches is the same as five and a half feet.
You could go around saying that the average distance between the Earth and the Sun
is 150 million km, or you could say it is 1 A.U. Which is easier? Well, duh, the A. U.
is easier to use. The A.U. is also a handy distance for other objects in the solar system;
you could say that Mars is 1.5 A.U. from the Sun, or that Jupiter is a bit more than 5
A.U. from the Sun. By giving the distances in A.U., you have their distances scaled to
the Earth's distance (Mars is 1.5x further from the Sun than the Earth, while Jupiter is
a bit more than 5x further from the Sun). This a lot easier to comprehend then to say
Mars is 223 million km and Jupiter is 778 million km from the Sun, right?
Now to answer the question, what actually are the foci? They are a way that ellipses
are defined - the distance from one foci to the edge of the ellipse and over to the other
foci is always the same - see Figure 10 for this. Now here is a rather nifty thing - a
circle is actually a type of an ellipse. In this case,
both foci are located in the center.
Figure 10. The foci-edge-foci distance remains
constant - so that all of the different colored line
segments have the same total length.
To distinguish between the different types of
ellipses, a value known as the eccentricity is
used. The value of the eccentricity is defined as
the ratio between the distance between the foci divided by the longest distance across
the ellipse. These are shown in Figure 11. As the ellipse gets more stretched out, the
eccentricity gets larger.
63 | P a g e

Figure 11. The eccentricity would be the value of
the foci-foci distance (green line) divided by the
widest part of the ellipse (blue line). For the
circle, the green line has a value of 0, so the
eccentricity for a circle is 0. For the other two
ellipses, the eccentricity increases as the foci get
further apart.
Time for the second law -
2. A planet moves faster at perihelion than at
aphelion. Sometimes this law is stated as "equal
areas swept out in equal times," which is not the
most straightforward definition in the world. See
Figure 12. This law basically says that when a
planet is closer to the Sun (at perihelion), it is
moving faster in its orbit about the Sun than when it is far from the Sun (at aphelion).
The Earth moves around the Sun fastest when it is closest to it (this occurs in early
January). Kepler originally derived this law using wedges and triangles to measure the
areas so the old phrase with "equal area" is often quoted, though it is a bit confusing.
If the ellipse is very close to being a circle, there is little variation in the speed. As
long as it is an ellipse, the speed difference is there. If the orbit is very elliptical
(stretched out), then the difference in aphelion and perihelion speeds are great.
Figure 12. Kepler's Second Law - planets
travel faster in their orbits when they are
near the Sun, and slower when they are
away from the Sun. If the motion of a planet
for one month is shown (the planet moves
the distance designated by the green arrow)
then the planet nearer the Sun is obviously
moving faster. If you could measure the area created by a planet as it moves in its
orbit during one month, the area would always be the same size (the two black wedges
have the same size so long as they were created during the same time span).
Last but not least, the third law, and this one is a formula -
3. If you define the period of the orbit as P and the average distance from the Sun
as a then you get the following relation -
P
2
= k a
3

64 | P a g e

What's k? That is a constant (a non-changing number) for the situation you are
looking at. The basic upshot of this formula is, the more distant planets (big a values)
take longer to go around the Sun (big P values). Average distance is used in the
formula (remember a is the average distance), since the distance between the planet
and the Sun is always changing (Law #2). As P increases, so does a, and
as P decreases, so does a. You could say it the other way (as a increases, so
does P ...).
One of the neat things about Kepler's laws is that they can be used for anything
orbiting anything else - not just planets going around the Sun, but also for moons or
satellites going around planets, stars going around the galaxy, or entire galaxies going
around one another. We'll see this later. The trick, though, is to use the correct value
of k. The value of k will depend upon what the orbiter and orbitee are - the value
for k will vary from one system to the next, so you will need to know it before you can
use the formula. You could find it if you know what P and a are.
There is also a special situation that makes using the law really convenient. If P is
measured in years, a is measured in A.U.s and the object is orbiting the Sun then the
following statement applies (in this case k=1) -
P
2
= a
3

For the Earth we have P=1, a=1, so we get 1=1. I suppose you don't believe me. Let's
try another planet. For Mars P=1.88, which when you square it P
2
=3.52. What
would a be in this case? a
3
has to equal 3.52, so by taking the cube root of 3.52 you
get a, which in this case ends up being 1.52 - and this is the correct value for a (you
can verify all of this by looking up the value for P and a somewhere for any of the
planets in our solar system).
Here's another example: What would the period of a planet be if its average distance
from the Sun is 4 A. U.? You have a value for a, and you need to get P. You can use
the special short version of the formula, since you are orbiting the Sun. Now you
need to calculate a
3
= 4x4x4 = 64. This means that P
2
= 64. What times itself is
64? P must be 8 (since when you square it you get 64). It would take eight years for
an object to orbit the Sun if it has an average distance of 4 A. U.
What if the orbit is very elliptical, how can you determine the value of a? Actually it
is quite easy. By definition, a is 1/2 the longest width in an ellipse - in Figure
11, a would be 1/2 the length of the blue line. Take a peak at the ellipses in Figure 13.
An object orbiting along those paths would take the same amount of time to complete
one orbit, since they all have the same length for their longest axes.
65 | P a g e

Figure 13. Different eccentricity ellipses all with the
same length for their long axis (so also the same value for
"a" for each of them). Planets traveling in these orbits
will all take the same amount of time to complete one
orbit.
I should mention the fact that laws 2 and 3 deal with
distances, but not in the same way. Law 2 deals with how
the changing distance of a planet in its orbit affects its
speed in orbit, while Law 3 deals with an average distance
and how that relates to the time for one orbit. Don't get
these two mixed up. For example, looking back at Figure 13 with the different
eccentricity ellipses - the most circular one would have the most uniform speed - little
change in velocity at any point in a planet's orbit, while the most eccentric one would
have a wide range in velocity, from very fast at perihelion to very slow at aphelion. In
spite of these speed differences, since all the ellipses have the same width, an object
would take the same amount of time to travel around the Sun regardless of which of
these paths it was on.
With his 3 laws (which took him years to figure out), Kepler was able to calculate the
orbits of the planets very precisely, much more precisely than could be calculated
using Ptolemy's or Copernicus's models. Again, it should be noted that Kepler really
didn't know why his laws worked; he did not know what the force was that was
behind the motions. He actually was sort of favoring some kind of cosmic magnetism,
but that's not right.
Again, it should be emphasized that while Kepler came up with his laws so that he
could explain the complex motions of the planets, his laws apply to anything orbiting
anything else. So the Moon orbiting the Earth
obeys Kepler's Laws, and a distant star orbiting
another star obeys Kepler's Laws - it's not just for
planets!
Galileo Galilei - 17th Century
Galileo (1564-1642), who was born in Pisa Italy.
While Kepler's ideas were pretty much banned by
the Church (that was both a religious thing and a
political thing) someone found his ideas very
interesting, as well as Copernicus's theories. That
fellow was Galileo, a rather outspoken gentleman
66 | P a g e

who did a lot of work in physics trying to figure out how things moved. His main
contributions to astronomy were not his theories of how the planets moved but his use
of the telescope. He did not invent the telescope (someone in Holland did that), but he
learned about the invention, improved upon it and made one that was much better than
the original design. The telescope was originally used by sailors to view great
distances over the water to spot other ships and land. Galileo used his telescope to
look at the sky. In just a few short weeks his observations of the sky literally blew his
mind and gave him evidence that convinced him that Copernicus (and Kepler) were
right. What did he see? (If you click on the items below, you'll see the drawings that
Galileo made of his observations.)
The Moon was a place like Earth with mountains, valleys, and other features that also
exist on the Earth. It was not a perfect heavenly object. Remember, things in the sky
were supposed to be perfect and eternal - and with all the bumps, holes and rough
surface features visible on the Moon, Galileo could see that it was far from perfect.

The Moon does not generate its own light, but is bright due to light that it reflects
from the Sun. This is obvious from the way that the shadows are cast on the Moon by
the ridges and mountains there.
Venus goes through phases like the Moon. This is easy to explain if it orbits the Sun.
If it orbited the Earth, as in Ptolemy's model, then it would not have the range of
phases that Galileo saw. Galileo could also see that Venus's size changed along with
its phases, indicating that it was getting closer or further away from the Earth - how
could it do that if the Earth is in the middle and Venus is on a perfectly circular orbit
around it? Frankly, it can't, because the Earth is not in the middle and Venus is not in
orbit about the Earth. (See Figure 14).
Figure 14. Two ways that Venus and the Sun
could be viewed from the Earth. In the
geocentric view (bottom), the Sun and Venus
must orbit the Earth. In that case, Venus
would always be located so that most of its lit
surface would be visible from the Earth. This
is not what Galileo saw. His observations
confirmed the Copernican view - that Venus
orbits the Sun (as does the Earth). In this case
we would see Venus go through a variety of
phases and it would increase and decrease in
size as it orbited the Sun.
67 | P a g e

There are four objects orbiting Jupiter.This broke the old rule "everything orbits the
Earth." How's this possible? Galileo tried to make some money by naming the four
objects the Medici Stars, but we all know them today as the Galilean satellites (the
Medicis were the folks that ruled most of central Italy - sort of like the most powerful
multinational corporation of their day, but quite a bit more ruthless). If these things
don't orbit the Earth, then how can they exist in Ptolemy's or Aristotle's models? They
can't. They are okay in Kepler's model. In fact, if you watch the little "stars" move
around Jupiter, you can determine that they obey Kepler's laws of motion (those
further away from Jupiter have longer orbital periods).
There are things on either side of Saturn that sort of look like "ears". This breaks the
rule about everything being a perfect circle in the sky. Galileo's telescope was good,
but not good enough for him to see the rings of Saturn clearly.
Then Galileo did a big time stupid thing, he pointed his telescope at the Sun. While he
was damaging his eyes he noted that there were "blemishes" on the Sun (what we call
sunspots). This is another instance of imperfections in the heavens. Of course, doing
something as stupid as pointing his telescope at the Sun contributed to the blindness
that he had later in life.
Galileo's ego got him into a lot of trouble with the church. He published quite a few
books about his theories and observations but the one that really got him in trouble
was his book Dialogue concerning the Two Chief World Systems, published in
Florence in February 1632. It probably wasn't so much about what he said in
Dialogue, but how he said it. The book was basically a debate between three people,
one supporting Ptolemy and Aristotle, one supporting the Copernican view and one
who was on the fence. The person supporting Aristotle's view is named "Simplicio,"
and he's portrayed as a fool. Some people thought that the character of Simplicio was
an amalgamation of a bunch of different people, mainly scientists and church officials
who held on firmly to the Aristotelian view. I guess some people just don't respond
too well to criticism. Unfortunately for Galileo, these were people you do not want to
offend! He was put on trial and forced to recant all that he taught about the
heliocentric model. Galileo had to spend the rest of his life under house arrest. He
eventually went blind and died in 1642.
Isaac Newton -
Sir Isaac Newton, (1642-1727), though because of
revisions of the calendar system, Newton birth may also be
listed as occurring in 1643.

Newton was born in the same year that Galileo died. He
68 | P a g e

was pretty much a genius in his day, and he was able to do what Kepler could not -
figure out why the planets moved the way they did and apply that motion to
everything in the Universe. Apart from that he also had time to invent calculus
(though some think that another did that - he at least shares some of the credit, or
blame, for it), figure out all sorts of rules of optics and physics and basically set up the
basis for modern physics. Obviously the guy never married - he was the major nerd in
his time.
In order to understand how the Universe work, Newton had to set up some ground
rules that pretty much are the basis for all kinds of motion in the Universe - not just
motions of planets, but motions of everything. These are his Three Laws of Motion.
Law #1. Stated very simply "An object at rest stays at rest, and a body in motion stays
in motion. This situation can be changed if the object is messed around with by
something." Okay, so the original statement of this law did not say "messed around
with," but I think you understand what I mean. Stuff will keep doing what it is
currently doing if no one or nothing messes around with it. A pencil will stay laying
on a table for all eternity so long as nothing touches it (and the table lasts forever). If
an object rolls across the floor it will keep rolling forever. Wait a minute - that's not
correct. If you roll a ball down the hall it will eventually stop. It stops because
something messes with it, in this case, the friction of the floor. If there was no friction,
no resistance to the motion, then the ball would go down the hallway forever.
You may have heard this law called the Law of Inertia. Another thing about this law
is what the motion is like. The motion should be straight, assuming that the surface is
straight. An object should always move in a straight line at a steady speed unless
something alters its motion. You may have experienced this effect in a roller coaster
or a car, depending upon how fast you drive. If you are in a roller coaster and it is
going straight, everything is fine. Then the track curves to the side, and fortunately the
roller coaster car stays on the track and goes around the curve - what happens to you?
Your body wants to keep going forward - things want to do what they are currently
doing - but the car has changed direction. This results in you getting thrown to the
side of the car. Actually, another way of looking at this is that the car is the one that
ran into you; after all, you were going along just fine when it decided to change
direction. A change in direction is another way of looking at motion changing.
Law #2. This one is pretty much a definition. If m=mass (material, stuff)
and a=acceleration (a change in motion) then
F=ma
(Force = mass x acceleration)
69 | P a g e

More mass requires a greater force to accelerate it (get it moving, slow it down, or
change the direction of the motion). It is harder to get a locomotive moving than it is
to get a Volkswagen moving. You could also look at how the different masses act
under the influence of the same force. Let's say you have a baseball bat and you hit
two objects, a baseball and a bowling ball. What sort of motion results? If you said
that the baseball will move further than the bowling ball, then you were right. You
could solve this numerically. Let's say that the bowling ball has 100 times more mass
(m) than the baseball (I really have no idea; I don't bowl). How do their accelerations
compare? The baseball will be accelerated 100 times more than the bowling ball. If
the force applied to both is the same (same F), then the values relating
the m and a should be equal in size but opposite in their sense (m big, a small
and m small, a big).
In case you were wondering, mass is NOT weight. Mass is the measure of the amount
of stuff in objects. Weight is just how much pull you feel due to the gravitational
acceleration of the Earth. Don't you feel like you weigh more when an elevator is
starting to go up? It is accelerating, and unless you are eating a donut, your mass isn't
changing, yet you feel heavier - you are feeling a new force along with the regular
force of gravity. The Earth is constantly accelerating (pulling) you downward. We'll
get to that. That's why you always fall down rather than up.
Law #3. "For every action there is an equal and opposite reaction." You've probably
heard this one before. You might want to think of it as the "you can't get away with
anything" law. If you do something, you will get something done back at you. If you
hit your hand against the table, the table hits back at you - that's why it hurts, so don't
do it! If you keep hitting the table, you might want to note how you are hitting it -
you're exerting a downward force, while the table is exerting an upward force on you -
that's the "opposite" part of the rule. While you might not think that the law is "equal"
since your hand hurts and the table doesn't appear to be damaged - don't worry, it is
getting hurt. Just keep hitting it for about five hours - you'll eventually see some effect
of the force that you are exerting, though of course your hand will be pretty much
ruined by this time.

While these rules might not seem all that Earth shattering, they are. They are just the
basis for all the motions of the Universe. To actually figure out what was driving these
motions, Newton had to use these laws and the concepts behind Kepler's laws to
formulate a rule that could explain the motions of all objects in the Universe.
Newton's Universal Law of Gravitation
Here's how this one goes - Objects will attract one another by an amount that depends
only on their mass and their separation. There is a formula that goes with it -
70 | P a g e


F=Force of Gravity, you may want to think about this as the amount of "pull"
something has
G=Constant - don't worry about what it is, you usually won't have to know it to use
this formula.
M
1
, M
2
= masses of the objects, you need two to tango and two masses are also
needed for gravity.
R = the distance between the center of masses of the two objects. We'll simplify this,
since most important things in astronomy are fairly spherical, so R can be thought of
as the distance between the two objects centers.
Now you have to be careful with this formula, since there are several things that can
change the value you get for gravity. The basic upshot is that if the mass goes up the
force of gravity goes up. If the mass goes down, the force of gravity goes down. If the
separation changes, then the force of gravity changes in the opposite manner.

The law is "Universal," meaning that it operates throughout the Universe, on all
matter at all distances. If you have mass than you are gravitationally attracting (and
being attracted by) everything else in the Universe with mass. This would include the
chair you are sitting in, the person next to you, etc. Of course, you don't feel the
gravitational pull of the person next to you (unless, of course, they are very attractive,
but that is not a real force of physics, more like a force of biology), because the
strongest pull you are feeling is due to the nearest big object, the Earth. Everything in
the Universe with mass is pulling on you and you're pulling on it - that's a lot of stuff.

Let's look at the Earth's gravity for a second. The force of gravity that you feel
standing on the surface of the Earth depends on the masses of the objects involved
(your mass and that of the Earth) and the distance between your centers of mass,
which is basically the radius of the Earth, since you are pretty small compared to it.

If you were to stand on a mountain, you would be further from the center of the Earth
and the distance in the formula would be larger. If that was the case, then you would
feel a lower gravitational acceleration - you would actually weigh less!
Just remember, if you change the masses or the separation then the force of gravity
changes. Let's try an example. Let's say you travel to a distant planet called
Gumbyville. This planet has a mass that is four times the mass of the Earth. How does
its gravity compare to the Earth's? The only difference from the Earth's gravity is an
71 | P a g e

extra factor of four in the top of the formula. You would feel a force of gravity that is
four times what you feel on the Earth - you would weigh four times what you weigh
now. Let's suppose that this planet also has a different radius than the Earth. The
planet's radius is three times that of the Earth. Now here's where most people mess up
- the distance value, R, has a two next to it, indicating that it should be squared. If you
change the value of R, then the change gets squared as well. For Gumbyville, this
results in a value of 3x3=9 on the bottom of the formula. What does that give us? A
four on the top and a nine on the bottom - the overall effect is that the force of gravity
is 4/9 that of the Earth, or slightly less than 1/2. You would actually weigh less even
though Gumbyville has more mass than the Earth.
Here's something you don't want to forget - Gravity is the most important force in
the Universe! Yes, it is even more important than Obi-wan Kenobi and the Force!
Gravity doesn't go away; it is always there pulling things together or squeezing things
down. It is always ready to take over a situation if you give it half a chance. As you'll
see, without gravity there would be no galaxy, Sun or Earth - it pretty much is the
main driving force behind the whole Universe.
Using the formula for gravity that he came up with and some basic calculus, Newton
was able to redo Kepler's laws and explain the motions of the planets as an effect of
the force of gravity. A couple of things can be seen in how Kepler's laws behave if
you look at them in terms of gravity - like the second law, which says that planets
move faster in their orbits when they are closer to the Sun. If they're closer to the Sun,
what's the gravity like? The gravity must be stronger since the distance is smaller, so
the planets are feeling more of a pull from the Sun (and of course the Sun feels the
pull from the planets, but it's so big it doesn't really care). How can the planet prevent
itself from being pulled into the Sun due to the increasing pull of gravity? If it goes
faster in its orbit, it will be able to balance out the force of gravity - sort of like having
to swing a bucket full of water around fast enough to prevent any of the water from
falling out. Further from the Sun there is less of a pull and the planets don't have to
move as fast. You have to remember, Newton's laws and even Kepler's laws are not
confined to just planets; they can work all over the Universe on all sorts of objects
including the Universe itself.
Configurations of Planets
All this history stuff was to explain how we figured out the motions of the planets.
There are a few more bits about the planets that need to be covered, specifically some
of their special alignments. To understand the various alignments of the planets, we
have to look at them in two groups, since you can have only certain alignments
possible for planets in certain locations.

72 | P a g e

Superior planets are those located outside of the Earth's orbit. Amongst the planets
you can see with the naked eye, these include Mars, Jupiter and Saturn. Their main
configurations are shown in Figure
15.
Figure 15. The main named
locations for the Superior Planets.
The positions are sort of similar to
the Moon's various locations. For
example, the opposition location is
like the Full Moon; a planet there
would be visible high overhead at
Midnight, it would rise around 6
PM and set around 6 AM.
The quadrature locations are sort
of like the quarter phases of the
Moon, at least in terms of the
questions of "when do they rise,"
"when do they set," etc. This is
because quadrature and the quarter
phases of the Moon both are
positions when the object is 90 degrees away from the Sun. The two quadratures are
distinguished as being eastern quadrature or western quadrature if the object is 90
degrees east or 90 degrees west of the Sun respectively. Of course, the location
designated as conjunction would be most like the New Moon, but in this case the
planet is behind the Sun, not in front of it. A planet at conjunction would be doing
things at the same times as the New Moon or the Sun does them (rises at sunrise and
sets at sunset).

Inferior Planets are those located inside of the Earth's orbit and include Venus and
Mercury. Their main alignments are shown in Figure 16.
The position of conjunction depends upon whether the planet is in front of (inferior
conjunction) or behind (superior conjunction) the Sun. As with the superior planets,
a conjunction is an alignment in the direction of the Sun. The other two positions
indicate when the planet is at its greatest angular distance from the Sun. If it is west of
the Sun as far as it can get (from our perspective), it is at maximum western
elongation, while being east of the Sun puts it atmaximum eastern elongation.
Sometimes the phrases greatest eastern elongation and greatest western
elongation are used - either way it means the same thing. These planets are always
73 | P a g e

seen near the Sun and are really only visible when the Sun is out of the way, so either
before the Sun has risen or after the Sun has set. When a planet is to the East of the
Sun, it would be visible in the Evening sky after the Sun sets (you can remember this
by noting both begin with "E" - east, evening). A planet located to the West of the Sun
would be visible in the Morning sky before the Sun rises (you can remember this by
thinking of the M as an upside down W).
Figure 16. The main named
locations for the Inferior Planets.


Now that you've read this
section, you should be able to
answer these questions....
How did the early
assumptions about the solar
system limit the models that
were devised by the ancient
greeks?
What was significant about
Copernicus' model? Was it
better than Ptolemy's?
What contributions did
Brahe make to our
understanding of the motions of the planets?
What are the three laws of planetary motion and what can they be applied to?
What was it about the objects that Galileo observed that made him support
Copernicus' model?
What are the three laws of motion?
What is the law of gravity? What does it depend upon?
What are the special planetary locations (as viewed from Earth) and when
would planets in those locations be visible, or rise, or set?
Light

What's covered here:
How do we describe light?
What types of light are there?
74 | P a g e

How are light, energy and radiation related?
How does light interact with matter?
What are the rules that govern hot objects?
What types of telescopes are there?
What are the limitations of telescopes?
What other equipment do astronomers use?

It is very important to understand the properties of light - why? Light is the main
means by which we study most astronomical phenomena. Astronomy is different than
other sciences in that astronomers can't actually handle, touch, or alter in any way the
stuff they study (this isn't always true - those who study planets and meteorites can
sometimes touch what they study, but they're in the minority). Almost all the
information we obtain is via light, so the more we know about how light works, what
it is, and how we can wring out information from it, the more we can use it to discern
features about the stars and galaxies which are beyond our reach. Astronomy is the
most voyeuristic of all sciences - we like to watch!

Exactly what is light?
If you want the $25 term for it, you can say it is an oscillating electro-magnetic wave.
That's pretty much a mouthful, but what does it really mean? Actually, the most
important part of this definition is that light is a wave; it has wave-like properties. It,
like sound or water, acts like a wave. Think of it as a wiggling, squirming thing if you
like.
Since light does stuff like a wave, it has a certain size. You know that not all waves
are alike, and therefore not all types of light are alike. You can distinguish the
different types of light by measuring the distance between consecutive peaks or
consecutive valleys. This distance is known as the wavelength. Often we abbreviate it
with the greek letter "lambda" - . Since it is just a distance, we can measure it in
units such as centimeters (cm), millimeters (mm) or other normal units of distance.
What else do we know about light? It travels. It has a certain speed, since it travels.
The speed of light is denoted by the letter c, and is equal to 300,000 km/s = 186,000
miles/s. That's pretty fast.
Probably the most important aspect of light is that its speed in space is always the
same - c is a constant. Actually, in some strange physics experiments people have
been able to slow light down to very slow speeds, but those are under unusual
circumstances and not what we see when we look up in the sky. As far as we're
concerned, the speed of light is always the same! Don't forget that! Now, the speed of
75 | P a g e

light doesn't care about the wavelength of the light; it can be long wavelength, short
wavelength, or medium sized wavelength. In all cases the speed at which light travels
is the same.
Take a look at Figure 1. There are two different types of light shown, each with its
own wavelength. They both travel at the same speed. Let's say we are going to do
another experiment with the light. We are going to watch the two types of light pass
by and we are going to count the number of waves that pass by each second. We like
doing these sorts of experiments, since we are strange scientists. Will we count the
same number of peaks passing by each second? Here is a little animation showing the
experiment - remember, both waves of light have to move at the same speed,
regardless of their wavelength. This is another way to distinguish the different types
of light. The one with more peaks squeezed together will give us a higher count, or
you could say its peaks pass by more frequently, since both types travel at the same
speed. In other words, you will count more peaks passing by each second for the
shorter wavelength light. You could also say that it has a higher frequency compared
to the longer wavelength light.
Figure 1. Two different types of
light are shown - one with a short
wavelength and another with a long
wavelength. Or conversely a high
frequency and a low frequency.
The other way to distinguish light is
by the frequency, which is denoted
by the letter f. Frequency is often
given in units of "per second" such
as "peaks that pass by each second."
Generally it is harder to get a mental
picture of light using frequency, so
I'll just stick to wavelength to
describe the different types of light.
Now there is a really neat relation between the speed, frequency and wavelength of
light. Here it is -
c =f
This is pretty simple and straight forward (unlike some formulas). Since c must
always have the same value, when you change the wavelength, the frequency must
also change, but in the opposite sense. When the wavelength increases, the frequency
76 | P a g e

decreases. Long wavelength light has a low frequency and short wavelength light has
a high frequency.
As I mentioned, we'll distinguish the different types of light based upon their
wavelength. It is important to be able to measure the wavelength to distinguish the
various types of light since there are many different types of light, each with their own
wavelength value. However, wavelength is often so small that it is difficult to describe
in units as "large" as cm or mm. Often astronomers use the units known
as ngstroms, and in case you were wondering 1 = 10
-10
m. These are pretty small
units and are good at covering the range of wavelengths very well.
The following table gives the various types of light that astronomers study. Since
frequency is a bit harder to picture, its value will not be given, though its size is
denoted in general terms. The types of light listed in this table are all studied by
astronomers, though for the most part a lot of astronomy is done in the visible part of
the spectrum - what you can see with you eyes. The table is ordered in terms of
wavelength from longest to shortest wavelengths. Don't worry about the last column;
we'll get to that later.
Light
Wavelength in
meters
Wavelength in

Frequency
Blackbody Temperature
(K)
Radio 10
-2
and greater
10
8
and
greater
low 0.3 or less
Microwaves 10
-3
to 10
-2
10
7
to 10
8


0.3 to 2.9
Infrared (IR) 7 x 10
-7
to 10
-3
7000 to 10
7


2.9 to 4100
Visible 4 x 10
-7
to 7 x 10
-7
4000 to 7000

4100 to 7250
Ultraviolet
(UV)
10
-8
to 4 x 10
-7
100 to 4000

7250 to 300,000
X-ray 10
-11
to 10
-8
0.1 - 100

300,000 to 3 x 10
8

Gamma-ray 10
-11
and less 0.1 and less high 3 x 10
8
and greater
You have heard of all of these types of light in one way or another. Radio is just what
you think it is - the type of light that your car radio picks up. The range of radio light
is pretty wide, and the stuff that goes into your radio or tv is a pretty small part of the
entire range. A lot of radio wavelengths are taken up by various communication
systems, including military, CB radio, ham radio, and so on. Radio light is pretty
much all over the place, even though you can't see it with your eyes. Another type of
light that is vital to the existence of a typical college student is microwave light. Some
people don't distinguish microwaves from radio, but we'll consider them as different
77 | P a g e

types of light. Yes, they do cook your dinner in the microwave oven. However, most
microwave light is harmless; a lot of cell phones use it for communication. The next
type of light is actually the type given off by pretty much everything around you -
infrared or IR. This is the type of light that is also associated with heat. You and all
other objects around you are probably giving off IR light - things at room temperature
do just that. You may have seen military or police videos that were shot using IR
cameras - these are just picking up the light given off by humans and other warm
things.
The type of light that you experience most of the time is visible light. You might just
think of it as white light from a typical light bulb or the yellow light from the Sun, but
if you were to take that light and pass it through a prism you'd get a rainbow. Visible
light is made up of actually all sorts of different types of light (different wavelengths).
It is often surprising to people to learn that white light is actually made up of all these
colors combined together. If you have ever seen a rainbow then you have seen the full
range of visible light. It is easy to remember the order of visible light from long to
short wavelength since it is the same as the rainbow -
Red,Orange,Yellow, Green, Blue, Indigo, Violet. Of course, if you don't have a
rainbow handy, all you need to remember is Roy G. Biv, the world famous
astronomer! Most pictures of objects that you'll see in this course were taken with
visible light cameras, since we can easily produce those images and our minds are
geared to interpret them. Red has the longest wavelength of visible light (it comes
right after infrared, after all), while violet is the shortest wavelength visible light.
Continuing on down the light spectrum to shorter wavelengths, after visible light, or
in this case after violet light, comes ultraviolet. Don't add an "n" to the name and
make it ultraviolent, even though this type of light is rather nasty stuff. This is the type
of light that gives you a sunburn in the summer and can lead to unpleasant things like
skin cancer. Right on the heels of ultraviolet light comes an even nastier form, x-rays.
You know this stuff is nasty when the dentist leaves the room when they x-ray your
teeth. Finally, we have the shortest wavelength type of light that is studied by
astronomers, gamma-rays. These are very short wavelength and are also rather rare.
There is the whole range of light, from longest to shortest wavelength.

Most importantly, you should remember that all types of light travel at the same
speed, namely c.

I've been rambling on and on (as usual) about the nature of light, how it is a wave,
how it has a certain type of wavelength, how the wavelength can be used to
distinguish it, and so forth. When people were trying to figure out what light actually
was, it took many years to understand the nature of it, in part because it is sort of
78 | P a g e

schizophrenic - it has two different personalities. Sure, there are times when light acts
like a wave, but there are also times when it acts like a particle, sort of like a little spit
ball or b-b pellet. When physicists were doing their little experiments, there were
certain tests where the result was "Hey, this thing is a wave" and other tests where the
result was "Wow, this light stuff is a particle!" That's confusing. Yes, light acts like a
wave, but it also acts like a particle. Let's check out that side of its personality.

What type of particle are we talking about? Light can be thought of as a particle of
energy, which is known as a photon. The idea of loading photon torpedoes during a
space battle actually makes sense, because you'd be loading up a bunch of energy.
Light can be a chunk of energy, but exactly how much energy is in one of these
photons? This is where that wave-like nature of light comes in handy. This is because
the energy of the photon depends on the type of light you are dealing with, and the
best way to distinguish light is by its wavelength (or frequency). The energy is
directly dependent upon the wavelength (or the frequency), in the following relations -
E = hf
E = hc/
Don't freak out; the formulas aren't that bad. Both formulas define the same thing but
one uses frequency while the other uses wavelength. You first have
the Energy (the E in the formula) of the light (or you could say the amount of energy
in the photon) and it is related to the frequency - in fact it is directly related to the
frequency. If you double the frequency, you double the energy. The other symbol in
the formula, h, is a constant, and you can ignore it for the most part. You can also
make the energy related to the wavelength as is shown in the second formula, since
frequency is related to wavelength. Wavelength and frequency act in an opposite
manner, so that when one goes up the other goes down, that's why the wavelength
ends up on the bottom in the formula. All you need to remember are the following
trends - Energy UP, Frequency UP, Wavelength DOWN. The reverse is also true. If
you double the wavelength, the energy is cut in half as well as the frequency. If you
triple the frequency, the energy is tripled but the wavelength is three times smaller.
This whole thing relating the wavelength to the energy of the light kind of makes
sense. Take a look at the table showing the different types of light. All of the nasty
types of light are at the bottom of the table. This is the short wavelength end, as well
as the high frequency end. What does that mean for the energy of the light - you got it,
it's the high energy type of light, which kind of explains why some of it can kill you;
it's nasty stuff!
Astronomical Spectra
79 | P a g e

Light can be distinguished by its wavelength, but why stop there - there is a whole lot
more we can do with light, just by using a piece of cut glass. If you were to pass the
light from an object (a star, the Sun, a glowing gas cloud) through a prism, you would
produce a rainbow like feature known as a spectrum (plural: spectra). Figure 2 shows
a prism spreading out light; actually it's just
artwork, not a real prism.
Figure 2. The album cover from Pink Floyd's
"Dark Side of the Moon" showing a prism
spreading out white light into the colors of the
spectrum.
When we look at the spectra from various objects
out in space, we don't see the same spectra - there
are some with big gaps, some with colors missing,
and some that are a real mess. How is this possible? What causes the different types of
spectra?
To understand this, we have to understand how atoms interact with light and how
spectra are produced. Perhaps this would be a good time to get a nice strong cup of
coffee, or perhaps some highly sugar loaded caffeinated beverage - don't worry, I'll
wait. First of all, you have to remember some basic physics about atoms. A typical
atom has a specific structure based upon the type of atom it is. This structure includes
many possible orbits for the electron or electrons if there is more than one. These
orbits are located at various levels, which correspond to different energy levels. You
might want to think of this as a bunch of mountain climbers. The climber with the
most energy will go to the highest part of the mountain, up to a certain ledge, while
those with lower amounts of energy will not get that high up, but will stay on the
lower edges. If they get rid of their energy, they will go back down the mountain to
the bottom level (known as the ground state in an atom). Electrons like to be in the
ground state (the lowest level). You might want to think of them as mellow particles;
maybe they just like to lay around the house all day in their underwear. Whatever the
cause, they don't like to be excited and in those higher orbits (higher energy levels). If
they do get into a higher orbit, they really want to get back down to the ground
(lowest energy) state.
However, atoms are not exactly like mountains, since the electron can only go from
one level to another if it absorbs or emits the exact amount of energy (light). It can't
go to a place in between - there are only certain energy levels available to the electron
- sort of like there are only certain places on the mountain where you can stop and
stand safely. Electrons are very picky about the type of light (amount of energy) they
will use. Only certain wavelengths of light will interact with the atom, since light is
80 | P a g e

distinguished by wavelength. Let's say you have an atom with the electron in the
lowest energy level (a hydrogen atom is the simplest to picture). Most wavelengths of
light are ignored by the atom. If the exactly right wavelength comes along, the
electron will absorb the energy (absorb the light) and the electron will bounce up to a
higher orbit. The higher the energy of the light it absorbs, the higher the orbit it ends
up in. Remember, it has to be the exact right type of light for this to happen. It will
ignore all of the types of light that are wrong. When the electron moves to a higher
orbit (higher energy state), we refer to the atom as being excited.
Electrons don't want to be in higher orbits; they like to be in a low energy state
(mainly the ground state) - remember, they are lazy and like to have low energies.
They want to get rid of any energy they pick up. To do this they give off the light
needed to go down to lower energy levels. This light that they give off could be
exactly like the light they absorbed or a combination of different types of light that
add up to the same energy as the light that was originally absorbed. The end result is
the atom gets rid of the energy it sucked up (it gives off the light in a random
direction) and the electron is back down in the ground state. For a while the atom was
storing up extra energy - this is when you would call it an excited atom - not because
it won the lottery, but because it has extra energy in it. This scenario is shown in
Figure 3.
Figure 3. An atom is shown with the electron in the lowest
(ground) energy level. Light of just the right wavelength is
absorbed by the electron, causing it to go into a higher energy
level. As long as the atom has this extra energy it is referred to
as being "excited". To get back down to the lowest energy
level, the electron has to get rid of the energy and does so by
emitting light with the appropriate energy to make the
transition. Now it is just a plain unexcited atom again.
What happens if the energy the electron absorbs is too
much? Is there such a thing as too much? You bet there is. The
atom has a structure that allows for orbits up to only a certain
energy level - sort of like the top of the mountain. If the
electron absorbs too much energy and then goes up, there is no
level that it can land in, since it has too much energy. This
would be sort like having a rocket pack on your back as you
climb a mountain, like what often happens to the Coyote trying to catch the Road-
runner - he overshoots his target. The electron goes completely out of the atom. The
atom is then said to be ionized; see Figure 4 for an illustration. Now the atom is sort
of out of balance, since there was an electrical balance between the electrons and
protons. With the loss of an electron, there is one less negative particle. For an atom
81 | P a g e

like hydrogen, there is only one electron that can be lost. Other atoms have more
electrons and can lose more and more, so long as the energy available is high enough
to keep kicking the electrons out. Roman numbers are put behind an atom's name to
distinguish ionized and unionized atoms. For example, regular, everyday, normal iron
atom is refered to as Fe I. If it loses one electron, it becomes Fe II; lose another
electron and it becomes Fe III, etc. If iron lost all of its electrons you would have Fe
XXVII, since it has 26 of them. If the electron wants to get back together with the
atom, or any other atom, it has to get rid of the energy it sucked up. Under the right
circumstances it can give off light and then can go back to an
atom.
Figure 4. An atom gets ionized, when light with high enough
energy can knock the electron out of the atom completely. The
light is absorbed by the electron so that it is able to escape from
the atom.
This is all exciting (yeah, right) but what about the spectra? Just
a minute, I'm getting to that. Let's look at another situation.
What if you have a bunch of atoms that are not getting
bombarded by light but are just moving around a lot - they are
perhaps in a container that is getting heated up. When you heat up gasses, the atoms
move around faster. This is not good, since there will be a bunch of collisions. Does
this mean their insurance rates go up? No, this means that there is a lot of energy of
motion involved. You know what atoms like to do with energy, right? The electrons
can take the energy of the collisions and change their orbits. To get back down to the
low energy levels they have to give off light, so they do just that and go back down to
their ground state. This is a situation where you can get light from a bunch of atoms
without putting light into it in the first place; all you have to do is heat it up
(technically, heat is energy, which is what light is, but it's not what you think of as
light, right?). If the atoms are hot enough, they'll produce their own light. The atoms
could be ionized or excited in this case, depending upon how fast they are moving and
how much energy is involved in the collisions.
See Figure 5 for the scenario.
Figure 5. Two atoms head for a collision. The
energy of the atoms' motions is translated into
getting the electrons bumped up to higher
energy levels. In one case, the electron just
moves up to a higher energy level and the
atom is therefore excited. In the other case, the
electron is completely popped out of the atom,
resulting in an ionized atom. The electrons
82 | P a g e

will have to eventually get rid of the energy they have - by possibly giving off light -
since they want to get back down to the ground state.
What about solid objects? When atoms get really, really close together where they are
pretty much solid, or thick enough that they are pretty close to solid, then the levels of
the atoms get really screwed up. They are sort of mushed around so that they aren't
too picky about the type of light that they'll absorb - they'll absorb longer and shorter
wavelength light and even longer and even shorter as the material gets denser and
thicker. A solid can absorb pretty much all wavelengths of light, while something less
dense, like a gas, will only absorb (or emit) certain types of light. A solid can also
give off light at pretty much all wavelengths if it is heated up. For a strange reason
physicists refer to a solid light source as a black body. Yes, that's right, an object that
is giving off light because it is heated up is called a black body. This is in part due to
the fact that such objects also absorb light, like how dark clothing is worn in cold
weather to keep you warmer. There are a bunch of complicated formulas that define
how "true" black bodies should behave. In reality, a perfect black body doesn't really
exist in nature, but the rules that govern how a black body absorbs and emits light are
useful in describing a lot of things, so it is often easier to just pretend that a lot of hot
solid objects (like stars or heated up pieces of rock or metal) are black bodies.
Are you completely confused by now? If not, then you must have been skipping the
really confusing bits. If you are confused, let's take a look at the rules that explain the
various types of spectra. The ways in which light is absorbed and emitted by matter
were studied and explained by a fellow named Kirchhoff, so we refer to these rules
as Kirchhoff's Laws - these are the rules that explain how spectra are formed and
define the different types of spectra.
1. A hot, opaque source (black body) gives off a continuous spectrum - this is
like the complete rainbow, with no gaps or breaks in it.
2. A heated, thin gas will produce an emission spectrum made up of discrete
lines. In this case, the distinct types of light associated with the atom's energy
level are seen, so there are often a lot of empty spaces (dark areas) in this type
of spectrum.
3. When a hot, opaque source (black body) is viewed through a cool gas, light of
only certain wavelengths is absorbed and then randomly given off (scattered)
by the gas. This produces an absorption spectrum - this is the case where you
get most of the light, but there are dark gaps in places where the light is
missing.
83 | P a g e

Figure 6. The different types of spectra are shown - continuous, emission and
absorption.
Figure 6 shows the various types of spectra. Often to denote a black body, a light bulb
is used - remember, a black body is sort of a hot solid object like the filament in a
light bulb. Again, it should be noted that there really is nothing in nature that is a true,
perfect black body, but that some things do come close enough - at least close enough
for us to use the formulas that describe black bodies. For the absorption spectra it
should be noted that not all of the light that is absorbed by the cool gas cloud remains
in the gas cloud. When the atoms in the gas cloud absorb the light, the electrons will
be momentarily excited, but as you know, they don't like to be excited. Therefore, the
energy that the atoms absorb does get eventually released, but the light is given off in
random directions. The case of an absorption spectrum might be thought of as a
"deflection spectrum" since the light is mainly just deflected off in random directions
by its interaction with the atoms. In the case of the emission spectrum, something is
causing the light to be produced by the gas itself - usually heat - and this causes the
light to come out based upon the structure of the atom. By making the gas thicker and
thicker, you would get thicker and thicker emission lines until you get to the point
where the thing is solid and now you are looking at a black body and have a
continuous spectrum.
84 | P a g e

In the case of Kirchhoff's laws 2 and 3, the different atoms in the gas give different
patterns of lines corresponding to the type of light they absorb or emit. The type of
light an atom absorbs or emits depends upon the specific structure of the atom (how
far apart the energy levels are). In other words, a certain type of atom will produce a
certain pattern of lines (dark absorption lines or bright emission lines) and each type
of atom makes its own distinct pattern. An atom of hydrogen is different from an atom
of helium, which is different from a carbon atom, which is different... well, you get
the picture. See Figure 7 for a variety of spectra from some gases.
Figure 7.The absorption and
emission spectra for different
gases are shown. Notice how the
spectra are sort of negatives of
one another - what is absorbed
in one spectra for a particular
element is emitted in the
emission spectra for that
element.
If the pattern produced by the
emission or absorption spectrum
is different for each element
(and it is), then by looking at a
spectrum, you can identify the
gas that is involved in the
production of the spectrum. This is just like fingerprinting the atoms - here's a method
of obtaining element identifications! When astronomers try to determine the
composition of a cloud of hot gas, a star or a galaxy, they use the spectra to provide
information about the object. We don't even have to touch the star to figure out what it
is made of, which is a good thing, since those critters are really hot. This is how
astronomers can talk about what a distant galaxy is made of, though they have never
had the chance to sample it directly - but just had the opportunity to look at its
spectrum. In reality it is a bit more complex since objects are generally made up of a
bunch of different elements - so all of those elements would be visible in the
spectrum. In some cases it is even possible that the spectra have both absorption
features and emission features - which is really complicated. An example is shown in
Figure 8.
Figure 8. Emission spectra for
three elements are shown
(Hydrogen, Carbon and
Oxygen). If all of these elements
85 | P a g e

were in a hot gas cloud, they could produce the spectrum at the bottom which is a
combination of the three individual spectra.
More Radiation laws
The situation where you have a continuous spectrum is particularly interesting. To
produce such a spectrum an object must be a hot solid object - or at least close enough
that it appears solid for the most part. If it is a perfect emitter and absorber of
radiation, we call it a black body. As if I haven't already stressed this enough, in
nature, objects are not perfect, so there are no true black bodies out there, though
some objects are close enough that we can use the laws of black bodies to help
determine other characteristics of these objects. One of the laws has to do with the
way that energy is emitted by a black body. The energy outputs for different
temperature black bodies are shown in Figure 9.
Figure 9. Several energy output
curves for blackbodies. The amount
of energy given off at different
wavelengths is shown. Note how the
curves all peak at just one
wavelength. This peak corresponds
to the dominant type of light that is
given off by the black body.
You might note that there is a peak
for the energy output. The peak is at
a different wavelength for the
different temperatures of the black
bodies. The higher the temperature,
the shorter the wavelength of the
peak. Hotter things produce mainly
blue light, since short wavelength
light is bluer, while cool things
produce mainly red light. You may have seen this type of phenomenon while
watching something heat up in a fire, like a piece of metal. First, the metal is dark and
cool, then it is reddish, then orangish, then yellowish, then whitish-bluish - the hottest
it will get. This color change is due to the change in the energy output peak - shorter
wavelength for hotter things. Of course, very hot things could have their peaks in even
shorter wavelengths, such as UV or x-ray, while very cool things could produce most
of their light in wavelengths such as IR or radio. In these cases we would not be able
to see with our eyes the main type of light that is given off, so it would look quite
86 | P a g e

different to us. The relationship between the temperature and the peak wavelength (or
color) is known as Wien's Law, and is given by this formula -
max
= 0.0029/T meters
where the temperature is in K and the resulting wavelength is in meters. The basic
upshot of this formula is that when the temperature goes up, the wavelength where
most of the light comes out at gets smaller (bluer). By using this formula it is possible
to determine the temperatures of objects, or at least estimate them, so long as you can
observe the peak for the energy output and you assume that the object is acting like a
black body. Sometimes astronomers can do this, especially if the light comes out at a
"normal" wavelength that we can detect. If you go back to the table listing all of the
different types of light, the last column in the table refers to the temperature range
needed to produce that type of light.
I should mention that the energy outputs of objects are quite diverse. The curves
shown in Figure 9 indicate that energy is given out at many wavelengths, not just at
the peak wavelength. If you can't see the energy at the peak wavelength due to a lack
of the proper equipment, it is probably still possible to see the object - it may not look
as bright as it would at its peak wavelength, but it should still be visible to some
degree in the range of light that you can detect.

Now take a peak at Figure 9 again. You may have noticed that the curve for each
temperature is different. The sizes of the curves varies with temperature - the higher
temperature objects have much larger curves. This is an indication of the total amount
of energy given off. Hotter things are also producing a lot more energy or light than
cooler objects. Here's another formula telling you how much total energy is produced
-
E = T
4

where is a constant and the energy emitted is the number of Watts per square meter
given off. This law is known as the Stefan-Boltzmann Law and is primarily applied
to stars, but can be used to estimate the output of energy from most objects. A rather
interesting aspect about black bodies in general, is that the composition of the object
doesn't influence the energy output. Temperature is the only thing that influences the
energy output of a black body. You should make note of the effect of changing the
temperature in both these laws. As the temperature goes down, the peak wavelength
gets longer (goes up) and the total energy output goes way down (that power of four
on the temperature is pretty strong!). What if you were to triple the temperature? The
peak wavelength would be three times shorter (1/3 the original value), while the
87 | P a g e

energy output would be 3x3x3x3 = 81 times greater! A small change in temperature
has a big effect on the energy output!
Doppler Effect
Another thing that you can get from an absorption or emission spectrum is the
velocity of the object producing the spectrum. This is due to the Doppler effect. This
is the same effect you experience when an ambulance or a train passes by. The speed
of the object affects the sound waves coming from it, so the sound of the ambulance
or train whistle changes as it goes by. You could also get a similar effect if you are the
one that is moving, and the sound source is stationary; basically, the sounds when you
approach and recede are different.

The same effect is seen in light waves, since light has wave-like properties (you didn't
forget that, did you?). When an object is coming towards you, its light waves get
squeezed in, causing the spectral features to shift to shorter wavelengths (be bluer than
normal), while an object moving away would have its light stretched out and the
spectral features would be shifted to longer wavelengths (be redder than normal). This
situation is shown in Figure 10. In cases of very extreme velocities it is also possible
for the object to appear a different color because the shifts are so large.
Figure 10. A light source is moving towards
the left. The peaks of the light waves are
shown as circles that are moving away from
the location they where they were first given
off by the light source. So the largest circle
represents the first light emitted by the
source. In the direction of its motion, the
light waves are compressed, so that Observer A would see the light source bluer than
it actually is, while Observer B would see the light source redder than normal (since
the light in that direction is stretched out to longer wavelengths).
Astronomers describe the spectrum as being either red shifted or blue shifted. This
just means that the light has been shifted due to, respectively, relative motion away
from you or motion towards you. The degree of the shift depends upon the speed of
the motion. You can determine the speed using this formula -
V =c ,
where V is the velocity (in km/s), c is the speed of light (=300,000 km/s), is the
change in wavelength of the spectral features from the normal wavelengths and is
the normal wavelength. A larger shift results from a large velocity. Figure 11 shows
88 | P a g e

how the Doppler shift would cause the absorption features in a spectra to appear at the
"wrong" wavelengths.
Figure 11. Absorption spectra
shown under three different
circumstances. At the top, the
normal non-moving spectra with
the features appearing at their
appropriate wavelengths. In the
middle, the object is moving
away from the observer (the
person who obtained the spectrum) so that the absorption features appear at slightly
longer wavelengths than normal. The bottom spectrum is the case where the spectral
source is moving towards the observer so the features are blueshifted - occurring at
slightly shorter wavelengths than normal.
Telescopes
Telescopes are the main tools of astronomers and have been used since the days of
Galileo to study objects in the sky. Of course since there are different types of light,
there are also different types of telescopes. Regardless of the differences, telescopes
have three purposes:
1. To collect light. Believe it or not, this is the most important purpose. As we
like to say in astronomy "Size does matter - the bigger, the better." In fact, size
can make all the difference. You may want to think of a telescope as a light
bucket and the more light it collects, the better the view. The amount of light a
telescope collects depends directly upon the size of the light collecting surface -
the part of the telescope that gathers up the light. The light gathering ability
goes as the radius of the light gathering surface squared, since telescopes
usually have a circular light gathering surface and the size of this surface is a
function of radius squared (Area of a circle= R
2
). If you were to double the
radius, you would increase the light gathering area by a factor of four - you
would gather four times more light. However, bigger also means that you have
to have some pretty big machinery to operate the telescope, as well as a pretty
big wallet to cover the cost of the thing.
2. To see fine detail - to resolve features. The ability to see things more clearly,
or to have better resolution, is usually measured in arcseconds, and the smaller
the resolution, the better. This just means that you can see things that are pretty
close together or that you can see small details clearly. Resolution is important
since it helps astronomers get a better idea about what is going on out there -
small details could provide a great deal of information about objects and the
89 | P a g e

processes that affect them. Unfortunately, the resolving ability of a telescope
depends upon the type of light (wavelength) you are using and the size of the
telescope you are using. For the best resolution, a very large telescope is
needed, and in many cases this is difficult to manufacture (or afford). In some
cases having a large telescope doesn't really help - that's because the Earth has
this annoying atmosphere that not only blocks our view of some things
(especially on cloudy days) but also blurs our views of objects. There are some
ways to get around these problems, as you'll see.
3. To magnify objects. While many people think that this is what telescopes
mainly do, it is usually not a major feature. If you were to sit two inches from
your television everything would look bigger, right? How is the quality of the
picture? Pretty bad, isn't it? Often when the magnification is increased the
image looks pretty cruddy, so sometimes a lower magnification is better.
Usually only a high magnification is used when looking at nearby object like
planets or the Sun.
Types of Telescopes
The first telescopes built were made up of a series of lenses, sort of like a bunch of
eyeglasses lined up. These lenses would bend the light to bring it to a focus at one
location. I suppose we should call these telescopes benders, hmm? No, that's not very
scientific. When light is bent by passing through something like glass, it is refracted,
so these early telescopes are known as refractors. Figure 14 shows you a refracting
telescope design. You probably have a refracting telescope in your home even if you
don't know it - that is if you own a set of binoculars, since they operate mainly by
using this method. The old fashioned refractors used to be fairly popular, since it was
sort of easy to make lenses.
Figure 14. A refracting telescope operates by
bringing light to a focus using a lens or several
lenses. The light collecting area of these types of
telescopes is limited.
Here's a good question - why do we want to bring light to a focus? Let's say you have
a refracting telescope that has a 3 inch wide lens. All of the light falling in that 3-inch
wide lens is being gathered by your telescope. How are you going to put all that light
into your eye? Your pupil is pretty small, only a few millimeters wide at most. It's a
waste of time if you can't get all of that light you are gathering into a convenient
package that fits into your eye. The light needs to be gathered or focused so that it will
fit into your puny little pupil. That's why we use the lenses to bend the light into a
focus.
90 | P a g e

Unfortunately, there is a limit to the size of a refracting telescope, since the only
support of the lenses is on the edges, just like eyeglasses. If a lens was very large, its
weight would cause the glass to sag and the images produced by the refracting
telescope would be distorted. Refracting telescopes had a size limitation that could not
be overcome, at least not without sacrificing the light gathering area. Since size is so
important (remember,size does matter), this is a pretty bad situation. One of the
largest refracting telescopes is at the Yerkes observatory in Wisconsin (see Figure 15).
The lens in this telescope is 40 inches wide - about as big as you can get.
Figure 15. Yerkes observatory refracting telescope.
Note how long the scope is compared to the amount of
light that it gathers. This long length is needed to
focus the light properly. (Photo from the Yerkes virtual
tour)
To overcome the size limits, another telescope design
is usually used. If you have a curved surface which is
highly reflective (like a mirror), the light could be
brought into focus by having it bounce off the mirror
at a certain angle. Since a mirror could be supported
from behind, you can make a very large mirror! To
focus the light, the best surface to use is a parabola
(Figure 16), and this scheme is used in most large
modern telescopes, not just those that you look
through with your eye, but also radio, IR, etc.
Currently most modern telescopes are of this type -
reflecting telescopes.
Figure 16. Basic design for a
reflecting telescope. The curved
mirror is in the shape of a parabola
and this allows it to bring all of the
light to a focus in one location.
You might have noticed that the light
in a reflecting telescope is focused at
a point in front of the mirror. If you
wanted to look at the gathered light,
you would have put yourself in front
of the mirror. This doesn't seem like
a very logical way to observe
objects, since your head would be
91 | P a g e

blocking some of the light. Actually, with some telescopes, this doesn't really make
much difference, since the size of the light gathering area is huge in comparison to the
size of your head. Often instruments can be placed at the main focus location, or the
light can be bounced around by other mirrors to take it to a more convenient location.
This is shown in Figure 17.
Figure 17. Different types of
reflecting telescopes. Some are
configured so that instruments or
even observers are located
where the light is focused, while
other use mirrors to deflect the
light to different locations where
instruments or astronomers are
located.
Visible light reflecting
telescopes are currently limited
only by the weight of the mirror
and the ability to aim and
support them. Also the amount
of money available to build such
large instruments and to keep
them in working order. This last
item may be the only real limit
that we have now on the size of
telescopes - the budget. Physical
limits are overcome by a unique
design, such as using not just one mirror, but a bunch of little mirrors that work
together and mimic one huge mirror. Also, recent methods in making the mirrors
themselves lighter has helped the process. Among the largest telescopes currently in
operation or currently being built are the following -
The Gran Telescopio Canarias (GTC) is currently the largest single-surface
telescope. Since making one giant mirror in not practical, this 10.4 meter wide beast is
made up of 36 segmented mirrors that are placed together. And in case you are not too
familiar with the metric system, 10.4 meters is equivalent to 34 feet. So this
telescope's main mirror is a little bit wider than 34 feet. The telescope is located on the
Canary Islands, off the coast of northern Africa. It is operated by a European
astronomy group, and is usually only used by astronomers who work in Europe, so
you may never have heard of this telescope. The size of the building that houses the
telescope is just under 20 feet shorter than the Statue of Liberty in New York City -
92 | P a g e

I'm not sure if that is something you are familiar with, but it is certainly higher than
any grain silo in the state of Iowa.

One of the 36 mirror segments that will go
into the Gran Telescopio Canarias.

The schematic for the GTC, showing the overall
width of the main mirror of 10.4 meters (34.1 feet)
Keck telescopes - these telescopes are nearly as big as GTC at 10 meters. What is
unique here is that there are actually two telescopes, each 10 meters wide. They have
the same design as GTC and are located on a mountain in Hawaii.

The Keck Telescopes. The two spherical domes
contain the two telescopes that make up the
Keck Observatory, the observatory to the left
houses the Subaru Telescope. Image courtesy
of the Keck Website

The main mirror of the Keck Telescope -
multiple mirrors combined to act like a single
mirror. Image courtesy of the Keck Website
93 | P a g e

The Keck telescopes and GTC both use multiple pieces of glass to form the main
mirror. The Large Binocular Telescope (LBT) is actually two mirrors, each 8.4
meters wide, and each made out of a single piece of glass. These mirrors work
together, which effectively allows the collection of light to double.
Very Large Telescope - the winner of the "most unoriginal name in astronomy"
contest. This is a big project where not one, not two, but four telescopes were built so
that they can work together to mimic an even larger scope. Each of the scopes is 8.2
meters in diameter and when they are all working together, they operate as if they
were a single 16.4 meter diameter scope. The VLT is in Chile.

Large Binocular Telescope-
here is a view of the LBT
showing the two 8.4 meter
mirrors. Image courtesy of NASA

Very Large Telescope- The 4
telescopes that make up the
VLT are shown. Photo
courtesy of the European
Southern Observatory.

South African Large
Telescope- The largest single
telescope in the southern
hemisphere. Photo courtesy
of SALT.
SALT - The names stands for South African Large Telescope. This one is designed
like the GTC and Keck, so it is made of many mirror segments working together. It is
about the same size as Keck, but is of course in the southern hemisphere, so that
makes it the biggest single optical telescope south of the equator.
94 | P a g e

While these telescopes are relatively new, the old standard for a big telescope was 4 or
5 meters in size (13 - 16 ft). This used to be the size of the "big" telescopes available
for research, but with new mirror design and manufacturing processes, it is possible
that 10 meter mirrors will be built in the future. Actually, some people are even
working on telescopes with 30 or more meter wide mirrors! Stay tuned for record
breakers in the "biggest optical telescope" race over the coming years.
At UNI there are two observatories. Hillside observatory's telescope has a diameter of
16 inches, while the McCollum Science Hall telescope is 12 inches in diameter. They
aren't 4 or 5 meters in size, but they are pretty nifty and relatively powerful
instruments.
You may be wondering how you can have multiple telescopes working together; after
all, astronomers can't look through both scopes at the same time, can they? Actually,
they can. In the old days they just sat behind a single telescope and used their eyes to
get a view of objects. Generally they had to be pretty good artists since that was the
only way to record the information. They had to make sketches of everything that they
saw. Some of the sketches are pretty good, which is rather impressive considering
how cold, dark and difficult it is to draw sitting behind a telescope. Life for
astronomers got easier around the late 1800s, when photography became the main tool
for data collecting. Previously unseen, very faint objects were then detectable, since a
photographic image could be made from a long exposure of the film. Long exposures
are rather difficult to get, since you need to have your telescope aimed at the same
object over the course of the evening, and since the sky has this rather annoying
tendency to appear to move, it isn't easy. Probably more times than can be imagined,
an astronomer would stay up all night taking a seven hour picture, only to have
something come along and screw it up.
Nowadays, to increase the light gathering abilities of telescopes and to combine the
data (light) from different telescopes together, and to make it easier to manipulate the
data, electronic detection devices are often used. These are similar to the digital
equipment that people have in things like plain old digital cameras, webcams and
most video cameras. The most common type of electronic detector is a CCD,
or Charge Coupled Device, which is really just a very light sensitive digital camera
that translates images into a computer format which is easy to analyze or manipulate.
It is also possible to take several different images from different locations and
combine them with computers for a better image or to analyze the differences in the
images. The only real disadvantage of CCDs is that they are generally pretty small
and can usually only take a picture of a tiny part of the sky. Even small ones can cost
several hundred dollars, while the biggest ones can cost several million dollars! Often
when you see a modern astronomical picture of something in the sky, it isn't a
photograph but a digital picture.
95 | P a g e

(Right) One of the new CCD chips installed into the
Hubble Space telescope in 2002. Each chip is very
sensitive to light and can obtain high resolution, large
scale digital images of the sky. Image courtesy Hubble
Space Telescope.
There is a telescope project planned which will survey the
sky using a 8.4 meter telescope in Chile. This project,
the Large Synoptic Survey Telescope, will cover the
entire visible sky each night by taking two pictures every
30 seconds. This will result in the creation of 36
gigabytes of data every 30 seconds. Read that sentence
again - it said gigabytes. So in a few minutes enough data
would be obtained to pretty much fill even large iPads. How can they obtain so much
data? Well take a look at the camera they have - it is huge!!! And you thought your
own digital camera took good pictures. This project is producing so much data that
Google will be involved in helping to distribute and store it. The telescope is expected
to be ready to go by 2014.
(Left) A representation of the CCD
chips to be used in the Large
Synoptic Survey Telescope project.
Each of the 200 chips has the
capacity to take a 16 mega-pixel
image. Image courtesy LSST
Corporation
Other Types of Telescopes
The same parabolic shape used for
visible light telescopes can also be
used for radio telescopes, but since
radio wavelengths are larger and
easier to catch, we do not need a
solid surface (mirror) to reflect them
(see Figure 18). Many large radio
telescopes can use a wire mesh (basically, chicken wire). This structure has less
weight per square meter of light gathering surface, and since metal is easier to shape,
radio telescopes are the largest of all the different types of telescopes. If you have
satellite television or any one of those dish-tv systems, you have a radio telescope -
same design, though in this case you're getting 17 different football games
simultaneously as opposed to scientific information.
96 | P a g e

Figure 18. A typical radio telescope. Light is reflected off
the curved surface and collected at the receiver at the top.
This information is then fed into a computer to produce an
image of the sky. Notice that this is not a solid surface, but
a wire mesh. Photo by Seth Shostak from the Parkes Radio
Telescope (CSIRO).
The largest single dish telescope is in Arecibo, Puerto
Rico. This telescope is so large (300 m across) that it
doesn't even move - it just sits in a valley and the rotation
of the Earth allows it to view the sky. There is also some
minor pointing ability with the focusing part of the
telescope. Usually radio telescope dishes are 50-100 m in
size.
Arecibo Radio TelescopeCourtesy of the NAIC -
Arecibo Observatory, a facility of the NSF. Photo
by David Parker/Science Photo Library.
You have to remember that radio light is not
something that your eyes can detect. Once the
radio telescope has gathered up the radio light, it
has to show you the picture somehow. This is
where computers come in really handy. They can
translate the signal into an image and also allow you to do some really nifty stuff with
the data. Radio telescopes can be linked together over great distances - sort of like
having stereo vision. This is known as interferometry and it provides greater
resolution in the images. It is easy to combine signals, even if the telescopes are
spread out over great distances, since all of the data that radio telescopes get are
digitized. This results in very high resolution images. Radio telescopes can have
resolutions that are 1/10,000 arcseconds, while visible light telescopes like the Keck
or Gemini may have resolutions on the order of a 1/10 arcsecond. You can also use
the interferometry trick with visible light telescopes, since they are becoming more
and more digital-based.
97 | P a g e


Figure 19. Two views of the same object. An optical (visible light) image of a supernova
remnant, the remains of a star that blew up, is shown on the left. The image on the right is of the
same object, but taken with a radio telescope. This view shows more structure and the bubble
shape of the gas cloud; however, it doesn't show the stars that are seen in the image on the left.
This is because stars are generally not very strong radio light sources - they are usually too hot.
Image from the Chandra website, NASA/CXC/SAO.
You'll want to take another look at Figure 19. The image on the left is what your eye
can see, while the one on the right is a false color image based upon radio light
intensities. This is similar to what you see on the local news when they show radar
images of storms. In those images the maps are color coded to show the location and
concentration of rain, or snow or something similar. In astronomical images from data
that our eyes can't see (like most types of light), you are seeing usually a
concentration or intensity of light. The actual color of the object isn't being shown, so
don't expect there to be large clouds of blue, red and green material floating out in
space. It is also likely that images taken with visible light telescopes are manipulated
to show certain features, and they can also be enhanced in artificial ways. So basically
most modern astronomy images are "altered" in various ways to help astronomers
learn specific details.
A large group of radio telescopes that work as if they were one large dish is found in
New Mexico at the VLA (Very Large Array- another one of those really exciting
telescope names). In this case, there are 27 individual radio telescopes that can act like
98 | P a g e

a single large one. Currently an even larger group of telescopes is being built in Chile.
Eventually there will be 66 individual radio telescopes for the Atacama Large
Millimeter/submillimeter Array. And plans are underway to build radio telesope
arrays covering even larger areas, so stay tuned.
The Very Large Array (VLA)- Photo
courtesy of Dave Finley, National Radio
Astronomy Observatory and Associated
Universities, Inc.
You may have noticed something about
the telescopes shown so far - a lot of
them are located in rather remote
locations, especially the visible light telescopes. This is because astronomers are so
annoying that people can't stand them - no, that's not the real reason. Actually, to
really study the sky, you need a really good view of it. The best possible view is one
away from cities, preferably in a place where there are few cloudy skies and low
humidity. That's way there are so many observatories located on tops of mountains,
since the humidity is pretty low up there many times and there is less atmosphere to
look through. Amongst the best sites in the world for having observatories are in the
southwest US (Arizona and New Mexico), the top of the mountain Mauna Kea in
Hawaii (which is a hopefully extinct volcano), the mountains of Chile and the desert
regions of Australia and many other countries. Radio telescopes aren't so picky; they
can be found in more locations, since radio light isn't as sensitive to things like city
lights as visible light it, though you often need a large area for a radio telescope.
Figure 20. An image of the USA at night
as seen from space. The light seen is due
to city lighting for the most part. Of
course, all of this light seen from space
is wasted electricity - why would
someone want to light up the sky? This
also shows why astronomers would want
to put a telescope in a place like Arizona
or New Mexico where there is less light
pollution. If you click on the image you'll
see the full sized view.
Radio and visible light are only two types of light that exist. What about the others
wavelengths of light? How do astronomers study them? We would like to study them,
but there is this rather annoying thing known as the atmosphere. The atmosphere
prevents other types of light from getting to the surface of the Earth, and it also has
99 | P a g e

the annoying tendency to limit the quality of our view since it can blur stuff quite a
bit. Pretty much all of the other types of light are blocked out at sea level. It is
possible to detect some of the wavelengths of infrared light at very high elevations.
Apart from that, you can't detect any ultra-violet, x-ray or gamma-ray light sources
from the surface of the Earth. How do astronomers study those types of light? We
have to get above the atmosphere - we have to send telescopes into space.

There have been many satellites that have been used by astronomers, but we'll look at
only the "biggies" here. Except for the gamma-ray and x-ray telescopes, the telescopes
use the same old parabolic mirror design to gather light. Of course, since the telescope
is in space, the images are obtained generally by using digital cameras or a similar
detection device, so the data can be easily translated into a format that our eyes can
handle.
Infrared Satellites
IRAS, (operated in the early 1980s) - The Infrared Astronomical Satellite took many
images of gas clouds and other low temperature objects. This was the first major
telescope to operate at this wavelength. It was able to do an initial survey of the entire
sky in infrared and the data is still being used today.
The main infrared satellite up there is Spitzer (launched 2003), however it has
recently run out of coolant, so it is limited in its ability to measure some of the lowest
energy IR light. In May 2009, the European Space Agency launched Herschel, a
telescope with better resolution than Spitzer, but it has since run out of coolant, so it is
no longer working. Another recent IR satellite is WISE, which is a survey telescope,
similar to IRAS. It was launched in December 2009, but only lasted until February
2011. During that short time it did find may low temperature objects including some
of the coolest stars. Some IR astronomy work has shifted to the tops of high
mountains, since it is also possible to do some IR astronomy work in that format, and
it is cheaper than a satellite. Click here to see a comparison of a region in space where
stars are forming. Such regions have a lot of cool gas and dust, and the stars tend to be
rather cool, so they show up better at IR wavelengths than at visible wavelengths.
Note how some stars are only seen in the IR image. These two images are from the
Hubble Space Telescope (courtesy of STScI/NASA).
Ultraviolet
IUE (1978-1996) - The International Ultraviolet Explorer was one of the most
successful satellites to be launched. It outlived its expected life span and provided a
great archive of relatively high temperature objects in the universe. Why isn't it still
100 | P a g e

working? Money, money money - due to budget cuts the satellite was turned off.
What a waste!
The only current large UV telescope is Galex (launch 2003). Often UV satellites have
specific purposes, for example, Galex's primary targets are galaxies (hence the name).
Recently one of the detectors on this spacecraft has stopped working, but it is still
obtaining data, and plans are in place to keep it working as long as possible.
Ultraviolet telescopes tend to look at very hot objects, things like ultra hot stars or
stellar explosions are popular objects as well as material heated up to extreme
temperatures around stars.
X-ray Satellites
There have been many satellites in the past that were x-ray missions with a variety of
names, like Uhura, Einstein and ROSAT. Even though these missions are completed,
their data is still being studied and analyzed. One of the tricky things about these
satellites, as well as the gamma-ray ones, is that they don't use a regular mirror.
Remember, x-rays go through things, so they would go through a mirror. How does an
x-ray satellite work? It is sort of like how the police trace the path of a bullet - they
detect where it went through various surfaces and therefore they can trace it back to
its origin. The same is done with these telescopes. They work by tracing how the x-
rays traveled to figure out where the x-rays are coming from. It's sort of weird, and the
images aren't always that sharp, but it does work.
Chandra (1999-now) is a current x-ray telescope that is in orbit. Its views of the sky
show astronomers how various high energy objects, like active galaxies and
supernovae, are behaving. Also the NuSTAR mission (2012-now) will be looking for
very high energy x-rays, possibly indicating the presence of black holes.
101 | P a g e




Images of the same object taken with four different telescopes. In the upper left is the optical
(visible) light telescope image of a cloud of material from a star that exploded about 950 years
ago. In the upper right is the same object, but as seen with a radio telescope. The lower left
shows an infrared view of the object, while the lower right shows the x-ray view. Even though
this is the same object, different features are seen, or not seen, depending upon which telescope
is used. Images from the Chandra website, courtesy of Palomar observatory, NRAO, W. M. Keck
observatory and NASA/CXC/SAO.
Gamma-ray Satellite
Due to the very high energy nature of gamma-rays, there have not been many
satellites that work in this part of the spectrum, so the observations of any gamma-ray
observatory are of great interest. A current mission isIntegral (launched 2002), which
stands for International Gamma-ray Astrophysics Laboratory - talk about stretching
an acronym out! Objects such as the mysterious gamma-ray bursters and supernovae
are the targets of interest of this satellite. Another recently launched gamma-ray
satellite is Swift (launched Nov., 2004) which also has ultra-violet and x-ray detection
102 | P a g e

capabilities. Swift was designed specifically to catch gamma-ray bursts, which happen
so quickly that by the time a regular telescope is aimed at them, they are over. So
far Swift has detected dozens of these events and this information is helping
astronomers figure out what gamma-ray bursts really are. To see a all sky map of
recent gamma-ray bursts, just click here. Along with these there is the recently
launched (June 2008) Fermi mission, which has made some amazing studies of
gamma-ray and x-ray sources.
Visible Light Satellite
The Hubble Space Telescope (HST) -
image courtesy NASA, STScI. Many of
the pictures in this course are from this
telescope. You can check out many of the
images if you visit the HST website.
While it is not necessary to put a visible
light telescope in space, the need for one
is obvious. The atmosphere has a
tendency to blur images, so by placing a
telescope above the atmosphere, the
clarity of the images is increased a great
deal as well as the ability to gather more
light. The currently largest optical telescope is the Hubble Space Telescope -
HST (1990 - ). While the mirror of the HST is only 2.5 meters in diameter, its
location in space makes it a very powerful tool. Soon after it was launched it was
discovered that the telescope's big mirror was not in the correct shape so that the light
did not focus. To give you an idea of how picky these telescopes can be, the error in
the HST mirror was 0.000002 meters. Even though this is a pretty small amount, it
really screwed up the telescope. It had to be fixed in space. A later space shuttle
mission went up to fix it and now it works just dandy. Recently a new camera was
also installed, which will allow it to take even better pictures than before. The HST is
by far the most annoying telescope ever built! Why do I say that? Mainly because of
all of the new things it discovers (and I have to keep re-writing my notes!). The HST
is supposed to last until about 2010, but could last longer if we're lucky or unlucky if
you have to teach astronomy and it keeps discovering new stuff. Recent repair
missions have extended the life of the HST, hopefully until we can replace it with an
even bigger visible light telescope. And just to show how big things are now, the
previously mentioned IR space telescope, Herschel, has a 3.5 m mirror. The telescope
that is planned to replace the HST is the James Webb Space Telescope, which is
scheduled to be launched in 2014. But until it is up there, we don't have to worry
about it (and I don't have to revise my notes again).
103 | P a g e

Future Telescopes - Right now there are plans on making much larger and obviously
much more powerful telescopes. As previously mentioned there are plans on making
visible light telescopes that are larger than 10 meters in diameter, and even plans for
the next version of the Hubble Space Telescope. While bigger telescopes may be
better, there are other ways to improve images of objects, especially when viewed
from the Earth where the atmosphere causes so much blurring. The latest idea is
something called adaptive optics. This technique involves compensating for the
blurring effects of the atmosphere by altering the light that is collected so that it is not
as smeared out. This can be done in a variety of ways including, having telescopes
change their shapes to compensate for the atmospheric effects. In some cases the
images obtained in this method are better than those obtained by the Hubble - take a
look at the images shown below to see what I mean. Expect not only bigger telescopes
in the future, but also much more sophisticated methods of dealing with the light -
whether that involves different collecting techniques or analysis techniques, we'll just
have to wait and see.

Image from the Hubble Space Telescope
showing a bright star and surrounding stars.
Photo courtesy of the Hubble Space Telescope
(STScI/NASA).

Image from the Very Large Telescope taken
using Adaptive Optics. There are more stars
visible in this view and each star is clearly seen
- not blurry - since the telescope is not only
larger than the Hubble, but used adaptive
optics. Photo courtesy of the European Southern
Observatory.
104 | P a g e




Now that you've read this section, you should be able to answer these
questions.....
What are the different types of light, how are they ordered according to
wavelength, frequency and energy?
What conditions are required to cause the different types of spectra?
What are the rules that govern the behavior of black bodies and what do these
rules depend upon?
What influence does velocity have on light?
What is the difference between reflecting and refracting telescopes?
Where are the major observatories found on the Earth?
What telescopes must be located above the atmosphere, and why?
What are the major space based observatories in operation?
What is adaptive optics?
http://www.uni.edu/morgans/astro/course/Notes/section1/new4.html
The Sun

Astronomical Symbol for the Sun

What's covered here:
What are the observable layers of the Sun?
What are sunspots?
What other activities are associated with Sun's atmosphere?

In this part of the course we'll be looking at stars, and to start things off, we'll look at
the star that is near and dear to our hearts, the Sun. A lot of people forget that the Sun
is actually a star. It is just so dang bright when compared to the stars that you can see
with your eyes at night. The Sun isn't a really powerful star, or a very massive or hot
star, it is just a very nearby star. It is right next door, so to speak, so we can use its
proximity to learn a great deal about it and also other stars.
105 | P a g e


Figure 1. Images of
the Sun. The one on
the left is taken with
a filter that allows
images to be
obtained of the
surface features
safely - without
burning out the
camera or the
telescope. The
image on the right
is a false-color
composite image
taken with different
wavelength
sensitive cameras
on
the SDO spacecraft.
Both images show
the Sun, but not just
the surface layer.
Image courtesys of
SOHO/MDI
consortium. SOHO
is a project of
international
cooperation
between ESA and
NASA, and
NASA/SDO and the
AIA, EVE, and HMI
science teams.
Even though astronomers have been studying the Sun for centuries, there are still
some mysteries that we are perplexed by. One of the problems with studying the Sun
is that you can only clearly observe what is on the surface, not what is inside. Is that a
problem? Imagine how difficult it is to learn about things like frogs, plants or people
without taking a peak at what is inside of them. Actually, this used to be the rule about
humans - physicians were not allowed to study their internal organs until only a few
centuries ago. That is the situation with the Sun; we can learn all sorts of stuff about
the outside, but when we want to see what is going on inside we are really limited in
our methods - though we have some ways to work around these limitations, as you'll
see.

We'll start first by looking at the main surface features and later, when we get to
106 | P a g e

talking about the structures of stars, we'll delve more into the interior of the Sun
(though we won't get burned). When we observe the Sun, we see the atmosphere,
which is made up of several levels, which are, arranged going from the surface
outward, the photosphere, the chromosphere and the corona (a favorite amongst
students). We'll take a peak at each level and don't forget to wear your shades.
Photosphere
The main layer of the Sun that we see is the photosphere. Even though the Sun is
pretty much a big gas ball and has no solid surface, the photosphere is defined as the
surface, mainly because you can't see what's underneath it. A typical value for the
photosphere's temperature is around 6000 K, though some astronomers give
temperature values that are slightly higher or lower than this - but this is an easier
number to remember. We'll use this as the surface temperature of the Sun (and if you
are not a metric person, or just don't like the Kelvin temperature scale, then you might
want to know that the equivalent temperature is around 10,000 degrees F. Either way,
it is pretty darn hot). When you look at the Sun for longer periods of time (WHICH IS
A REALLY STUPID THING TO DO UNLESS YOU ARE A TRAINED, WELL
EQUIPPED ASTRONOMER) you may notice that the surface is not smooth. These
irregular surface features are sort of like bubbles. By viewing the photosphere with
special telescopes, this bubbled surface can be easily seen. These bubbles are actually
the tops of rather large bubbles - bubbles of gas that are bringing light (energy) from
the interior of the Sun to its surface. Of course, we can't call them bubbles; that
doesn't sound very scientific. The bubbles give the Sun a rather grainy appearance, so
the effect is known as granulation and individual bubbles are granules. What you are
actually seeing is a process known as convection, which is like a boiling pot of water
where heat is brought from below the pot to the surface (the steam you see). Like a
pot of water, the bubbles are continually changing. Here is a little movie showing how
the granules change over time - this movie actually covers about 35 minutes of
observations of the Sun. The granules come and go, and in the process they release the
energy (the light) from the interior, cool off and sink back down. That is why they
have a rather patchy appearance - the cooler areas (the areas that are going
downwards) are slightly darker than the areas that are coming up. Typically, the
cooler parts are about 300 K cooler than the surroundings. The average size of a
granule is 1000 km in diameter, though there are instances where they can form much
larger granules. The granules are also useful in providing us with information about
the inside of the Sun, since the granule sizes and the way they change can give us
some information about what is happening deeper inside
the Sun.
Figure 2. Granules on the surface of the Sun. Each bubble
is around 1000 km in size. The differences in the shades is
107 | P a g e

due to temperature differences, with lighter areas being hotter. Image courtesy
of NASA's Solar Physics Research Site.
The photosphere is pretty much the "surface" of the Sun, so it is the layer from which
the light that is produced inside the Sun is released. When we look at the light from
the Sun we see that it produces an absorption spectrum. The Sun's surface is thick
enough to look solid, so why doesn't it produce a continuous spectrum? Why are there
absorption lines?
Figure 3. The Sun's visible light spectrum
showing absorption lines. The entire
spectrum extends over a wide range of
wavelengths, which are seen here as the
different spectral colors (remember ROY G.
BIV?) and are chopped up so that it doesn't
stretch off too far. You can click on the
image to see a larger version. Image
courtesy of Nigel Sharp, NOAO/NSO/Kitt
Peak FTS/AURA/NSF
You have to remember how an absorption spectrum is produced. You need a cooler
layer of gas between us and the light source (where the light comes out) to absorb
certain wavelengths. The photosphere is pretty hot, isn't it? How can there be a cooler
layer of gas? As in many things in astronomy, it's all relative - the layer that is cooler
just has to be slightly cooler than the layer from which the light is emitted. In this
case, the cooler layer has a temperature of about 4800 K, which is about as cool as
you'll get in the photosphere. This is still hot by your standards, but in comparison to
the other, deeper layers in the photosphere, this is cool enough to do the trick - absorb
light and produce an absorption spectrum.

If you were to also look at the spectra given off by other stars, you would see that
about every star (unless there is something seriously wrong with it) has absorption
lines in its spectrum. It is likely that they also have a slightly lower temperature layer
in their atmospheres, just like the Sun. We'll get back to the photosphere in a bit, but
first we should finish off the other layers of the Sun's atmosphere.
Chromosphere
The next layer up is the Chromosphere. Even though this layer is further from the
surface of the Sun, the temperature of the chromosphere is greater than that of the
photosphere, typically about 10,000 K. The easiest way to see the chromosphere is
during a solar eclipse when the brighter photosphere is blocked out. At that time, the
108 | P a g e

chromosphere will appear as a multicolored or reddish pinkish layer during the
eclipse. If you don't want to wait for an eclipse, you can see the chromosphere using a
specially filtered telescope - the most common view today is with the SOHO satellite
that has several ultraviolet filters that are tuned specifically to view the chromosphere.
The gas is thinner in the chromosphere and harder to detect with most Earth based
telescopes.
Figure 4. Chromosphere of the Sun visible
with the SOHO satellite. Notice how the
chromosphere has a rather wispy appearance
and is not uniform in shape. Bright areas
indicate the location of solar activity. Image
courtesy of SOHO/EIT consortium. SOHO is
a project of international cooperation
between ESA and NASA.
Corona
The outer most layer of the Sun's atmosphere
is the Corona. While this is a rather large region in terms of how far it extends, it is
rather thin and diluted (of very low density). It is also very hot, so hot that most
electrons are striped off of elements (the gas is highly ionized). Temperatures in the
corona range between 1-2 million K. A layer with a temperature this high is a strong
source of x-rays. The corona is also the source of the solar winds, steady streams of
particles that are blown off from the Sun. These particles can on occasion interact
with the Earth's atmosphere and produce the aurora, or northern lights. During a large
expulsion of particles (as you'll see below) the northern lights can become very
spectacular.

Figure 6. Aurora visible due to the particles
109 | P a g e


Figure 5. The Corona of the Sun visible
during a solar eclipse. Image courtesy of
NOAO/AURA/NSF
ejected from the Sun. Image courtesy of Jan Curtis
Here's a question - why are the outer layers of the Sun hotter than the photosphere?
The atmosphere thins out (gets lower in density) as you get further from the
photosphere, so shouldn't that make the temperature go down? This is one of those
mysteries about the Sun that was only recently solved. There are mechanisms that
pump energy into the upper layers of the Sun, particularly through the interaction of
the Sun's magnetic field with these upper layers. We'll get into the magnetic field in
just a bit, but you might want to think of the heating up process as kind of an oil
gusher of energy from below the surface of the Sun extending all the way to the
Corona. The end result is very high temperatures far from the surface of the Sun. The
photosphere is the coolest layer of the atmosphere and the corona is the hottest - rather
strange, but that's the set up.
Figure 7. A graphical
representation of the
temperature in the various
layers of the Sun's atmosphere.
Note how the temperature
steadily increases throughout
the chromosphere and peaks in
the corona. You can also see
where the temperature is at its
lowest point, near 4800 K in the
photosphere.
Sunspots and related acne
Various phenomena observed
on the Sun can extend through
all of the layers of the
atmosphere, though it may be
more apparent in some layers
more so than others. Generally, most types of solar activity start in the photosphere
and then extend all the way up through to the corona and beyond. Of course, some of
this activity can have an influence on the Earth.
The most common type of solar activity is found easily in the photosphere, and these
are the things known as sunspots. Sunspots usually come in pairs or groups (they're
110 | P a g e

pretty friendly in that respect). Sunspots actually are rather complex in their structure,
usually having two main visible regions; the darker interior is known as the umbra,
while the not so dark outer region is known as the penumbra. Yes, these are the same
words used in association with eclipses - okay, I'll admit it, astronomers aren't too
original in naming things. Any ways, the penumbra has a rather streaky look to it, and
may not be as easy to spot (or is that sun'spot'?) as the umbra. Why are they dark? It
has to do with temperature. Sunspots are cooler when compared to the surrounding
surface of the Sun, with temperatures of around 4200 K being typical. For a high
resolution image of a sunspot, click here.
Sunspots are located on the surface (photosphere) of the Sun and can be used to
determine the rate at which the Sun rotates. If you were to follow the motion of the
sunspots, you would notice that the equator spins around faster than the poles. This is
not surprising since the Sun is a big ball of gas, after all, and different parts can move
around at different rates. Such motion is known as differential rotation, since
different parts rotate around at different rates. It takes about one month for the Sun to
rotate around, with the equator taking slightly less time and the polar regions taking
slightly more time.

Figure 8. A large sunspot group is shown, with the main
components, umbra and penumbra, indicated. Image
courtesy of SOHO/MDI consortium. SOHO is a project
of international cooperation between ESA and NASA.

Figure 9. The different latitudes of the
Sun rotate around at different rates, so
that the equatorial regions get around
before the polar regions (since the
equator rotates faster). This is the
concept of differential rotation. Spots
that initially started all lined up will not
stay that way very long.
People wanted to learn as much as possible about sunspots, so they did everything
they could to determine what causes them, why they have different sizes and such.
Typically sunspots will last for days, weeks, or even months depending upon the size.
The bigger ones last longer. One thing that was determined early on was that every 11
years there would be a maximum number of sunspots. This is a pretty regular trend
111 | P a g e

and can be seen in graphs that go back for hundreds of years (since people started to
count sunspots after Galileo described them in his observations). A graph showing the
average number of spots is shown in Figure 10. Some people think that there is a
relation between sunspots and the weather on the Earth, and this is possible. When the
sun has a large number of sunspots there are all sorts of other things going on as well
(I'll get to that in a minute). The end result is that there is quite a bit of extra energy
given off by the Sun when there are a large number of sunspots (this time is
called solar maximum). Of course, that energy can come to the Earth and heat up our
atmosphere. You can't blame sunspots alone for all the bad weather on the Earth,
since our weather is also influenced by many other things, like pollution, volcanic
eruptions and
such.
Figure 10. The
number of
sunspots varies
with a peak
occurring every
11 years. This
graph is based
on data from
the National
Geophysical
Data Center of
NOAA.


That's one
thing we know
about sunspots
- they have a maximum number popping up every 11 or so years. What else do we
know? If you were to obtain a spectrum of a sunspot, you'd see pretty much the same
spectrum as you see for the rest of the Sun, but the individual spectral features would
be split apart. Instead of seeing a single absorption feature at a certain wavelength,
multiple absorption features would be seen. This is what is known as the Zeeman
effect and it is caused by strong magnetic fields. This tells us that the areas of
sunspots are regions with strong magnetic fields. In fact, in large sunspots it is
possible to detect which magnetic polarity they have (north or south). Sunspots are
often in pairs or groups, so one part of the group tends to be a south magnetic pole,
and the other part is a north magnetic pole. That's pretty cool - but what else can we
112 | P a g e

learn? Take a peak at Figure 11. This shows a magnetic "image" of the Sun with
several sunspot groups. Very dark regions are associated with "north" magnetic poles,
while the lighter regions are "south" polarity
areas.
Figure 11. A magnetic plot of sunspots showing
the polarity alignment. Image courtesy of the
National Solar Observatory in Kitt Peak.


Do you notice how they are arranged in each
hemisphere? The spots in the Northern
hemisphere are lined up so that the spot with
northern polarity is on the left (or west) and the
one with southern polarity is on the right (to the
east). The spots in the southern hemisphere are
just the exact opposite, so that the southern
polarity spots are to the west and the northern polarity spots are to the east.
Throughout the entire sunspot cycle (which starts when there are the fewest number of
spots) this alignment will remain. For 11 years all spots that appear in the northern
hemisphere will be arranged with the polarity N-S, while in the southern hemisphere
the spots will be ordered S-N. Then when the next cycle starts (after the solar
minimum - when fewest spots are seen), the situation is reversed - the spots in the
northern hemisphere will be arranged so that their polarity is arranged S-N, while the
spots in the southern hemisphere will be arranged N-S. If someone were to ask you
how long it takes the Sun to go through a sunspot cycle, the answer wouldn't be 11
years, since after 11 years, the spot polarity isn't where it was at the beginning of the
cycle - an entire sunspot cycle takes 22 years - you have to wait for the magnetic
polarities to get back in order. Figure 12
shows the cycles.
Figure 12. The changing polarity of the
sunspots. During each cycle, the polarity of
the spots in each hemisphere stays fixed. When
the next cycle starts (11 years later), the
polarity is switched. It isn't until 22 years after
the first cycle that the Sun is back to where it started from.


113 | P a g e

Somehow the magnetic field of the Sun is involved in sunspots. Another thing that
people noticed with the sunspots is where they are seen on the surface of the Sun.
When there are very few spots (at sunspot minimum), the spots tend to be seen about
30 or more degrees north and south of the Sun's equator. As the years go by, they start
to appear closer to the equator. By the end of the 11 years since they started
appearing, they are located very close to the equator. When the spots start appearing
far from the equator again, we known that another sunspot cycle is starting. You can
check the spots' polarities to be sure - so this is another way of checking sunspot cycle
beginnings and endings. A cycle starts with the spots far from the equator and
concludes when they are close to the equator. You can have some overlap, but you
can distinguish which cycle the spots belong to not only by their location on the Sun,
but also their polarity. A map of sunspot locations over the years produces a diagram
known as the Maunder Butterfly Diagram. This is shown in Figure 13. You can see
that the current cycle started in about 1997 (when there were very few spots). In 2001
the cycle peaked - this would have been the time of solar maximum.
In July of 2006, astronomers at a solar observatory noticed that spots started appearing
on the Sun that were of opposite magnetic polarity than the other spots in that
hemisphere of the Sun, and these spots were also located far from the Sun's equator.
This indicated the start of a new sunspot cycle on the Sun - this happens when there
are very few sunspots, and as expected it happened about 5 years after the sunspot
maximum (peak number of spots). So over the coming years, look for more, and more
sunspots.
Figure 13. The
Maunder Butterfly
Diagram - the
locations of the
sunspots on the
surface of the sun
varies over the 11
year cycle. At the
start of a cycle, the
spots are far from the
Sun's equator while at
the end of the cycle
they are found close
to the equator. Image
based on data from NASA's Solar Physics Research Site.
What causes sunspots? We don't know for sure, but we suspect that the differential
rotation and convection going on under the photosphere can wrap up and tangle the
114 | P a g e

Sun's magnetic field. As a magnetic field line gets twisted and stretched out by the
Sun's wacky rotation, a part of it can erupt near the Sun's surface, producing a
sunspot. It is possible this happens when the magnetic field lines get really jumbled up
by the differential rotation. The polarities of the sunspots sort of point to this kind of
situation, since the magnetic flow continues from one spot to the other - they just sort
of pop the magnetic field outside of the Sun for a while. This situation will eventually
get worse and worse (more and more spots) until something happens that puts the
magnetic fields back in line (untwisted). The only problem is, the magnetic polarity
isn't the same as when the stretching originally started - but is reversed, which of
course causes the sunspots to have a different order of polarity from the previous
cycle. We're not exactly sure how the whole thing works since it is very difficult to
compute the behavior of magnetic fields. You might want to think of it as a big rubber
band that gets pulled, stretched and then eventually snaps.

Figure 14. The model for the production of sunspots. The rapid rotation of the sun and the fact
that it has differential rotation and bubbling convection cause the magnetic field lines to get
wound up and jumbled until they are so distorted that they buckle and break through the surface.
How the sun gets back to the original situation is still a mystery.
When there are many sunspots visible on the surface of the Sun (at the time of solar
maximum), other features such as solar flares and prominences are also visible.
These are eruptions from the surface, thought to be associated with sunspot activity,
but what actually causes them is not completely understood. These eruptions result in
the release of large amounts of energy and particles into space. Many of the particles
are electrically charged, since the outflow can contain electrons or protons. This flow
of particles can cause disruptions of power, communication and some really intense
aurorae on the Earth. With the SOHO satellite, astronomers have been able to detect
even more energetic eruptions known as Coronal Mass Ejections (CMEs). These
blast out excessive amounts of particles and energy at very high speeds (1,500,000
miles/hour). They are mainly associated with the upper layers of the Sun's
atmosphere, so they are not always easy to see - at least not with visible light
telescopes.
115 | P a g e


Figure 15. A large prominence seen on the
Sun. The Earth is shown to scale.

A rather twisted prominence. Both images
courtesy of SOHO/EIT consortium. SOHO is a
project of international cooperation between ESA
and NASA.
Observations by the SOHO satellite can alert folks on the Earth when one of these
CMEs goes off. We have an advantage, since the particles that cause most of the
problem take longer to get here than the light from the event. This is sort of like how
you see a firework display before you hear it - it takes time for stuff to get here from
the Sun - more than a day, while only a few minutes are needed for the light to get
here. This time delay is useful for folks who have to maintain communication
satellites, since the particles that come in can disrupt the electronics in satellites. The
Earth's magnetic field pretty much protects us from the worst of these blasts, and we
can be treated to a really nice display of aurorae. The Space Weather site can provide
warnings of such blasts and also predict possible auroral activities.


Coronal Mass Ejections
116 | P a g e

Figure 16. The movie above shows
multiple CME events. The main
part of the Sun is blocked out in the
center. The little white circle
indicates the actual size of the Sun.
The CMEs that are directed
towards the Earth cause static like
features on the images. The stars in
the background are also visible in
these images. To the right is a
typical large CME. Images and
movies courtesy of SOHO/LASCO
consortium. SOHO is a project of
international cooperation between
ESA and NASA
Less prominent but also visible are the spicules - which extend from the photosphere
up through the chromosphere. While the spicules are about 10,000 K, they extend up
into hotter layers near the corona (500,000 K) and can appear dark in various filtered
views of the Sun. Unlike other events on the Sun, the spicules are very short lived,
sort of like goose bumps that come and go on a cold day.
All the major activity events are linked with sunspots - prominences tend to erupt
from sunspot groups, CMEs come from sunspot groups, and flares come from the
areas of sunspot groups. I am noticing a trend here. When the sun is at solar
maximum, there is a greater chance for these things to occur and for there to be effects
on the Earth, particularly from the high amounts of charged particles (electrons and
protons) that come our way. This would also be the time to watch out for aurorae at
lower latitudes, like Iowa. While all of these events do occur at other times, the
likelihood of them happening is greater at solar maximum, and the scale of the events
are also greater at solar maximum.
Most of the discussion to this point has been about the atmospheric layers of the Sun.
While there is a connection between the atmosphere and the interior, it isn't very
obvious. How can we learn about the inside of the Sun by studying its exterior? You'll
learn a few tricks on how we do this later, but there is a way that we can study the
inside of the Sun in a manner similar to the way that geologists study the inside of the
Earth without drilling a hole. On the Earth, geologists study how earthquakes travel
through the Earth's interior and this provides clues to its internal structure. In a similar
manner astronomers can observe how the Sun "vibrates" - yes, that's right; it vibrates
sort of like a big bell. This is part of the relatively new field of study known
as helioseismology. As the name implies, this is like studying "sunquakes," since
seismology is the study of earthquakes. However, it is more complicated than that.
117 | P a g e

Unlike earthquakes, the Sun's vibrations are continuous - they don't stop. If you have
ever watched a bell, taunt string or wire make a noise, you see it vibrating quickly.
The Sun does a similar thing, except not as fast. It takes about five minutes for one
vibration of the Sun. The way something vibrates depends upon the object's structure;
astronomers study these vibrations in the Sun to get clues about the interior of the
Sun. There are actually several different vibrations occurring in the Sun all at the
same time, making it a very complicated thing to work out.


Figure 17. Above left is a model showing the vibrations of the Sun. One color corresponds to the
parts of the Sun that are going inwards, while the other color indicates the parts that are moving
outwards. To the right is an animation of how the surface of the Sun moves, though of course the
motions are greatly exaggerated and speeded up. Images from the GONG Project.

Now that you've read this section, you should be able to answer these questions....
What are the characteristics of each layer of the Sun's atmosphere?
What types of light do the different layers predominantly produce?
What can be learned from granulation?
What are the characteristics of sunspots and how do they change over time?
What causes sunspots?
What are the various outbursts from the Sun's surface and what do those
outbursts lead to?
How are the outbursts related to sunspots?
Stars

What's covered here:
What are the observed characteristics of stars?
118 | P a g e

How do we determine/measure brightness, distance, temperature, energy
output, radius, and mass for stars?
What can spectra tell us about stars?
How do we classify/categorize stars?
What can binary stars tell us?

While the Sun is a star, and most stars are similar in some way to the Sun, they are not
all like the Sun - they can be hotter or cooler, more massive or less massive, more
luminous or less luminous, and so on. This of course means that all stars are unique in
various ways. You do not even need to use a telescope or other instrument to see how
stars are different. What differences can you see with just your eyes?
One of the most obvious differences is that they have different brightnesses. Some
people equate this difference with size, but that isn't necessarily correct as you'll
see. Why do stars have different brightnesses? There are actually two things that can
influence how bright a star appears to your eye, the star's actual brightness and its
distance from you. By actual brightness, I mean how much power a star has - you can
think of it as wattage - just like light bulbs. Some stars have a higher wattage than
others. Various combinations of actual brightness and distance give you the range of
observed brightness that is seen in the night sky. Of course, if you were to just look at
any star and see that it was brighter than the stars around it, the exact cause of the
difference in brightness isn't obvious - is it a nearby star? Is it a really bright star?
How can you figure that out?
The other difference that you can see amongst stars is that they can have different
colors. Not everyone can see the color differences that well, but there are many stars
that have very obvious color differences - some look rather reddish, some look
yellowish, some look white-blue. Why do stars have different colors? The color
difference is due to a basic characteristic of a star, the temperature of the star's
surface.
Figure 1. The range of colors seen in stars is
shown here in this group of stars with a range
of brightnesses as well as colors. Image
courtesy of Don Figer (Space Telescope
Science Institute) and NASA.
Now remember, we can't touch stars or
sample them directly. Everything we learn
about them is by looking at the light that
comes from them and by applying certain
119 | P a g e

laws of physics. By observing the features that we can, such as the colors,
brightnesses, and spectra, it is possible to derive information on the masses, radii,
motions, distances, temperatures and chemical compositions of stars. However, as you
will see it is not always possible to find out everything about a star; sometimes very
little can be discerned.
Brightnesses - the Magnitude Scale
One of the easiest things to note about a star is how bright it looks. Astronomers like
to make catalogues of everything and keep track of all of the stars in the sky, so they
had to find a way to rank the brightnesses of the stars. Why did they need to do this? I
don't know, maybe they were bored, or maybe they had too much free time on their
hands; either way, they wanted to assign some sort of number to the brightness of the
stars. This was done in the ancient times generally by assigning a number to the star to
rank its brightness, sort of like saying that a certain star has a rank of 1 since it is
really bright, while one that isn't as bright gets a rank of 2, and an even fainter one is
ranked 3.This is sort of like putting them in order with the brightest as number 1 and
the faintest as number 6 or so. The only thing that was confusing about this was that
the smaller the value for its brightness (which we call the magnitude), the brighter the
star. A star with a smaller magnitude is brighter than one with a larger magnitude
value. That's silly, isn't it?
Astronomers assigned these numbers to a bunch of stars and eventually got the system
worked into a standardized format, so that everyone was using the same numbers for
the same stars. The scale was also calibrated a bit better but this resulted in some stars
being assigned negative magnitudes. Actually, this magnitude system is pretty
flexible. You can assign a magnitude value to not only stars, but also other things in
the sky, including planets, comets, asteroids, galaxies, the Moon and of course the
Sun.
The dimmest magnitude that you can see with your eye is about 6th magnitude. Big
telescopes can see objects as faint as maybe around 20th magnitude. The Hubble
Space telescope can see down to a magnitude of about 25 or fainter. What does that
mean? Is a star with a magnitude of 4 really "weaker" or not giving off as much light
as one with a magnitude of 2? No, this scale is just based upon how it looks to our
eyes.

The values in the magnitude scale are spaced out so that a difference of 5 magnitudes
corresponds to 100 times more energy being detected. If you observe a star with a
magnitude of 8 and another with a magnitude of 13, the star with a magnitude of 8 is
100 times brighter (more energetic) than the one of magnitude 13. You could also say
that we detect 100 times more light from a magnitude 8 star compared to a magnitude
120 | P a g e

13 star. What's the difference between a star of magnitude 5 and magnitude 6? The
magnitude scale is logarithmic, so it isn't a simple linear scale - the energy difference
for a 1 magnitude span is about 2.5, so we detect 2.5 times more energy from a star
with magnitude 5 than one of magnitude 6. We would detect about 2.5 x 2.5 = 6.25
times more energy for one of magnitude 5 compared to magnitude 7, and for each
increment of magnitude just multiply the energy amount detected by 2.5. For a
magnitude difference of 5, you would multiply 2.5 by itself 5 times and get a number
close to 100 - which is what I said originally. Basic upshot - the bigger the difference
in magnitude, the bigger the difference in energy detected.
The ancient astronomers defined magnitude as just a measure of how bright a star
appears to us in the sky, so the value of the magnitude depends on both how close the
star is and how much energy it is emitting. Obviously the Sun is at the top of the list
of all objects in the sky in terms of brightness - it has the most negative value for a
magnitude based upon how they appear in the sky. Here are some typical values for
magnitudes -
These magnitudes are referred to as the
object's Apparent Magnitude (and that's what
the m stands for). Again, it should be
remembered that this is how we rank the
brightness of the object as it is viewed from the
Earth - it isn't really meaningful if we want to
determine which star is really giving off the most
energy.
To see how bright a star really is (how much
energy it is giving off), it is necessary to remove
the distance differences between stars. Unfortunately, this is not an easy task.
Actually, it is easy, but we can't do it very well - but it really isn't our fault! Let me
explain how you can find the distance to a star.
The easiest method to use to get a star's distance is to measure a star's parallax. No,
that is not what you call two laxatives. What is parallax? If you remember the history
of astronomy stuff, parallax is the shifting location of nearby objects compared to
more distance objects when you change your viewpoint. If you still don't remember,
then you can see how parallax works by taking your thumb and holding it at arm's
length. Line it up with an object in the distance and view it with one eye. Now switch
your eyes (I don't mean take your eyes out of your head and switch them, I mean close
one eye and open the other - sheesh!). What did you see? Did it look like your thumb
moved? Didn't you hold your thumb steady? If you held your hand steady, your thumb
didn't really shift - but your perspective shifted; you viewed your thumb from a
Star Apparent Magnitude (m)
Sun -26.8
Arcturus -.06
Sirius -1.47
Vega 0.04
Betelgeuse 0.41
Polaris 2.0
121 | P a g e

different location. If your eyes were further apart, guess what would happen? First,
your head would probably hurt really badly, but as for what you would see, it would
make the shift even larger. You can get a bigger shift if you can view the nearby
object from very widely spaced viewpoints. However, there isn't really much that you
can do to get your eyes further apart - at least nothing that I would recommend doing.
What would happen if you were to bring your thumb closer? What happens to the
shift? You may notice that the amount of the shift has changed. The shift should get
larger as you decrease the distance to your thumb. If you were to increase the distance
between your eyes and your thumb, though that might be rather painful, you would
get - yup, you guessed it, an even smaller shift. Now you can see how the size of the
shift is related to the distance of
your thumb from your eyes.
Figure 2. Stellar parallax. As the
Earth goes around the Sun, the
position of the nearby star appears
to change relative to the more
distant background stars. The size
of the shift is related to the distance
of the star.
The same thing can be done with stars. For the best results you want to use the most
widely spread apart viewpoints possible. The biggest scale that we have available to
us is the orbit of the Earth. Basically, you observe a nearby star at two times during
the year, like in January and July. (See Figure 2.) When you compare the location of
the nearby star relative to the distant, background stars, you may note that the position
of the nearby star has shifted slightly relative to the background stars. The parallax
shift of stars can be related to the shift you saw with your thumb. In this case, the two
locations of the Earth correspond to your eyes (your two different views) and the
nearby star corresponds to your thumb. The parallax concept was known even to
ancient people, and it was one of the reasons they gave for the Earth to be in the
center of the solar system and to be stationary. They thought that if the Earth did have
an orbit, then they could see the parallax shifts of stars. They didn't. This is because
parallax angles are very, very, very small and can not be seen without a pretty good
sized telescope. Take a peak at Figure 3 to see the situation.
Figure 3. The size of the parallax
shift is related directly to the
distance of the object. The
parallax angle for the nearer
122 | P a g e

object, p1, is larger than the shift for the more distant object.
The size of the shift is denoted by p. It would be ridiculous to measure its size
of p with degrees or arc minutes, since p is usually so small. Instead it is always
measured in arc seconds (remember one arc second is 1/3600 of a degree, a very
small angle indeed). We know that the size of the parallax shift is related to the
distance of the star, so you can use the following relation
d = 1/p
to find the distance to the star. d (distance) will be in units of Parsecs, and p is of
course in arc seconds (small fractions of a degree, whose symbol is "). In case you are
wondering, a parsec is just a regular unit of measure that was based upon the parallax
concept - get it? For those of you who are fans of science fiction, 1 parsec
(abbreviated as pc) = 3.26 light-years, and yes, light-years are also legal units of
measurement, just like kilometers, inches and miles. Don't get confused by the name,
a light-year is a unit of distance, not time. A parsec is actually a fairly large distance,
about 3.09 x 10
13
km, which translates to about 20 trillion miles. A light-year is a
mere 5.9 trillion miles. Astronomers use parsecs and light-years pretty
interchangeably. Distances to stars are usually measured in the tens or hundreds or
thousands of parsecs or light-years. When we get to galaxies we'll be using millions
and billions of parsecs and light-years - but we have a long ways to go before we get
there.
The closest star (apart from the Sun) would have the largest parallax angle. This star is
alpha or Proxima Centauri, which has a p=3/4", giving it a distance of
d=1/(3/4) = 4/3 pc
If a star has a parallax angle of 0.01", then it has a distance of
d = 1/0.01=100 pc.
Remember, the angle gets smaller as the distance gets larger. The largest angle we can
measure for a star (that for Proxima Centauri) is so puny that it isn't even the size of
1". All other stars have even smaller angles (are at greater distances). 3/4" is not very
big; it is about how wide a pencil lead would look if you were to stand 1.5 football
field lengths away from it. That's pretty small. It is possible to measure the parallax
angle for only about 10,000 stars using even the best telescopes on Earth. You might
remember that the Earth's atmosphere has this annoying tendency to smear out the
light in the atmosphere, making it hard to get precise position measurements.
123 | P a g e

To help get past the problem of the atmosphere, a special satellite, named Hipparcos,
was launched with one main task, to measure the parallax shifts of over a million
stars. It finished its job in the late 1990s and these improved distances have really
helped astronomers figure out distances to other stars. The Hipparcos satellite was
able to measure very precise parallax angles, in some cases down to 0.001" accuracy.
There are future plans for satellites to do more angle measurements from space, with
in some cases the parallax angles of billions (yes, BILLIONS) of stars being
measured.
While the parallax method is the most direct method of measuring the distance to a
star, there are other methods which depend upon various special circumstances or
characteristics. We'll look at some of these later.
What has all of this distance determination gotten us? Remember, we were trying to
figure out how to find out which stars are really the brightest. Once the distance to a
star is known, it is possible to compensate for the distance and we can figure out what
the actual brightnesses of stars are. The best way to do this is to move all stars (not
actually move them, but account for their distances in some mathematical ways) to the
same distance and then compare their brightnesses. This is like having a group of
people all stand in a line so that you can tell which ones are taller or shorter. If we
were to move all stars to a distance of 10 pc from the Earth and then measure their
brightnesses, we could determine which stars were actually brighter and which ones
were actually fainter. The magnitude that a star would have if it were placed 10 pc
from the Earth is known as the Absolute Magnitude. To distinguish it from apparent
magnitude, we use M.
Here are some stars' apparent and absolute magnitudes. Of course, the Sun is a lot
closer than the other stars, so its apparent magnitude is quite a bit different from its
absolute magnitude. In the cases of the other stars, some have to be brought in closer,
since they are further than 10 pc away, while some stars have to be moved away since
they are closer than 10 parsecs. Which stars in the list below had to be moved closer
and which had to be
moved further away?
Star Apparent Magnitude (m) Absolute Magnitude (M)
Sun -26.8 4.83
Sirius -1.47 1.4
Arcturus -.06 -0.3
Betelgeuse 0.41 -5.6
Polaris 2.0 -4.6
124 | P a g e

The stars that had to be
moved away from the
Earth to place them at 10 pc are the Sun, Sirius and Vega. You know this because
their absolute magnitudes have a larger numeric value than their apparent magnitudes
- the stars became fainter. The other three stars had to be brought in closer, so their
absolute magnitudes have smaller values than their apparent magnitude values.
Remember, the apparent magnitude and the absolute magnitude scale is sort of
backwards - the larger the number, the fainter the star. Unlike the apparent magnitude,
the absolute magnitude of a star is a realistic measurement of its energy output. By
comparing absolute magnitudes, you are comparing the energy output differences
between stars. Which star in the list is producing the most energy? Is it the Sun? No -
that's actually the weakest star (it has the largest absolute magnitude value).
Betelgeuse is the most powerful (highest energy producing) star in this list since it has
the most negative value for its absolute magnitude.
The formula which relates the magnitudes and distances is a fairly straight forward
formula, which is
m-M = -5 + 5 log(d),
where the m and M values are the magnitudes, and d is the distance in parsecs. This is
a pretty handy formula for converting things. Some people get a bit confused since the
distance gets to go through the log function - don't worry about that - it's a pretty
simple function on most calculators. To un-log something, you just take it to the
power of 10. If you have a star with an apparent magnitude of 7 and an absolute
magnitude of -2, how far away is it? Just put the numbers into the formula
7- (-2) = -5 + 5 log(d)
9 = -5 +5 log(d)
9+5=5 log(d)
14/5 = log (d)
2.8 = log (d), so d=10
2.8
=630 parsecs. That wasn't so bad, was it?
That is one thing we can learn about stars - absolute magnitudes can tell us which
stars are producing more energy. What does this energy output depend on?
Let's go back to the rules for black bodies. Remember, these are the hot, solid objects
that produce continuous spectra. Now stars aren't really black bodies, but they are
pretty close, or at least close enough so we can use the rules for black bodies to make
our lives easier. One of these rules was how much energy a black body produces -
the Stefan-Boltzmann law -
E = T
4

Vega 0.04 0.5
125 | P a g e

where is a constant and the energy emitted is the number of Watts per square meter
given off by the black body. The energy output from this formula is given in terms of
the energy per unit surface area, so the size of the black body (or star) would influence
the over all energy output. A larger star (one with a large surface) would produce
more energy. To account for that we use the following formula:
Luminosity = L = 4 R
2
T
4

where (=3.14 or so) and are constants, T is the temperature of the star's surface,
and R is the radius of the star. This defines the luminosity, the way that we talk about
the energy outputs or brightnesses of stars. Unlike the magnitude system, the larger
the number for the luminosity, the brighter the object is. Luminosity is measured in
units of Watts, just like light bulbs, and it is one of the more important characteristics
for stars that astronomers like to determine; I guess we're just silly that way. There are
ways to translate absolute magnitudes into luminosity values, so at times you may see
either one of these scales used in comparisons of stellar brightness. Either way you
look at it, a star's brightness may be given in terms of its luminosity or absolute
magnitude.
To make things simpler, astronomers often talk about the luminosity or temperature or
radius of a star in terms of the Sun - if you do that, then you can use a simplified
version of the above formula.
L = R
2
T
4

If a star has a temperature that is two times that of the Sun's, then it's luminosity
would be 2x2x2x2=16 times greater. If a star has a temperature that is two times
greater than the Sun and it is 1/3 the radius of the Sun, then its luminosity would be
(1/3)
2
2
4
= 16/9 = 1.78 that of the Sun. This is much easier to do than dealing with all
of the messy symbols that are in the original formula. Once you have the value of
luminosity and temperature for a star, this formula can be used to determine the
radius, or if you have the radius and temperature you can get the luminosity, or.... I
think you get it, if you have two of the three things in the formula you can figure out
the missing value.
The luminosity depends upon the temperature of a star, so it would be a good idea to
be see how astronomers are able to determine a star's temperature. One way to
determine the temperature is to use Wien's Law (
max
= 0.0029/T) to determine the
temperature of a star, provided you can actually observe
max
- the wavelength at
which most of the light is emitted. This is not possible for the hottest or the coolest
objects, since they would produce most of their light at wavelengths beyond the
visible part of the spectrum, and unless you have access to a special telescope, you're
126 | P a g e

out of luck. You would need other types of telescopes to study them, such as X-ray,
UV, IR or radio. However, most stars do have their peaks in the visible part of the
spectrum, so we can determine the temperatures of most stars by using Wien's law.
Another method, that is a bit easier than Wien's law is known as photometry. While
this sounds like a complicated method, it is really quite simple. Remember how we
went over the differences of stars, how their colors are different because their
temperatures are different? What if you could measure the color of the star? That is
sort of what photometry is. Astronomers use special filters on their telescopes to note
how the brightness of the star changes when viewed with different color filters. A
really cool star would be very bright when viewed through a red filter but not very
bright when viewed through a green filter. A hotter star would have more light
coming through the green filter, and an even hotter star would be giving off more light
in a blue filter compared to a red filter. By measuring the relative differences of the
light observed through the various filters (by measuring their magnitudes),
astronomers can assign a temperature value to the star. This is the most common
method of determining a star's temperature.
What have we been able to figure out so far? - the distances, the luminosities and
the temperatures of stars. What else can we learn about stars? How about a star's
spectra? To see what can be learned from a star's spectrum we'll learn about a rather
nifty bit of stellar research history. Early in the 20th century, astronomers at the
Harvard College Observatory started to catalog various spectra. How do you catalog
spectra? First you need to obtain spectra, which is pretty easy and was done using
photographic techniques and a prism-like device (spectroscope) to spread out star
light into spectra. Stars are so far away, and even with the biggest telescope they look
like dots, so their spectra end up being really tiny - so tiny, in fact, that you have to
view them with a microscope. Nowadays with computers this eye-strain is avoided,
but in 1910 this was the way it was done. Who was going to spend hours peering into
a microscope at tiny little spectral features? The astronomers weren't going to do it
since they had more important things to do. The students weren't going to do it; after
all, they were paying to go to college. The folks at Harvard needed some workers who
would be willing to work for pretty low wages, do very meticulous work and work
with very delicate objects (since the spectra and other astronomical pictures were
often produced on glass plates). Obviously, the best pool of laborers would be women.
Those women worked like the dickens! There was one lady, Annie Jump Cannon who
classified more than 250,000 stars herself.
Figure 4. Typical stellar spectra - note that
these are all absorption spectra. Also notice
that the pattern of the spectral features are
different from one star to another. You can
127 | P a g e

click on the image to see a larger version of it. Image courtesy of NOAO/AURA/NSF.
You've got a bunch of spectra to classify; how do you go about doing that? The
obvious thing is to look for trends in the spectra. One trend was that the strength of the
spectral features associated with hydrogen appeared to be very prominent in some
stars and not so prominent in other stars. Hydrogen is a pretty important element, so
let's call those stars with really prominent hydrogen spectral features 'A' type stars.
Those with slightly weaker hydrogen spectral lines are 'B' type stars. Then you could
also classify stars that have other elements with other letters of the alphabet. 'C', 'D',
and a whole alphabet soup of star types (spectral types) seemed to pop up.
After awhile, they determined that the classification system was not an accurate
portrayal of the physical characteristics of the stars. 'A' type stars didn't really have
more hydrogen than other types of stars, but the conditions in their atmospheres made
it a very strong feature in their spectra. At the time astronomers thought that the stars
were made of the same stuff as the Earth was - lots of rock and such. Someone wrote
up their Doctoral thesis and proposed that stars were actually made up mainly of
hydrogen and helium. This was Cecilia Payne, the first person to ever get a Ph.D.
from Harvard observatory. At the time, most astronomers didn't think she was correct,
but eventually it was shown that she was correct about the nature of stars. Stars are
made up of pretty much the same stuff (mainly hydrogen and helium), so the variation
in the spectra has to be due to something else. It was determined that the primary
cause of the variations in the spectra is the temperature of the star's surface. Now the
astronomers had a bunch of stars classified by an alphabetical system that wasn't
really in a logical or useful order. It would be better to arrange the stars in order of
temperature, since that is what makes the various spectra unique. They had already
labeled most of the stars, and didn't want to relabel them, so they just rearranged the
sequence of star types into an order based upon temperature. There were also some
redundant star types that needed to be removed and eventually the way that the
spectral classification were ordered was put into a logical format.
Once everything was rearranged, the spectral classification system was defined. The
order of the letters that are used to classify different stars were initially arranged as
follows - OBAFGKM (ordered from high to low temperature). There is also an old
alternative sequence that uses some rather obscure spectral types, and this
is OBAFGKMRNS. Again, the stars are arranged from hottest to coolest surface
temperature (though R, N and S types are sort of the same as K and M types). These
orders of letters are not the easiest things to remember, so a memory aid would be
good to have for these things. It was the male astronomers (probably very lonely male
astronomers) who devised a way to remember the order of the stars by using the little
saying Oh Be A Fine Girl, Kiss Me. Of course, now you can say Oh Be A Fine Guy,
128 | P a g e

Kiss Me or, with the longer saying,Oh Be A Fine Girl/Guy Kiss Me Right Now
Sweetie. Figure 5 shows various spectra with their types labeled.
Figure 5. Typical examples of
the different spectral types.
The hottest stars, O-types, are
near the top, while the
coolest, M-types, are at the
bottom. There are also some
unusual stars included. The
names of the stars are
indicated on the right - most
are just their catalog
designations. Note how the spectral line patterns change with changing temperature.
Click on the image to see a larger version. Image courtesy ofNOAO/AURA/NSF.
The spectral classification system has been pretty steady for more than 70 years.
However, there have been some studies that have come up with some stars that are
even cooler than M types. These have been labeled the L and T type stars. L and T
types are usually not included in comparisons to the other types since these stars are
so cool and faint - they are primarily visible at only infrared wavelengths. Right now,
the complete spectral classification system is OBAFGKMLT. This is not the easiest
sequence to remember, but it is an important one nonetheless. Remember, even
though this is based upon spectra, it is a temperature sequence.

Even with the introduction of the L- and T-type stars, there are still only nine classes
of stars in this system. This is not a very diverse classification system, so to be more
precise, it is possible to subdivide each of these types into further groups. There are
usually 10 subdivisions for each type, so there are A0, A1, A2,... up to A8, A9, F0,
F1... F8, F9, G0 and so on. In some cases there are even fractional types - like the
O6.5 shown in Figure 5. Again, the order is from hottest to coolest. You may have
heard of the Sun referred to as a G2 type star. Once you've figured out the spectral
type of a star, you know what its temperature is.
While it is true that stars all have pretty much the same chemical make-up (mainly
hydrogen and helium), there are some subtle differences in their compositions. Stellar
spectra can help astronomers find these differences, usually by looking at things like
the amount of iron or other heavy elements in the spectra. Generally, stars have a
composition of about 97%-99.999% Hydrogen and Helium combined, with the
remaining fraction comprised of all of the other elements. You might not think that
this is a very diverse range in composition, but the fact that the metal content (non-
hydrogen and helium part) ranges in value from 3% to 0.001% does show that not all
129 | P a g e

stars are made out of exactly the same stuff. As you'll see, there is a very good reason
for this.
Bringing it all together - the H-R Diagram
Where are we now? We have methods of getting distances and energy outputs
(luminosities), ways of getting the temperatures and ways of checking the chemical
compositions of stars. Of course, once astronomers have a large amount of data, they
want to combine it together to see if there are any relations between the measured
quantities.
Once enough information about stellar temperatures (from photometry or spectral
classification) and luminosity (from absolute magnitudes) was known, some people
started to see relations between these quantities. Two fellows took the information on
the stars' temperatures, often in the form of a spectral type or color, and the
luminosities, often in the form of an absolute magnitude, and made up a diagram
relating these two quantities. These two fellows did the same thing at the same time,
quite independently, on different sides of the Atlantic Ocean, so the diagram is named
after both of them - the Hertzsprung-Russell Diagram, or, more simply, theH-R
Diagram. Figure 6 shows a typical H-R diagram. One thing that is confusing about
the H-R diagram is that the temperature scale increases towards the left. This isn't
normally how you would graph things, but since they often used the spectral
classification system to set up the temperature scale, and that goes from hot to cool,
you get a 'backwards' temperature scale. This is just another one of those annoying
things that astronomers like to
do to confuse poor
undergraduates.
Figure 6 Typical H-R
diagram. Notice that the
vertical axis can be scaled by
either the luminosity or the
absolute magnitude values.
Also note that the temperature
scale increases towards the
left. The main types of stars
are also included.
It was noticed that stars were
not scattered randomly about
the diagram but were found in
various distinct groups. Each
130 | P a g e

group has its own characteristics, and it is possible to use the Luminosity-Radius-
Temperature relation to expand upon these characteristics.
Main Sequence (M.S.) Stars - The diagonal through the middle is big since most
stars are of this type, about 90% of all stars, in fact. The Main Sequence stretches
from the low luminosity, low temperature stars in the lower right to the high
temperature, high luminosity stars in the upper left. A very wide range of
characteristics are found amongst stars on the Main Sequence, as you'll see. This is
also where you would find the Sun.
Giant Stars - These tend to be more luminous than stars on the Main Sequence and
often have lower temperatures than stars of comparable luminosity on the Main
Sequence. For them to have lower temperatures but not significantly lower
luminosities they must have a really big radius. How big are they? They are big
enough so they are called Giants! They are often at sort of low temperatures, so they
are usually named Red Giants. In the old days, people sometimes referred to the stars
on the Main Sequence as "dwarfs" since they were so much smaller in radius than the
Giants. However, this is not a very common name anymore, though you may still see
it pop up on occasion.
Supergiant Stars - These are just really big stars. They come in both hot and cool
varieties - Blue and Red Supergiants - and they are just really, really luminous, so you
find them hanging out in the upper part of the H-R diagram. They can also have pretty
high radii values as well. They are just BIG!
White Dwarf Stars - These are stars found in the lower left corner of the graph. They
are generally on the left side, so this means that they are pretty hot. They are also very
faint. To get low luminosities with high temperatures, they must have very small radii.
That's why they are called white dwarfs - hot and puny.
You may have noticed that stars can have the same spectral type (temperatures) but
may have vastly different luminosities - often one star's luminosity is thousands of
times greater or less than another with the same temperature. This difference in the
luminosities does have a subtle influence on the spectra so that astronomers can use it
to classify stars in another way (oh goodie, another classification scheme!). This
classification is known as the Luminosity Class. The various classes are shown in
Figure 7. Stars on the Main Sequence are type V, while various giants are types IV,
III, and II, and Supergiants are type Ia or Ib. You can describe the Sun as being a G2V
star. What about the White dwarfs? Generally they just get a "D" attached to their
spectral type - like DA3. A combination of the spectral type and the luminosity
classes allows you to determine where any star is on the H-R diagram.
131 | P a g e

Figure 7. The various
luminosity classes are shown.
Main Sequence stars are
denoted with a V, those
slightly above the Main
Sequence are IV, Giants are
III, Bright giants are II, and
Supergiants are either Ia or
Ib. White dwarfs are just
denoted with a "D".
Once the H-R diagram was
popularized, a new method of
determining the distances to
stars was found - that
of spectroscopic
parallax. Actually, this is a
rather confusing term, since there is no parallax angle measured. What it should be
called is spectroscopic distance, since it is done by taking the spectrum of a star and
determining where the star belongs on the H-R diagram. The H-R diagram can be set
up using absolute magnitude on the vertical axis; when you classify the star according
to its spectral type and luminosity class you can read off the value of absolute
magnitude. That can be compared to the apparent magnitude to get the distance. Let's
say you have a star which you obtain the spectrum for. When you compare to other
stars, you see that the star's spectrum is a K1 IV type star. A K1 star has a temperature
of around 5000 K. Looking at Figure 7, such a star would have a corresponding
absolute magnitude of around 1. This can be compared to the star's apparent
magnitude and the distance can be determined.
Make sure you understand H-R diagrams pretty thoroughly, since you'll be seeing a
lot of them for this part of the course. They are the main tools used to show how stars
relate to one another and they help astronomers to map out groups of stars for
comparison.
Binary Star Systems
So far we have been able to determine distances, luminosities, temperatures and radii
of stars. What about their masses? That is where binary starsystems come into play.
Most stars are in some sort of group, with the most common grouping being a binary
system (two stars). Stars in a binary system are orbiting one another, so they must
obey the rules that govern how objects orbit - Kepler's Laws! You thought those laws
132 | P a g e

only applied to planets! Actually, you need to use special forms of Kepler's laws when
you apply them to stars, but they are really just the same laws.
Remember, Kepler's law (the 3rd law specifically) has in it terms for the distance
between the objects and the period of the orbit. Usually when astronomers view
binary star systems they can determine how long it takes to orbit, while the distance
between the stars is a little bit more difficult to figure out. Generally there are some
estimates that can be made about the separation of the stars, and then these estimates
are carried over to the masses, so that generally we only have a good estimate of the
mass. We usually don't have incredibly precise values for the masses, just good
estimates.
Kepler third law when applied to stars is actually still pretty simple -
M
1
+ M
2
= a
3
/P
2

where M
1
and M
2
are the masses of the two stars (in solar masses), and a is the
average distance apart (measured in A.U.) and P is the period of the orbit (measured
in years). If you wanted to use this law to determine the masses of the individual stars,
then this won't get it for you. This will only get you the sum of the masses, not their
individual masses. You need another formula to get the masses. This formula is called
the Center of Massformula, or some such silly thing, but I like to call it the see-saw
formula. Here it is
M
1
a
1
= M
2
a
2

where a
1
and a
2
are the average distances each star is from the center of the orbit (See
Figure 8) and I should also mention that a
1
+ a
2
= a. As with the modified version of
Kepler's third law given above, the masses are in solar masses and the distances are in
A.U.s.
Figure 8. Binary star system - the orbits of two stars is shown;
both go around the center of mass. The distance each is from
the center of mass changes as they orbit, so the average
distance each is from the center of mass is used in the
formulas.
What's the center of the mass? It may be easier to think of it as the center of the orbit.
Think about it - if you have one object orbiting a fairly stationary object, it's pretty
easy to figure out the orbit size and where the center of the orbit is. In a binary star
system, both objects are moving - so how can you figure out the orbit size if neither
object is standing still? We have to define a location for the center of the orbit. This
location is known as the center of mass. Just think of a binary system as a see-saw or
teeter-totter. If you have two people on a see-saw, and one weighs much more than
the other, then how should the people sit so that they balance? Have the big person sit
closer to the pivot point. Figure 9 shows the arrangement. This is just like what
133 | P a g e

happens in the binary star system. The bigger star is closer to the pivot point - the
center of the orbit (or center of mass). Actually, if you look at the formula for the
center of mass you can rearrange it so that the masses are on one side of the formula
and the distances are on the other. In this case you would end up with
M
1
/M
2
= a
2
/ a
1

which is pretty nifty. If you have a mass that is five times greater than another mass,
then that mass has to be five times closer to the center of mass (its a value has to be
five times smaller). The ratio of masses is inversely proportional to the ratio of
distances.
Sometimes the big star is so close to the pivot point that the pivot is actually enclosed
within the star. It looks like the big star doesn't move - it really does, but not enough
to be obvious. This also applies to the Sun and the planets. The Sun actually does
move a little bit, mainly due to the influence of Jupiter. Anyways, with the two
formulas, it is possible to solve for the individual values of mass, though it will
require some algebra.
Figure 9. The center of mass depends upon the masses of the
object involved. The center of mass is closer to the object with
a larger mass. The way that the masses relate to one another is
the inverse of the way that the distances to the center of mass
are related to one another.
There are two main types of binary star systems. The first is the Optical Binary. This
is not a "true" binary system. The stars just appear to be next to one another in the sky,
but are in reality very far apart. If you have ever looked at the two stars in the handle
of the Big Dipper, Alcor and Mizar, they look like they are close to one another. This
is actually an optical binary system, since these stars are really very far apart from one
another. They are just lined up in a way that makes it look like they are next to each
other. This type of binary system is pretty useless for getting information about the
masses since the stars are not in orbit about one another. The truly useful binary
systems are the Physical Binary Systems. This is where you have two stars orbiting
about one another so that you can apply the modified versions of Kepler's Laws, and
they can be used to determine the masses of the stars. To
make things just a little more complicated, there is more
than one type of Physical Binary system (you knew it
wouldn't be so easy, didn't you?).
Figure 10. Optical Binary - the stars are not anywhere
near one another, but because of their alignment in the
sky appear to be close to one another.
134 | P a g e

There are 3 types of Physical Binary Systems. The first is the most obvious - the case
where you see two stars actually moving about one another. This is known as a Visual
Binary.Actually, this is a fairly rare type of binary system, since you have to be able
to see the motion and only very nearby stars will show motion in a binary system.
You can see their orbits, so you can estimate the sizes of the orbits as well as the
periods. Once you have done that you can apply Kepler's Third Law and the center of
mass law to determine the masses of the two stars.
Figure 11. A visual binary system is shown. The stars are in
orbit about the center of mass of the system. The sizes of their
orbits depend upon their masses - the more massive, the closer
to the center of mass.
Binary stars move and motion can often be detected in the spectrum of a star (via
the Doppler effect), so the next type of Physical Binary system is theSpectroscopic
Binary. In this case you may not actually see two stars, but the spectrum reveals the
presence of two stars orbiting about one another. In this instance, you would see
Doppler shifts due to orbital motions - one star moving towards you and the other
moving away. Again, by following the motions you can derive the orbital periods and
the sizes of the orbits, and using Kepler's laws you can get masses. It is also possible
that one of the stars is so faint that you don't even see its spectral features, but you
only see one set of spectral lines going from redshifted to blueshifted and back again
as it orbits the other. Even with this little bit of data it is still possible to get some idea
of the masses involved.
Figure 12. A spectroscopic binary is shown. The motion
is seen in the spectrum by how it is affected by velocity.
The spectrum of the star coming towards us is blue
shifted (seen at shorter than normal wavelengths), while
the spectrum of the star moving away from us is red
shifted (longer wavelengths). The stars switch direction
as they orbit about the center of mass, so the spectral
features associated with each star also switch from being
red to blue shifted and from being blue to red shifted. In reality you would see the two
spectra combined into one so the lines would go back and forth across one another.
The third type of Physical Binary system is the most useful. This is the Eclipsing
Binary system. In this case you see a light variation as the stars pass in front of one
another and/or behind one another. The overall brightness of the star system changes
over time in a repeated, periodic manner. By following the change, you can determine
the period of the orbit and the size of the orbit, and apply Kepler's laws to get the
masses. This binary system has an added bonus! The duration of an eclipse will
135 | P a g e

depend upon how wide the stars are, so it is possible to also determine the radii of the
stars in these binary systems. This type of binary is particularly useful.
Figure 13. An eclipsing binary system. The variation of
the brightness (the graph at the bottom) due to eclipsing
stars depends upon the brightness of the individual stars.
When the stars are both visible, the brightness is at a
maximum value. During the eclipses, the brightness goes
down when something is being covered up. The sizes of the
stars (radii) can be seen in the widths of the eclipses.
With the masses obtained from binary star systems, it is possible to compare the
masses (M) and luminosities (L) of individual stars. When we look at these
characteristics for stars on the Main Sequence, we note that there is a very good
relationship between M and L. The relation is pretty simple
L= M
3.5

where the mass and luminosity are given in terms of the Sun. Remember, this formula
works well only for stars on the Main Sequence, and units for mass and luminosity are
in terms of the Sun - how many times the star's mass or luminosity is greater or lesser
than the Sun. Figure 14 shows this relation.
Figure 14. The masses and
luminosities of 250 binary stars are
plotted up. The values for mass and
luminosity are given in terms of the
Sun's mass and luminosity. Many of
these stars are not found on the
Main Sequence, so there is a great
deal of scatter in the data points. A
relatively straight line relation can
be seen here, which indicates that
there is a simple relation between
the masses and luminosities,
especially for Main Sequence Stars.
The Main Sequence is very well
ordered in several respects - stars
range from low temperature,
luminosity and mass (those in the
lower right corner) to stars of high temperature, luminosity and mass (upper left
corner). Why is it like that? How do the other groups of stars fit into all this? Do stars
136 | P a g e

ever change their characteristics? Do you really have to know all of this stuff? Yes, of
course you do.
As more and more people started to classify stars according to their spectral types and
luminosity classes, they noticed some interesting trends. When you look at the number
of stars of the different spectral types out there, you may note that most are located at
the low temperature end of the Main Sequence. These stars are faint, small, and red,
so they tend to be referred to as red dwarfs. While these are not the brightest stars out
there, they are the most common. K and M types easily outnumber the O and B types
by a wide margin. It is likely that the L and T types actually outnumber all other
types, but the problem with them is finding them - they are just so faint and are very
difficult to detect so we don't have accurate statistics about them.

Now that you've read this section, you should be able to answer these questions....
What is the difference between apparent magnitude and absolute magnitude?
How is parallax related to distance?
What methods are used to measure the temperatures of stars?
What is the spectral classification system and what makes it vary from star to
star?
What does the luminosity of a star depend upon?
What is shown on an H-R diagram?
Where is the Sun located on an H-R diagram?
Where are most of the stars located on an H-R diagram?
How can Kepler's laws be used with binary stars?
How does the luminosity of Main Sequence stars relate to their masses?
Stellar Birth and Middle Age

What's covered here:
How do stars form? What are the basic ingredients? How does it occur?
How did a star like the Sun form?
What happens during the Main Sequence phase of a stars life?
How do stars create energy?
What is the internal structure of a star like?
What determines the life span of stars?

137 | P a g e

Star formation
Before we start with the ways that stars are formed and evolve, there needs to be some
explanation as to how astronomers can actually talk about this stuff. Typically stars
take millions or billions of years to live their lives, so it is not possible to sit down
behind a telescope and watch a star live out its entire life. We just don't live that long.
However, we do have some things to help us understand what is going on. One of
these tools is the H-R diagram. We can't actually see an individual star evolve, so we
need to look at a large number of stars at once. When you observer many stars, you
have a pretty good chance of catching stars in different stages of their lives. The more
stars you observe, the greater your chances of finding a star in one of the shorter,
harder to catch phases. If you observe thousands or millions of stars and describe their
characteristics on the H-R diagram (temperature and luminosity), you will see that
there are patterns to their distributions on the diagram. The information in H-R
diagrams, coupled with computer models and physical theories, help us to understand
processes that we can't even see - or at least they give us a good starting place to work
from. By observing as many stars as possible, we have a better chance of piecing
together the picture of stellar evolution. What characteristics of a star change over
time? We can see the changes in temperature and luminosity, which results in their
changing positions on the H-R diagram. These are both coupled with radius, so that
changes as well. The masses of some stars are observed to change, and to do so
requires certain processes to operate inside the stars. We also know that the generation
of energy inside of a star will change its internal composition, especially in the core.
When you change its composition, you change its internal structure, which of course
has an influence on how it gives off energy and how hot its surface is and how large
its radius is and so on. In this way everything is sort of tied together - mass,
temperature, luminosity, radius, and chemical composition. As you'll soon learn, one
of these physical characteristics will really have the greatest influence and basically
determines most of the other characteristics of a star (you'll have to keep reading to
find out which one is the most important; you think I'd give away the answer this
early?).

When we describe the changes in a star's temperature and luminosity, its position on
the H-R diagram changes - we often stay that it "moves" on the H-R diagram. THIS
DOES NOT MEAN THAT IT MOVES IN SPACE!!! The "motion" is just the
changes in the temperature and luminosity of the star that we can observe. Stars do
move through space but this doesn't usually effect their lives. Don't get confused by
the references to "moving" on the H-R diagrams - it's just showing how they change
their surface temperature and luminosity over time.
Raw Material
138 | P a g e

As you just learned, stars are made primarily of hydrogen and helium with only a
small fraction of their mass (about 2%) made up of all of the other elements. So in
order to make stars you need a lot of hydrogen and helium. When we look out into
space, we can easily detect many clouds of hydrogen floating around in our galaxy.
Helium isn't easy to detect since it is a noble gas and doesn't show up in surveys of the
sky using any type of telescope.
While there is mainly hydrogen floating around between the stars, in what we call
the interstellar medium, there is also a little bit of dust also there - that's part of the 2%
that makes up stars. This dust is very important for several reasons. Since each dust
particle is more massive than a gas particle, it has a greater gravitational pull and
would help in the process of forming stars. Also it can block energy/light that could
break apart molecules. Another very important aspect of dust is its influence in our
view of the sky. Dust is so good at blocking light that star light can not easily travel
through clouds of dust and because of this many stars are not visible to our eyes. And
if the dust does not entirely block the light it can just make stars look dimmer - this
can really mess up our calculations of their distances since we think that the stars are
fainter than they really are.
When you look up at the fuzzy band of the sky that we call the Milky Way you are
looking at the thickest concentration of the gas and dust in our galaxy, and also where
a lot of star formation is occurring - where the hydrogen and helium gets concentrated
into stars. But enough about the raw ingredients, let's get to making some stars!
Stellar Birth
Observing stellar birth is rather tricky, since it happens so quickly (relatively
speaking), and it is often not directly visible but is buried under layers of gas and dust.
There are actually two main types of star formation - the large and small scale, or
simply the formation of many stars at once and the formation of only a few stars.
Obviously the large scale star formation will be easier to observe and it is sometimes
really apparent. We'll look at that process first.
Large Scale star formation
When we look into the sky, we often observe stars in groups - these groups are often
dominated by massive, young, hot stars. By this I mean these hot, big stars outshine
all of the other stars, so that you might not even notice the small stars that are there.
These stars are gravitationally bound together, so it is reasonable to conclude that the
stars were formed together. To form so many stars (hundreds, thousands or more)
requires a lot of resources. What is needed for large scale star formation? You need to
139 | P a g e

have a lot of the basic material that goes into stars, hydrogen and helium (or H and He
for short).
Looking out into the galaxy, we can find many Giant Molecular Clouds (GMCs).
As the name implies, these are large clouds of gas and dust. They have very
distinctive characteristics:
Masses of these clouds are typically on the order of millions of solar masses,
and on some occasions up to billions of solar masses (that's the Giant part).
They are cool, around 10 K (this is logical since there are a lot of molecules in
them which would not normally exist in a hot environment).
Gases are generally found in molecular form, with such molecules as H
2
, CO,
CO
2
, CH, H
2
O, SiO, etc. Actually, about 150 different molecules have been
found in GMCs, some with up to 70 atoms! Some of these molecules are linked
with the formation of life (as seen in this recent finding from
the Herschel telescope).
There are a few thousand GMCs in our galaxy, and they tend to be found in the areas
of large scale star formation and near very massive, hot stars. GMCs are so cool that
the only way that they can be detected is with telescopes that detect their very low
energy light - in this case radio telescopes would be best. They are very good for
detecting complex molecules in space. You might know these regions as the Spiral
Arms of a galaxy (we'll get to those later on in the course). GMCs are basically the
raw materials that make up stars. There is only one problem - they don't want to be
stars (obviously they should not go to Hollywood). Cool gas clouds tend to want to
stay as they are - as cool gas clouds. In this way, GMCs are sort of like Nerf balls -
they resist squeezing. If you were to squeeze them together just a little bit, they would
tend to heat up, and this increases the motions of the material in the clouds (atoms and
molecules), which increases the gas pressure, and this will counter-act the squeezing.
If you want to make stars, you have to squeeze the GMCs so hard that they can't resist
or overcome the force of gravity! Somehow or another a very strong force of
compression must be exerted upon a gas cloud, so much so that it can't counteract it
with an increase in the gas pressure. After all, without the compression, there would
be no stars formed and if that were the case, then we won't have anything to talk about
- and that would be really boring, wouldn't it? Where does the compression come
from? It doesn't really matter where it actually comes from, it just has to be strong
enough to counteract the ability of the cloud to resist compression. You'll later learn
about some of these compression mechanisms that are out there that can do the trick.
Let's say we get the thing compressed by some strong force. What happens to a GMC
when you do compress them?
140 | P a g e

As with anything, when you start to compress it, material gets closer together.
According to Newton's law of gravity, a decrease in distance leads to an increase in
gravity (remember, gravity goes up as distance goes down). The material is feeling a
higher gravitational pull (the gravity that each bit of the cloud feels from all of the
other bits). If gravity is strong enough, it will bring the material together even more,
which will increase the gravity and bring the stuff even closer, which increases the
gravity... you sort of get the idea. Once you overcome the barrier of the cloud's
resistance to the compression, it will pretty much give up and start collapsing down.
This will continue until the individual stars form - remember, stars are just big gas
balls, so as various parts of the cloud clump together, they will form clumps of gas
which become stars. Figure 1 shows how this happens. You might want to make note
of the fact that only part of the GMC will get compressed and have star formation
occurring - not the entire cloud. With each batch of new stars forming, there could be
hundreds or thousands of stars created at one time.
Figure 1. The compression of a GMC occurs only in
a small part of the cloud. The rest of the cloud is not
effected. It should be noted that the colors shown
here are not accurate - in general you can't even see
the clouds since they are so dark. They would
appear like a black region to the eye.
What kinds of stars will form? Will they all be big stars? Will they all be little stars?
The variety of stars that form is sort of like the variety of pieces of glass you get when
you break a window or a drinking glass. Usually there are many small pieces - these
would be the very low mass stars, which (when they reach the the Main Sequence) are
the K, M, L and T types (less massive than the Sun). There will be few of the more
massive stars, the A, F and G types (which have masses similar to the Sun) and very
few of the really big stars, the O and B types. This kind of makes sense, because we
see very few O and B stars out there - they are pretty much outnumbered by the little
stars. As you'll see there is another reason why O and B stars are so rare, while the
little ones are so common, but we'll get to that
later.
Figure 2. The part of the cloud that was
compressed breaks up into stars in a distribution
with few large, hot stars (OB stars); more middle
sized stars (AFG types); and many, many cool,
small mass stars (KLMT types). The coolest stars
would not be visible to the eye since they are mainly infrared sources.
141 | P a g e

Even though they are greatly outnumbered, the big stars, the O and B types, are the
most important ones in the group. Why are they so important? What characteristic do
these stars have that set them apart from the other stars? They are very hot! These
beasts are so hot that they give off a lot of UV radiation. UV radiation is very
important since it can ionize the gas around the stars (it ionizes the hydrogen mainly).
Remember, when light ionizes something it means that it knocks an electron off of an
atom. This happens quite a bit, as does the reverse process (the electron getting back
into an orbit around the atom). The end result is that the gas glows. You get a lot of
hot glowing hydrogen gas and perhaps other types of gas that are glowing, though
there is much more hydrogen around the star formation region than anything else.
Now we have a region around the hot O and B stars that is just a huge cloud of hot
glowing ionized gas (hydrogen). Such a region is called an H II region - the Roman
numeral II means that one electron has been lost. How would you pronounce the name
of this region? You say "H two." That makes it a bit confusing, since it sounds the
same as saying H
2
, which is how we refer to molecular hydrogen. Yes, it is confusing,
but we like it like that. Just remember, H II indicates that you are talking about hot,
ionized gas, while H
2
indicates that you are talking about cool, molecular gas.
We have to get back to the star formation region and see what's happening there. The
O and B stars are quite effective at ionizing the hydrogen. They can ionize a large area
around them, producing a very bright H II region. When you observe these regions
you are basically seeing hot, thin, glowing gas (mainly H). If you were to look at the
light from the H II region, you could obtain a spectrum of it - what type of spectrum?
An emission spectrum - remember, that is what is produced by a hot gas. The
emission spectrum of hydrogen has a very strong red line in it, so there is often a
rather pink-ish glow to these regions. Another tell tale sign that you are looking at an
H II region is that there are often traces of the cool molecular gas that usually look
like dark blobs. This is because it has dust in it and the gas in these dark clouds isn't
hot enough to emit light, like the H II region gas can. It is sort of neat that these very
different gases can be right next to one another. Other things to look for in the H II
region are the culprits that are causing all
the trouble - the O and B stars. These
will be the brightest of the stars in the
area, so they often stand out quite well.
You may also note that these stars tend
to be very bluish. Figure 3 shows the
characteristics in and around an H II
region.
Figure 3. The Triffid Nebula (M 20) is a
good example of an H II region since it
142 | P a g e

shows the different features around it. First of all is the H II region itself -
distinguished by the pinkish glow. Next is the presence of dark dust that blocks light.
This is referred to as a dark nebula or just basically dust. However, under certain
conditions, it will not look black but instead will appear blue as is seen on the right. A
reflection nebula is just a region of dust that is reflecting blue light towards you.
These bluish regions are not always seen around H II regions but are seen often
enough to be recognized. They can also be seen in other circumstances. At the center
of the H II region are the hot stars that are maintaining the high temperature of the
gas in the region. The image is over exposed so that these stars are not visible
amongst the hot gas. To see another view of the Triffid Nebula, just click here. This
shows both the visible light view as seen at the left and an infrared view, as seen by
the Spitzer telescope. The IR view shows the extent of the gas and dust, often in areas
that aren't seen in the visible light view. AAO, photo by David Malin
There are quite a few examples of H II regions, though one of the most spectacular is
the Orion Nebula (Figure 4), which is visible in the winter sky. Other star forming
region are shown in Figure 6.


143 | P a g e

Figure 4. A large view of the Orion Nebula,
visible in the winter sky. This image is about
1 degree wide. AAO/ROE, photo by David
Malin
A Hubble Space Telescope view of the inner
region of the Orion Nebula. Compared to the
other image, this covers only a tiny fraction of
the bright inner core. To see how the image
was obtained by the HST, you can view a little
movie of it here. Image credit: NASA and C.R.
O'Dell and S.K. Wong (Rice Univ.)

Figure
5. Two
Hubble
Space
telescope
views of
the central
part of the
Orion
Nebula.
The image
on the left
shows the
visible light
view. In
this view
you can see
the four hot
stars in the
center that
form the
"Trapezium
" and
provides
most of the
energy that
keeps the
nebula hot.
Also,
individual
gas and
dust clouds
are visible.
The image
on the right
was
obtained
144 | P a g e

with an
infrared
camera on
the Hubble,
and in this
instance,
the many
small,
newly
formed
stars are
visible.
Notice how
many of
these stars
are not
visible in
the image
on the left.
Credits for
near-
infrared
image:
NASA; K.L.
Luhman
(Harvard-
Smithsonia
n Center
for
Astrophysi
cs,
Cambridge
, Mass.);
and G.
Schneider,
E. Young,
G. Rieke,
A. Cotera,
H. Chen,
M. Rieke,
R.
Thompson
(Steward
Observator
y,
145 | P a g e

University
of Arizona,
Tucson,
Ariz.)
Credits for
visible-
light
picture:
NASA, C.R.
O'Dell and
S.K. Wong
(Rice
University)
.

Figure 6. The Eagle Nebula (M 16) - this is
another region of star formation, but unlike
the Orion nebula, the stars are not formed on
the inside of the cloud but are being formed on
the outer edges. The huge pillars of gas shown
are quite large, with the one on the left being
about 1 light year long from top to bottom.
There are bright stars located above the

Close-up This shows the top of the pillar on
the left. The long tube-like structures on the
edges of the cloud are parts of the cloud that
are breaking away and forming into stars.
These areas appear to contain a concentration
of gas and dust that doesn't get entirely blown
away by the hot massive stars located above
146 | P a g e

A bunch of stars have been formed. What happens now? It is possible that the
formation of stars will lead to further star formation. How does that happen? The new
stars, again the O and B ones especially, have some rather strong winds. These winds
can compress other parts of the GMC and you know what happens when you
compress a gas cloud, right? Star formation! Tt is possible for the whole process to
continue like a domino effect until the cloud is gradually used up. In a way, the star
formation process slowly eats away the GMC. This also tells us that when you see a
region of star formation, there should be a pretty big gas cloud in the area, much
larger in size than the H II region that is visible to your eye. There can also be a rather
negative side effect to having such massive stars involved. As just mentioned, these
stars have pretty strong winds. In Figure 6 the effects of some of these strong winds
are obvious, they can blow apart gas clouds - so it is possible that some stars-in-the-
making could be destroyed in the process. Take a peak at Figure 7 for a view of some
of these wanna-be stars (which we call proplyds), and you may also want to check
out this animation showing the inside of the Orion nebula and the future evolution of
the stars in it (we'll get to that later). In the animation you may notice how many of
these blobby stars are elongated - stretched out - by the strong winds from the hot
stars near the center of the nebula. Here is an animation showing the zoom into
the Hubble image and revealing the proplyds as they really are - some of these are
shown below in Figure 7. In general these objects are several times larger than our
solar system (hundreds of A.U.s in size), but as they evolve, they'll lose their cocoon
of gas and dust.


Figure 7. Two stars in the process of
forming in the Orion Nebula, as seen by
the Hubble Space Telescope. These
"proplyds" (as they are called) are more
gas cloud than star at this point and are
feeling the effects of the strong stellar
winds near them. Each proplyd is several
times larger in size than our entire solar
system and they have a long way to go
until they become stars - mainly they have
to shed a lot of their outer layers. Image
credit: NASA, J. Bally (University of
pillars (which are visible in x-rays as can be
seen in this image) which are blowing away
gas from the pillars. The gas that remains
behind forms into stars. Image credit (both
images): Jeff Hester and Paul Scowen
(Arizona State University), and NASA.
the cloud. Whatever remains behind will
eventually become a star.
147 | P a g e

Colorado), H. Throop (SWRI), C.R.
ODell (Vanderbilt University).
Of course, the formation of so many stars at once is easily noticed, particularly
because many of these stars stay in groups or clusters. Sometimes these clusters are
sitting right next to H II regions - which also helps us see where star formation is
going on in the galaxy. Where the stars form in a relatively loose group that
eventually breaks down (what's known as an association), it is harder to track the
point of origin for the stars.

The Orion nebula has been mentioned as a good example of an H II region, and it is
also one of the most studied ones out there. This is due in part to its relatively close
distance to us (about 1500 light-years away) and the fact that the part of the nebula
facing us is the part where all the action is. You are sort of looking straight in at the
star formation, so it is rather obvious what is going on. The hot stars really give it
away. Often, we don't have such a good view. If we were on the wrong side (the cool
dark side) of the GMC, we would not be able to see any hot stars or H II regions since
the cool gas and dust would block our view. To see anything that is made up of cool
gas and dust, we cannot use visible light telescopes, but rather IR telescopes. The IR
images that are of star forming regions show not only the stars and gas associated with
the star formation but also the gas that isn't currently involved in any star formation,
since cool gas and dust show up easily at these wavelengths. IR telescopes reveal the
presence of a lot of "invisible" material, which can eventually go into the formation of
stars. Figure 5 shows many of these strong, newly formed infrared stars in the Orion
Nebula. Here is an image from the VLT (visible light) and the Spitzer space telescope
showing a region of star formation. The cloud's dark dusty appearance hides all of the
star formation that is going on, which can be seen with the Spitzer's infrared eye. It
also helps that the stars in this cloud are also "misbehaving" by blasting material out -
a good way to spot young stars as you'll see in the next section. The basic lesson here
is don't just trust what you can see with your eyes - use other "eyes" like radio and
infrared to get the entire picture!
GMCs are needed for large scale star formation. The formation is so massive that it
can be seen over a great distance - even in other galaxies, since OB stars and H II
regions are very bright. However, not all stars are found in such groups, so there must
be a way to do small scale star formation. After all, the Sun is not part of a large group
of stars - so we need to make stars on a smaller scale. We'll look at how that happens
next.
Small Scale Star Formation
148 | P a g e

It is possible to have only a few stars form, rather than thousands of stars forming.
These are formed from obviously much smaller clouds of gas and dust. From such
clouds only a few stars or maybe only one star will form. Such clouds are typically
only around 10 solar masses in size. These gas clouds are pretty small and difficult to
find due to the fact that not only are they small, but they tend to be cool - they don't
give off any or hardly any visible light. I guess we could call them PMC - puny
molecular clouds.
Figure 8. A Hubble Space Telescope image
of some dark, small mass clouds. While
these clouds look pretty dusty, they really
don't have that much dust in them; maybe
only 1% is dust. The dust, though, is very
efficient in blocking out the light. Image
credit: NASA and The Hubble Heritage
Team (STScI/AURA).
They are already small to begin with, so it is
easier to get them to compress, though there
is less mass involved (remember, gravity
depends strongly on distance, with a small
distance having a very large gravitational
pull). As in the case of large scale star formation, these clouds will give off light
primarily in IR wavelengths. However there wouldn't be any easy to see H II regions,
since it is unlikely that large stars will form in this small scale process. The stars that
do form are buried inside of the dusty cloud and are rather cool protostars. Protostars
are what you'd call something that isn't quite a star yet, but is more star-like than
cloud-like. The reason it isn't a star yet, is because it hasn't started any fusion
processes - the mechanism by which stars make energy. I'll get to fusion in a little,
just hang on a bit. It is very difficult to spot these critters, since not only are they cool
stars, but they are also buried deep inside the dust and gas cloud.

As the small gas cloud compresses, it often flattens out into a disk. Why does it do
that? It is sort of like the way that good old fashioned pizza is made. When the dough
is being shaped, it is often spun around - this stretches it out. In the case of these small
gas clouds, they are already spinning around, and spinning things that are compressing
have a tendency to flatten out into disk shapes. It is possible to find these stars by
looking for the disks around them - since the disk can be much larger than the star.
Figure 9 shows some of these disks around young newly forming stars. In case you
were wondering, yes, this is a way that planets can get formed - the material in the
disk can eventually form into planets. Now you should be wondering about whether
149 | P a g e

disks would form around the big stars - you were wondering that, weren't you? It is
possible that disks could form around them, but since they are in groups, there is also
a greater chance that the other stars in the area could mess up and destroy any disks.
Also, the very strong winds that are produced by massive stars would disrupt a disk of
material very quickly.
Figure 9. Disks around young
protostars are shown in these
two Hubble images. In both
cases the disks are tilted so
that they are seen edge on.
There is dust in the disks, so
they tend to look very dark and
are hard to see. This is
especially true for the disk on
the right. Some hot gas around the newly forming star is seen above and below the
disk, while material getting shot out from the poles (which look green here) indicates
that there is a young star in the middle (this will be explained further down the
notes). Image credit: Chris Burrows (STScI), John Krist (STScI), Karl Stapelfeldt
(JPL) and colleagues, the WFPC2 Science Team and NASA.
A star with the mass of the Sun in this stage of its life would have a surface
temperature of only about 4000 K and a radius about 20 times the Sun's current
radius. The luminosity is rather large, mainly due to the large radius, typically about
100 times the Sun's.

These stars are cool and luminous. Are these stars Red Giants? No, as you'll learn,
Red Giants are stars near the end of their lives, while these protostars haven't even
ripened (haven't "turned on" yet). They do have similar outward appearances, but
what is going on inside a Red Giant and a protostar is quite different (of course we
can't see inside of them directly; we just think we know what is happening inside the
stars - I'll talk about this more later). Also, protostars in this stage of their lives evolve
very fast, so it is hard to catch them during this stage. By "fast" I mean around a few
million years for the really small mass protostars to get their act together (become real
stars) and only a few thousand years for the very large mass protostars to become real
stars. Figure 10 shows where these stars are on the H-R diagram and how they change
their temperatures and luminosities as they evolve.
150 | P a g e

Figure 10.Early stages of
stellar evolution, pre-main
sequence. As gas clouds get
converted into protostars
and the protostars gradually
stabilize into stars, they
evolve in a generally left-
ward motion on the H-R
diagram (towards higher
temperatures). The lines
indicate the paths on the H-
R diagram that protostars of
different masses will take as
they head towards the Main
Sequence. Graph based
upon stellar evolution
computer models of Siess, Dufour and Forestini.
Also with small scale star formation, it is very difficult to find protostars when they
are still surrounded by gas and dust. Once they do get close to the Main Sequence,
they often start making their presence known by being rather active. I like to think of
this as their "terrible twos stage," sort of like a young child having a temper tantrum -
these small stars do sort of the same thing. This stage, known as the T Tauri
Stage (TT - just like "terrible twos") only happens to small mass stars like the Sun.
What they do is develop very high velocity winds which blow off their outer layers
and the surrounding material (mainly hydrogen). On occasion, this material will run
into other material in the region and form shock fronts, what we call H-H Objects.

We can see that the various H-H Objects are moving either towards or away from us,
so they have a pretty good velocity away from the protostar. Figures 11 and 12 show
some views of T Tauri stars and their associated activities, including H-H objects.
151 | P a g e


Figure 11. Some time lapse images of T Tauri
stars. These are images obtained from the
Hubble Space Telescope showing the mass
ejections due to strong winds from young stars
(really still protostars). On the left is XZ Tauri,
actually two stars in orbit about one another.
The bubble of material bursting from them is
moving at a speed of about 300,000 mph. On the
right is HH 30 - the T Tauri star that has a disk
of material around it and is erupting matter in
two directions - you're just seeing one gusher
here. Image credits: (left) John Krist (STScI),
Karl Stapelfeldt (NASA JPL), Jeff Hester
(Arizona State University), Chris Burrows
(ESA/STScI); (right) Alan Watson (Universidad
Nacional Autonoma de Mexico, Mexico), Karl
Stapelfeldt (NASA JPL), John Krist (STScI), and
Chris Burrows (ESA/STScI).
Figure 12. Several T Tauri
stars producing some H-H
objects. These Hubble images
show mainly the H-H objects,
since the T Tauri star is not
visible behind a layer of gas
and dust or inside the disk of
material. However, the stream
of material being ejected by
the star is visible, and the pile
up of the material in the form
of the two H-H objects is also
easily seen. You can click on
the image to see a larger view.
The scale in each image is the equivalent to 1000 A.U., or 1000 times the distance
between the Earth and the Sun. Credit: C. Burrows (STScI & ESA), the WFPC 2
Investigation Definition Team, and NASA.
It is interesting to note that the material is being ejected from the young star only in
two directions, known as a bipolar flow. Why doesn't it just blow material out in a
uniform direction or into a random direction? It is possible that there is some
influence from the magnetic field of the star and/or the presence of a disk of material
that would obstruct the flow of material in any random direction. In some cases disks
are visible, but not all. As you'll see, a lot of things have a bipolar outflow - this is a
152 | P a g e

very popular way for objects to spew material out into space. The T Tauri stage is
short lived, and the star eventually settles down and becomes a boring, normal Main
Sequence star.
On the Main Sequence (MS)
The "adult" stage of a star's life is when it becomes a Main Sequence (MS) Star.
When does this happen? A star becomes a MS star when it startsfusion reactions in its
core. This is the process that produces the energy in a star's core. Along with these
fusion reactions, there are other characteristics of a star that help astronomers
understand the nature of stars. Some of these properties make MS stars distinct from
others and help us understand the way in which stars evolve. You have to remember
that we have never seen inside a star, that we can only study what is on the surface,
but there are many ways of figuring out what is going on deep below the surface.
Fusion
How is energy in a star produced? Originally people thought that the Sun was actually
something that was on fire, like a big block of wood or some sort of "normal" fuel.
When people determined how long a big block of wood, gasoline or some other
"normal" fuel would last, they determined that the age of the Sun would be really
young, since these types of fuels don't last very long. There must be some other
mechanism that is creating the energy. We know that gravitational compression
(squeezing) produces heat, so could gravity be doing it? No, there isn't enough mass
in the Sun to produce all of the energy we get from it. Eventually, Albert
Einsteinprovided the answer in his Special Theory of Relativity (1915). While there
were several different formulas and concepts in this theory, the most familiar is the
famous equation
E=mc
2

What does this mean? All it really says is that if you convert mass into energy, you get
a certain amount of energy out of the reaction that only depends upon the mass
involved - the more mass, the more energy. The formula doesn't tell us how the energy
is produced from the mass, but just gives us a way to measure the amount of energy
produced. Other physicists came along and determined the processes where the mass
is converted into energy, but Einstein is the one who showed that mass and energy are
related. As you'll see this formula can go either way - you can convert mass into
energy or energy can be converted into mass.

How do you convert mass into energy? I'm glad you asked. The process is known
as fusion, or the fusing together of atoms. This is not what goes on in nuclear power
plants, and actually it is rather difficult to do even the simplest fusion reactions, which
sort of explains why we have no nuclear fusion plants on the Earth. These reactions
153 | P a g e

require very high temperatures and pressures to work - the main reason it is so hard
for people to duplicate this process on Earth. Sometimes we use the word burning to
describe the energy production in a star, but this is not a very accurate term, since the
material isn't really burning - it just sometimes makes it easier for people to visualize
the process.
The main reaction operating in the Sun and other low mass MS stars is the Proton-
Proton chain (or the p-p chain). To get this reaction to work you need a temperature
of at least 13 million K and a density of about 100 gm/cc. If you are not familiar with
density, then you might find it interesting that the density of metal is around 7 gm/cc.
This is much denser than most metals that you come into contact with. Something you
might also want to remember is that a proton is really just a hydrogen atom without an
electron (an ionized hydrogen atom). The core of a star is very hot, and the atoms are
colliding around quite a bit, so it is really easy for the atoms in the core to get ionized.
There are a whole bunch of protons just bouncing around in the core - but you can
also think of them as hydrogen atoms bouncing around in the core - it doesn't matter
how you visualize it.
Here is the step by step process:

First have two protons come together and form
into deuterium (H
2
), with by-products of a positron and
a neutrino. Deuterium is an unusual form of hydrogen, sort of
like over-weight hydrogen. The positron (e
+
) is just like an
electron, except that it is positively charged. A neutrino ( )
is a weird little particle that is very difficult to detect. Here's
the reaction written out:
H
1
+ H
1
H
2
+ e
+
+

Now have the deuterium fuse together with another proton to
form a type of helium that is a bit light weight (He
3
) and also
a gamma-ray photon . The gamma-ray is the energy that the
star has produced. Here it is written out:
H
2
+ H
1
He
3
+

Take the light weight helium and have it combine with
another light weight helium to produce normal helium (He
4
)
with two protons left over .
He
3
+ He
3
He
4
+ H
1
+ H
1


154 | P a g e

Basically, you start with four protons (hydrogen nuclei) and end up with a helium
nucleus (which contains two protons and two neutrons) and some other stuff (the
positron and neutrino, oh yes, and some energy). If you were to put this stuff on a
scale and measure the mass of what went into the cycle and what came out of the
cycle, you would see that they don't weigh the same. The difference in weight
between these items is the mass used to create the energy in the reaction. Even though
only a minuscule amount of energy is produced in each reaction, stars do this reaction
so many times that there is a great deal of energy given off. A star like the sun does
this reaction about 100,000,000,000,000,000,000,000,000,000,000,000,000 times each
second. Even though a single reaction doesn't produce enough energy to keep a gnat
alive, stars produce a huge amount of energy, since the reaction is performed so often.
In stars more massive than the Sun, another fusion reaction is at work, the CNO cycle.
This does basically the same thing as the p-p chain but uses other elements, namely
carbon, nitrogen and oxygen, to bring about the production of energy and the other
byproducts (Helium, positrons, and neutrinos).
Fusion is the source of energy (photons) in a star. The density of the interior of stars is
so high that the radiation can't go very far without being absorbed, deflected and
bounced around by all the stuff on the inside. Actually, it is rather difficult for the
photons to go very far at all without hitting something. The path that the light takes
from the core to the surface is rather chaotic.
The light produced in the core will take a Random Walk towards the surface. This
little walk can take around 200,000 years or even more. At the onset of the walk, the
photon is a high energy gamma-ray, but it loses most of its energy in all those
collisions it has as it works its way through the Radiative Zone. In this region,
radiation (energy, light, photons, etc.) is transported according to the rules that govern
how radiation interacts with materials - all of those collisions, absorptions, and
emissions that go on in the dense material. That's why this layer in a star is called the
Radiative Zone.
Figure 13. Energy produced in the center of
the Sun (or any other star) bounces around
for years as it works its way towards the
surface. This Random Walk is due to the way
that matter and energy interact. The process
will eventually stop once the density of the
material is too low for there to be significant
interaction - once it gets out of the radiative
zone in the Sun.
155 | P a g e

Further out from the center, the material is less dense, so the radiative processes are
not as important. Here the energy is transported via convection in what is called
theConvective Zone (isn't that clever?).
Remember, convection is also the process you see when you boil a pot of water on the
stove. Energy is transported by the motion of the material - the churning and bubbling
of the gas. At the top of the convective zone, the gas is thin enough for radiation to
easily escape - these are the layers of the star's surface that we see - the atmosphere. If
you remember the stuff about the Sun's surface, the tops of the convective bubbles are
what we call the granules. The situation for larger stars is actually quite different -
they have the convective zone near the core with the radiative layer near the surface.
This just has to do with the way that the material acts under different circumstances -
whether there is convection or not going on. Either way, energy produced in the core
of a star, where it was originally a high energy photon, will eventually get to the
surface of the star where it can then fly off into space in the form of a relatively low
energy photon, usually in the visible part of the
spectrum.
Figure 14. The interior of a star like the Sun is
shown. The three main layers of the core,
radiative zone and convective zone are depicted.
For larger stars, the order of the various regions
in the interior does not appear to be the same as
is shown here. Those stars tend to have
convection near their cores and their radiative
layer just below the surface.
Stellar interiors
How do we know what the inside of a star looks
like? How can we really talk about fusion and such with any confidence when it is
pretty much impossible to see anything below the surface of a star? There are several
ways to study the interior of a star. One already mentioned is helioseismology, which
can also be applied to other stars (where it is referred to as asteroseismology). The
study of a star's light is not a good indicator of the interior since it is altered so much
on its journey from the center, and it is also old light - it takes a long time to travel
through all those layers. The only thing that comes immediately from the center of a
star is the neutrino, one of the by-products of the fusion reaction. Neutrinos are really
weird little critters. These are particles that we don't know much about, in part
because we have a very hard time detecting them. There are thousands of neutrinos
going right through your body right now. Does it hurt? No, because neutrinos travel
through pretty much everything without causing any problems - they don't interact
156 | P a g e

well with matter. That's why it is so difficult to study them - we can hardly ever catch
them! It isn't really impossible to study them, but it is just very difficult. Neutrinos are
so elusive, that we are not even sure if they have mass. Our best current estimates put
there mass at less than 5 x 10
-37
kg, though we aren't certain about the lower limit to
the mass. For comparison, a proton has a mass of around 1.67 x 10
-27
kg, which means
that a proton is at least 3 billion times more massive than a neutrino.
Actually, there was a major problem with neutrino studies of the Sun. There have
been neutrino experiments going on for some time, some working for over 30 years,
and all of the detectors got basically the same result. We detected fewer neutrinos than
we were supposed to detect based upon the laws of physics. This used to be a major
problem in astronomy since it has some rather nasty aspects. Why were too few
neutrinos detected? Was there something wrong with the inside of the Sun? If that
were the case, it may be bad news for us, since our existence depends on the Sun. Was
there something wrong with all these theories that we have to explain what happens
inside of stars? If that were the case, we would have some serious conflicts between
the rules that we are using and the experiments where the results do come out okay. Is
there something wrong with the neutrino detectors? Well, every neutrino detector is
different, and every detector seemed to get the same result - that there were too few
neutrinos coming from the Sun.
Figure 15. The Super-Kamiokande Neutrino
detector in Japan. This large vat of ultra pure
water will detect different types of neutrinos.
The spheres seen along the side of the tank are
light detectors that register the rare
interaction of a neutrino with a water
molecule. Image courtesy of ICRR (Institute
for Cosmic Ray Research), The University of
Tokyo .
What's the answer? It seems that neutrinos
aren't all alike. Actually, we knew this
previously, but we didn't know that the
neutrinos coming from the Sun do change in
subtle ways, becoming a type of neutrino that
is incredibly hard to detect. Before, neutrinos
were just really difficult to observe, but this
other type of neutrino was so close to
impossible to detect that no one ever saw it.
That was until a really amazing neutrino
detector was built (see Figure 15), and it started to pick up all these very difficult to
157 | P a g e

detect neutrinos. It seems that the neutrino problem is no longer a problem.
Sometimes it takes better theories to figure out what's going on; sometimes it takes
better equipment.

Looking at neutrinos helps us to figure out what the Sun is like, but other stars are so
far away that it is nearly impossible for us to look for their neutrinos. How can
astronomers figure out what is going on inside all of the other stars out there? How do
we know if there is convection or radiative energy transport occuring inside of them?
In order to understand how stars work, astronomers must apply several laws of
physics to them and these laws, combined with observations, provide astronomers
with a better pictures of stellar physics. These rules are
Hydrostatic Equilibrium - This is the balance between weight (gravity) and
pressure (air pressure or gas pressure). This is like how the old dome on the
UNI-Dome worked (and how the Metrodome in Minneapolis works). The cloth
dome was held up by air pressure inside of the UNI-Dome, while the weight of
the dome was always exerting a downward force due to the pull of the Earth's
gravity. The air pressure and the gravity need to be evenly balanced so that the
roof doesn't cave in - which happened at those times when it wasn't balanced. A
star is similar. The outer layers of the star are being continually pulled in by the
star's gravity. These layers don't fall in because the high pressure of the gases
inside of the star prevent that. The star is balanced - it doesn't cave in on itself
or blow itself apart.
Conservation of Energy - Energy is given off but it is continually produced.
Another way of saying this is that stars keep on shining; there are no gaps in the
production of energy. This is important because it helps regulate the
temperature and density inside of a star. The energy from the core flows out
from a star continuously - it doesn't get trapped inside various regions but flows
through them. This is a good thing, because if it didn't happen, if energy got
stuck in a certain layer, then that layer would heat up to an abnormally high
value, which would screw up the internal structure of the star.
Conservation of Mass - A star's total mass is the sum of all of its layers. This
one actually sounds kind of silly, because it is saying if you take all the parts of
a star and add them up, you'll get the total mass of the star. What this rule
really does is account for the fact that not all layers of a star have the same
mass - some layers are thicker than others; some are thinner since the density
inside of a star changes from one layer to the next. You have to account for this
change in density when you do add up all the layers. It is actually rather
important.
Energy Transport laws - These are the rules governing energy flow
mechanisms such as convection and radiative transport. These rules also tell us
158 | P a g e

how much energy is being transported by these various mechanisms; in some
layers it is all radiative, in some layers it is all convective, and in some layers it
is a mix of both.
By combining these rules with other laws of physics and observations, astronomers
can make computer models of stars which can then be altered to show how stars will
live their lives. Of course, the results from the computer models can be compared to
observations of stars. This makes life pretty easy for astronomers since stars take a
very long time to get from one phase of their lives to another, so sitting behind a
telescope and waiting for a star to change is pretty much a waste of time. A computer
can speed up the process quite a bit.

What is the inside of a star like? We know that various physical characteristics like
temperature, density and pressure change as you go from the center of the star to the
surface. In the center the temperature and density values are the highest (temperatures
of several million K, for example), which is a pretty good thing since there would be
no fusion without these very high temperatures and densities. As you get further and
further from the center of the star, the temperature and density decrease quite
drastically. By the time you get to the surface of the star, the temperature is in the
range of only thousands or tens of thousands degrees, quite a bit cooler then the core.
The density is really low at the surface. Actually, the density is so low that it is lower
than the density of air molecules in the room that you are sitting in now - the gas is
really thinned out at the surface. Remember, in the core, densities are more than 10
times the density of lead! Stars are objects of very extreme conditions. See Figure 16
for some results from a computer model for the Sun.


Figure 16. Results from computer models of the Sun. These graphs show the calculated
temperature and density for a star like the Sun. The graphs are set up with radius increasing to
the right, and extend from the center (where the radius =0) to the surface (radius=1). The
temperature and density values vary over a wide range of values. Click on each image to see a
larger version. Data is from the solar model of J. Christian-Dalsgaard et al (1996).
I think that is enough talk about physics and computers. Let's get back to the Main
Sequence discussion. When a star finally turns on (starts fusion reactions), it is at the
159 | P a g e

starting point of its life on the Main Sequence. While on the MS, it is burning
hydrogen in its core - and nothing else. Initially stars are said to be on the Zero-age
MS, the point when they have just started to burn their fuel - brand new stars. Most
stars we see in the sky are not on the ZAMS, since they have generally been burning
fuel for some years. The ZAMS is really only a theoretical concept that is most useful
when we make computer models to study stellar evolution.
Figure 17. The Zero-Age Main
Sequence (or ZAMS) is shown - it is
the green line. This is only a
theoretical location on the H-R
diagram and is used to show the
likely location where a newly
formed star would be found before
it has burned any hydrogen. As
stars undergo fusion they gradually
follow the paths indicated by the
light blue lines until they reach the
end of their Main Sequence phase -
indicated by the dark blue line. The
masses of the individual stars are
indicated. This graph is based on
computer model data and is not
based upon actual stars.
How long will a star be a MS star? A looooooonnnnngggg time! Stars spend 90% of
their life on the MS. Boy, is it dull! Being on the MS is sort of like middle age. You
get up, go to work, come home, watch tv, go to bed, repeat. This type of routine is
boring! Except, a star never stops working; it is constantly fusing hydrogen into
helium (and making energy) while it is on the Main Sequence. Main Sequence life
spans for a variety of stars are given in the table below.

It is easy to figure out how long stars will last on the Main Sequence, since it depends
on how much fuel they have (mass) and how quickly they use it (luminosity). The
time a star spends on the Main Sequence can be approximated using the following
formula
time on MS = (1/M
2.5
)x 10 billion years
where M is the mass of the star in units of solar masses. If you put the Sun's mass in
there - you'd get the time the Sun will spend on the Main Sequence, about 10 billion
years. If you put in greater masses, the fact that the mass is being divided into10
billion will always give you a shorter time. For example, a 5 solar mass star would
160 | P a g e

give you an age of 1/5
2.5
x 10 billion = 1/55.9 x 10 billion = 10 billion/55.9 = 179
million. This formula is only an approximation, but you can see that a greater mass
means a shorter Main Sequence life. Even though high mass stars have more mass,
they burn it much more quickly and end up having very short lives (remember the
mass-luminosity relation of MS stars? Massive stars have very high luminosities and
use their fuel up very quickly).
Stars on the MS are arranged in very orderly way -
At the top left end of the MS- High Temperature, High Luminosity, High Mass,
Short Life (millions of years)
In the middle of the MS - Medium Temperature, Medium Luminosity, Medium
Mass, Longer Life (billions of years) - this is where you'd find the Sun
At the lower right end of the MS - Low Temperature, Low Luminosity, Low
Mass, Long Life (10s or 100s of billions of years).
The entire history of a star, how quickly it forms, where it resides on the MS, its life
span on the MS, the mechanism it uses to produce energy, its internal structure and
what happens to it after it leaves the MS is dependent primarily on MASS! Pretty
much everything about a star's life depends on its MASS. Don't forget that! The table
below gives some values of time that a star will spend on the Main Sequence based
upon computer models. You would not get these exact ages using the formula given
above, since that is sort of an approximation for all stars, so it isn't too good in some
cases.
Mass Spectral Type Years on Main Sequence
25 O7 6.4 Million
20 O9 8.1 Million
15 B0 11.6 Million
12 B1 16.0 Million
9 B2 26.4 Million
7 B4 43.1 Million
5 B6 94.3 Million
3 A0 351.7 Million
2 A5 1.1 Billion
1.5 F2 2.7 Billion
1 G2 9.4 Billion
0.8 K0 22.8 Billion


Now that you've read this section, you should be able to answer these questions....
161 | P a g e

What is needed for large scale star formation?
What types of stars (mass, spectral type) are formed in large scale star
formation?
How are H II regions formed?
How does small scale star formation differ from large scale star formation
(apart from the size)?
What are TT Tauri stars and H-H Objects?
How does the mass of a star influence the pace of its formation?
What do Main Sequence stars do that makes them Main Sequence stars?
What happens during fusion in the Sun?
How does the energy from center of the Sun get to the surface?
What are neutrinos?
How are astronomers able to understand the workings of the interiors of stars?
What does the Main Sequence life span depend upon?
How do characteristics of stars vary along the extent of the Main Sequence?
Stellar Evolution

What's covered here:
What causes stars to eventually "die"?
What happens when a star like the Sun starts to "die"?
What happens to the Earth when the Sun starts to "die"?
What is the final fate for the Sun and other low mass stars?
What is a nova?

Stars spend most of their lives on the Main Sequence with fusion in the core providing
the energy they need to sustain their structure. There is a price for this. As a star burns
hydrogen (H) into helium (He), the internal chemical composition changes and this
affects the structure and physical appearance of the star. The older the star, the greater
the amount of helium in the core.
The Sun is currently not on the ZAMS, since it has been burning hydrogen into
helium for about 5 billion years. This is one of the reasons that the MS appears as a
wide strip when it is plotted up. Most of the stars have been doing fusion for some
time, and therefore altering their internal structure, such that they are removed from
the ZAMS (since the internal changes effect their appearance - their luminosity and
surface temperature).
162 | P a g e

Figure 1. The core of the Sun (and other stars) over time. The top shows
how it started out, with 70% H and 27% He. Over time the size of the
helium core increases, so that it gets larger - as shown in (b), and larger -
as shown in (c). Remember, this is just the core of the star, while the rest
of the star keeps pretty much the same composition while on the Main
Sequence.
The Sun and most other stars originally have a composition of 70%
hydrogen and about 27% helium in their cores. This is sort of the standard
composition of stars like the Sun, at least when they start out their lives.
This is also the current composition of the layers outside of the core since
there is no fusion going on out there, but in the core things have changed.
Now if you were to look inside the Sun's core today, you would see that
there is about 35% hydrogen and 62% helium. Helium is denser, so it
sinks to the center of the core. What does it do there? Nothing but take up space.
Actually, you can think of helium like a lazy roommate - just sitting around in the
middle of the house, doing nothing and getting bigger each day. Actually, that would
be a pretty gross roommate, but you get the idea.
The helium core is getting larger every day, since the helium that is produced in the
fusion process is just taking up more and more space, since it can't really do anything
else. If you think this is bad, you're right. The helium, just by its presence, is sort of
crowding out the area of energy production. Remember, fusion can only occur in the
hot, high density area of the core. Outside of this area fusion (and energy production)
will not be happening. With helium taking up more and more space, there is less space
to go into the production of energy. In a way, the burning region is being forced
outward away from the star's center, it's getting crowded out by the helium.
What is it like as you get further from the core? It's cooler and less dense (remember
those temperature and density graphs from the previous set of notes?). This is a region
that is not hot enough to sustain the same rate of energy generation as in the hot, dense
core. Is that important? Of course it is important - the energy from fusion helps to
hold up the outer layers of the Sun and maintains the various forms of stability within
the star (like hydrostatic equilibrium and that other stuff).
What happens when the upper layers are not being held up as effectively as before?
Gravity raises its ugly head, and boy does it have one ugly head. You have to
remember that gravity is always there, but when you are not fighting its influence very
effectively, you pay the price! The layers outside of the core would start to pull
inward, and the Sun's core and the area around it would slightly contract. While this
might not seem like a good thing, it actually is, since the contraction will help to heat
up the area around the core and get the temperature and density to go up to a level
163 | P a g e

where fusion can start in regions that were previously too cool or too low density for
fusion to operate.
The basic upshot of this whole thing is that the burning region (energy production
region) of a star is gradually moving further out from the center of the star as the
fusion by-product (helium) is taking up more and more space in the center.
Does this mean that the Sun is getting larger? No, because you have to remember that
the mass of the Sun has been going down steadily over the years, since all the energy
that it gives off as sunlight was actually mass at one time. The Sun is losing mass
steadily by just giving off light - a lot easier than Weight Watchers, isn't it? Helium is
denser than hydrogen, so the core is actually getting slowly denser with more and
more helium being produced.
The slow squeezing process seems to have solved the energy problems of the Sun,
right? A little more helium causes the energy production region to be slowly pushed
out from the center - so everything is just fine, right? Not really, since the Sun (and
stars like it) can only do this for so long. Eventually it will get to the point where the
contractions will not be able to heat up the interior regions high enough to enable
them to produce energy to sustain hydrostatic equilibrium. Even though there is
gravity keeping things hot and dense, it won't be enough to help the situation. There is
a limit to how tightly you can squeeze stuff and how hot you can get the material.
At this point hydrogen burning in the core is no longer significant, and there is only a
thin shell of hydrogen burning around the large helium core. The star is pretty much at
the end of its Main Sequence life.
Figure 2. The outer layers of a star (like the Sun)
get pushed out by the compression and heating of
the core. Even though there isn't much fusion going
on, the gravitational heating of the core causes the
heat flow to increase and the outer layers of the
star to swell up.
The shell burning of Hydrogen doesn't produce a
lot of energy, so it isn't really helping in holding up
the outer layers of the star. Yet the outer layers are
being held up - how? It is pretty much due to all of
the hot gas in there, or Thermal Energy, which is
just the heat produced by the squeezing of the core.
Remember the influence of gravity - it will compress and squeeze the core so that the
core gets smaller and hotter. The heat of the core is so great that it will start to actually
164 | P a g e

overdo the "holding up the outer layers" bit. It will, in fact, sort of puff out the outer
layers of the star. As the outer layers spread out, they will cool off and the observed
surface temperature of the star will decrease. What do we have here? We get a Red
Giant - better start a new section.
Red Giants
Finally we get to the red giant stage; I know you have been waiting for this for some
time, so just try to stay calm. First, some clarification: I know that red giants are a bit
confusing since there seem to be two opposite things happening at the same time - the
core is getting smaller and hotter due to compression, while the outer layers are
getting more spread out and cooler. You sort of have things going in two opposite
directions - but these things are related. Without the heating of the core, there would
be no expansion of the outer layers.
In the case of the Sun, it will expand to a size greater than that of Mars's orbit, or
become about 430 times larger than its current size, and will have a surface
temperature of 3500 K and a luminosity that is 20,000 times its current value. Of
course, the core temperature will be increasing (remember, the squeezing is still going
on), getting closer to around 100 Million K. The size of the Sun will be as big as the
size of Mars's orbit? What will happen to the Earth? - nothing good, obviously. The
Sun will get larger and larger each day, and the luminosity will increase, so the
surface temperature of the Earth will rise (this is a really big time global warming
event). Eventually, after all life on Earth is destroyed by the intense heat and
radiation, the planet will enter the expanding outer layers of the Sun. If it doesn't burn
up right away, it will probably spiral in toward the center of the Sun, so it will
eventually burn up. I guess not even SPF 40 will help at that point. Not only will this
happen to the Earth, but Mercury, Venus and Mars will most likely share this fate.
Jupiter and Saturn may be the safe places to be at this time, so you better buy some
real estate out there now while the prices are still low.
The core is being compressed continually, getting hotter and hotter as well as denser
and denser. Is there no limit? Will it ever end? Yes, it will end; there is a limit. The
limit comes about due to some of the laws of quantum mechanics. One of these laws
tell us how tightly we can pack things in, such as atoms and electrons. Once a star's
core gets to the point where the electrons are packed in as tightly as possible, the
material is said to beelectron degenerate. The core of a red giant is therefore
becoming more and more electron degenerate. One of the unusual properties of
electron degenerate material is that once it is electron degenerate, you can't make it
any denser. No matter how hard you squeeze and compress it, it will not get any
denser - it will get hotter, but not denser.
165 | P a g e

Another thing that happens with a red giant is that the outer layers become very
convective. Actually, there are huge convective bubbles reaching down all the way to
the core and then back up to the surface. For a star that is as large as the size of Mars's
orbit, these are some seriously large bubbles! One neat aspect of these huge
convection bubbles is that they can mix up material really nicely. Sometimes they can
even pull some stuff from the core onto the surface. Red giants can have rather
peculiar chemical compositions in their surface layers, since the high density stuff in
the core can get mixed in with the material in the outer layers, which we can see when
we obtain a spectrum. This is something that astronomers have seen with red giants,
and is a feature that is predicted by our computer programs as well.
It's time to get back to the core. When the core temperature reaches 100 million K, it
is hot enough for helium fusion to occur. Believe it or not, this is a good thing.
Actually, it isn't so good at first, since the core of the star is electron degenerate when
the helium ignites, and the resulting ignition is a rather catastrophic event.
This Helium Flash marks the rapid onset of helium burning in the core for a low mass
star. A helium flash is a rather violent explosion that can significantly alter the
internal structure of a star, in part because the material in the core is so electron
degenerate. We can't actually see the helium flash, since it is buried deep in the core,
and it happens relatively quickly, but it does get the star back on the fusion track.
Stars more massive than the Sun usually don't have an electron degenerate core, so
when their helium ignites it's no big deal. In the case of the low mass stars, the helium
flash also has the added bonus of removing the electron degeneracy in the core.
Now we have sort of a happy star - it has gotten its second wind. The helium fusion is
making a good amount of energy in the core and the core is no longer electron
degenerate. Unfortunately, the helium burning is not as energy efficient as hydrogen
burning, so you get less bang for the buck. You might want to think of it as a lower
quality fuel, perhaps a lower octane than hydrogen fusion. The basic upshot is that not
as much energy is produced in each helium fusion reaction as was produced in each
hydrogen fusion reaction, so the star has to burn the helium faster to produce a
sufficient amount of energy. The core is still extremely hot, so it is producing a lot of
thermal energy as well, which will keep the outer layers puffed up, so the star is still a
cool red giant.
Figure 3. The interior of a red giant after helium
fusion starts. The size of the star is on the order of
a couple of hundred A.U.s, while the core is only
about as large as the Earth. The close up for the
core shows that multiple layers exist. At the center
are the by-products of the helium fusion - carbon
and oxygen. Above that there is a layer of helium
166 | P a g e

that is undergoing fusion and another layer of helium that isn't fusing (too cool or low
dense). Above the helium layers is the layer of hydrogen fusion in a very small shell.
This layer marks the outer edge of the core. The rest of the star is for the most part the
same composition that the star had to begin with (mainly hydrogen and helium).
Now helium fuses into carbon (C) and oxygen (O) through what is known as
the triple alpha process. This is sort of a silly sounding name, but it comes from the
old fashioned name for helium (alpha particle), and when you combine three helium
atoms together, you make carbon (as well as energy), while four helium atoms will
combine to make oxygen (and energy). I guess it could be called the quadruple alpha
process, but that isn't as easy to pronounce. Any ways, let's get back to the fusion of
helium. Helium fusion is less efficient than hydrogen fusion, so it goes pretty fast and
the carbon and oxygen will start to take up more and more space in the core of the
star. Does this sound familiar? It should. Again, the burning will move out further and
further away from the center of the star and the core will start to get degenerate again.
As new regions of helium start to burn, the star can experience Helium Shell
Flashes (also called thermal pulses). These are similar to the explosive helium flash,
but less powerful since less material is involved and it isn't as degenerate as the core
was when the helium flash occurred. These helium shell flashes are rather
troublesome, though, since they do release a fair amount of energy - kind of like
violent, uncontrollable hiccups.
After about 1 billion years (for stars similar to the Sun), the build up of carbon and
oxygen in the core will prevent the production of a significant amount of energy, since
the darn carbon and oxygen are taking up so much space. This marks the end of the
line for the star. The core will start contracting, but the contraction will not produce
high enough temperatures or densities to ignite the next fusion process. A star will
spend only about 10% of its life as a red giant - not much for a second lifetime, but
you take what you can get.
Is the end of helium burning the end for all stars? No, stars more massive than the Sun
may be able to continue into the next burning stages once the helium fusion stage
stops. Remember, more mass leads to more gravitational pull (or contraction), which
leads to more squeezing of the core. More squeezing leads to higher temperatures and
densities. For these larger mass stars carbon would burn next, and then heavier
elements after carbon, but only if the star is massive enough to raise the temperature
and density to the high levels required for the fusion of these heavier elements. The
star will remain a red giant and basically just hang around in the upper right area of
the H-R diagram until it finishes all of its fusion stages. Different mass stars tend to be
arranged with the higher mass stars further up on the H-R diagram, but that's not
always true - the arrangement of the stars in the red giant stage isn't as nice and
167 | P a g e

orderly as it was during the Main Sequence stage. It is rather difficult to precisely
determine the mass of a red giant star.
Figure 4. The various steps of the
Sun's evolution are outlined here.
This is also the way that most
other low mass stars similar in
mass to the Sun will evolve. The
current location of the Sun is
indicated by the yellow dot along
the evolutionary path. The
direction of evolution is indicated
by the arrows.
Before we find out how a star like
the Sun ends its existence, let's
recap the fusion history of a low
mass star like the Sun. The
various points listed below are
marked in Figure 4.
1. Hydrogen is undergoing fusion in the core, which produces helium and energy.
The star is happily doing this for 90% of its life, which is the amount of time it
spends on the Main Sequence. This amounts to a total MS life of about 10
billion years for the Sun.
2. Hydrogen fusion in the core ceases for the most part, with only hydrogen fusion
in a thin shell still occurring. The star from this point on is a red giant. While
the outer layers get puffed up, the compression in the center starts to form an
electron degenerate core.
3. Hydrogen shell burning continues in a small layer around the degenerate core.
The fluffy red giant is rather unstable, and the outer layers get very convective.
This can bring up some of the heavy elements from the center to the surface,
which produce some rather bizarre spectra.
4. Helium Flash signals the start of helium fusion in the core.
5. Helium core fusion producing energy and the byproducts of carbon and oxygen
begins. There is also a thin layer of hydrogen shell burning around the helium
core, though this isn't producing as much energy as the helium fusion.
6. Helium core burning decreases as the carbon and oxygen build up so there is
only helium and hydrogen burning in thin shells. This is also the time when
helium shell flashes would occur, causing disruptions in the structure of the
star.
168 | P a g e

7. What comes next? That depends upon mass, with higher mass stars continuing
on with further stages of fusion - maybe one more fusion stage, maybe two, or
maybe three. It all depends on the mass of the star - more mass, more fusion.
For a star like the Sun, after step 7, it is pretty much dead. Follow this link to see the
evolutionary paths of several other stars as they evolve from the Main Sequence
through their helium fusion phases. But what happens after the helium fusion? How
does a star like the Sun actually die? I'll get to that in a moment, but first, a small
diversion.
Brown Dwarfs
Believe it or not, there are some stars out there that are more boring than the Sun. This
would be those that have very low masses compared to the Sun. How low? Low
enough so that they don't even get to do any helium fusion - these stars only have a
hydrogen fusion stage and after that, they're pretty much dead. This is likely the fate
of stars less than 20% of the mass of the Sun - likely those that are Main Sequence
types M5 and cooler.
Figure 5. The coolest brown dwarf yet
discovered, WISE 1828+2650, with a
surface temperature of only about 300
K, (80 Fahrenheit). The colors are not
true color - click on the image to see
the larger view. Image credit:
NASA/JPL-Caltech/UCLA.
What's the limit for hydrogen fusion?
The lowest mass that a star can have
and be able to fuse hydrogen is about
8% the mass of the Sun. Remember
that when stars form, the vast majority of those that form are very low mass, so there
should be a lot of these very small objects out there. What do we call these really low
mass not-quite-stars? These are known as Brown Dwarfs. These stars have surface
temperatures of only about 1000 K and less, and have a fraction of the Sun's
luminosity. Such objects would be very difficult to detect, and seem to have the
characteristics of the L and T type stars. So are L stars really brown dwarfs? No
necessarily. Some L stars are large enough to fuse hydrogen and therefore aren't
brown dwarfs, while other L stars are too small to fuse hydrogen and should be
viewed as brown dwarfs. It is more likely that all of the T types are brown dwarfs. It is
really hard to determine which of these stars are fusing hydrogen and which aren't
since they give off very little light and it is very hard to see them and accurately
169 | P a g e

measure their characteristics. We have to use IR telescopes like Spitzer andWISE to
detect them. In fact, the Spitzer has found not only brown dwarfs but also disks of dust
around them - which sort of stand out since thees are quite large disks. So far the
coolest (and perhaps the smallest) stars are classified as T9 types, with temperatures
down to about 500-600 K. Astronomers have also suggested that there be another
spectral type beyond T stars, and these cool objects would be known as "Y" type
brown dwarfs. Unfortunately so few Y type stars have been found that it is difficult to
understand their characteristics and also to determine if they should actually be
planets and not stars.
While it is very difficult to just find a brown dwarf, it is also very difficult to measure
their masses. This was recently done, where astronomers were able to measure the
masses of two brown dwarfs and found masses of only about 3% the mass of the Sun.
Brown dwarfs may also play a role in the structure of our galaxy since they should be
very common and could contribute substantially to the overall mass of the galaxy. For
stellar evolution, Brown Dwarfs are pretty much a dead end, since they have no
fusion, they don't change or evolve in any way and just cool down over time, which is
pretty boring.
Since we want to see things that are more exciting than brown dwarfs, we need to
return to the discussion of a star like the Sun to see what happens after it finishes all of
its fusion stages.
Stellar Death
The life stages of a star will depend mainly on one factor - its mass. How long it lives,
how it will die, what it will fuse in its core - all of these depend upon the mass.
Though different mass stars will evolve in slightly different ways, many will do the
same sort of general things so that we can group them together. We'll first look at the
case of the small stars like the Sun.
Low Mass Death (for stars like the Sun)
Eventually the red giant phase will end when the star can no longer burn anything in
its core. For the Sun, the last element it will burn is helium. Other, more massive stars
will burn carbon, oxygen or neon in their cores and other heavy elements if they are
massive enough. Of course, each time a star burns an element it produces heavier
elements as by-products. Eventually they will have to stop burning since their cores
can't get hot or dense enough for the next burning cycle to start.
When a star stops fusing an element it evolves towards the upper right corner of the
H-R diagram. This is because of the compression of the core and the resulting
170 | P a g e

expansion of the outer layers. The star is getting more and more puffed up. This is
also a time when the convection in the outer layers is very extreme, so that the entire
star is almost completely convective. If that weren't bad enough, in low mass stars like
the Sun, the Helium Shell Flashes will keep occurring every once in a while and these
will release a lot of energy in short bursts. This is rather like having a very nasty case
of the hiccups - with the outer layers of the star actually being pushed outward.
Another aspect of stars in this stage of their lives is the development of very strong
stellar winds (really super winds). With the stretched out surface layers, the Helium
Shell Flashes, and the super winds, there is so much going on to push the outer layers
even further out. The combined energy of all these processes can help to blow off the
outer layers of the star. Actually, this is not too difficult to do since the outer layers of
the star are already stretched out quite a ways from the core (remember, it is a red
GIANT) and so it is harder for the star to hold onto them effectively (the further from
the center, the lower the gravity).
The end result is that the outer layers get blown off. As the outer layers get blown off,
there will be a lot of hot gas flowing away from the star, which used to be the outer
layers and there will only be the core of the star left behind. Sometimes the material
will be ejected as a ring, sometimes the material gets blown off in two directions
(remember bipolar outflow); it is also possible that the material will be blown out in
bubbles. The animations provided are based upon various theories of how material is
ejected. In the case of the bipolar outflow, it is believed that the interactions of two
stars cause the funnel shaped outflow. In any event a large amount of material
(basically all of the stuff outside of the core) gets blown out into space. This produces
a structure known as a Planetary Nebula - and in the great tradition of naming things
in astronomy, it has nothing to do with planets! It's just that in the old days when
people looked at these things with telescopes, they saw that most Planetary Nebulae
had disk shapes, and the only other things visible with such shapes were planets. Yes,
that's a pretty lame reason, but we're stuck with it.
It's time to get back to killing off stars. During the rather nasty Planetary Nebula
stage, stars like the Sun will lose about 40% or more of their mass. More massive stars
will lose an even greater fraction of their mass. Some Planetary Nebulae are not nice
and perfectly spherical - some have rather unusual shapes, possibly due to irregular
mass loss or several different phases of mass loss. Typically Planetary Nebulae are a
few light-years in size and the gas is still moving away from the star at relatively high
velocities (a few thousand km/s). The gas in a Planetary Nebula will stay rather hot
due to the heat from the hot, dense core, so they are visible for many years after the
start of this phase.
171 | P a g e



Figure 6. Examples of Planetary
Nebulae. All images are from the
Hubble Space Telescope. The top row
shows the common bipolar outflow
shaped nebulae, where material is
ejected in two directions (assumed to be
out the poles). The different shapes may
be due to how the nebulae are tilted
toward or away from us, or how well
aligned the ejected material is. The two
images at left are of the circular variety
of the planetary nebulae. Initially it was
thought that these circular shapes were
due to a circular or spherical ejection of
material, though some now think that
the circular shape could be due to a
end-on alignment of the bipolar
types. Image credits: NASA, ESA, The
Hubble Heritage Team (STScI/AURA),
Bruce Balick (University of
Washington), Vincent Icke (Leiden
University, The Netherlands), Garrelt
Mellema (Stockholm University), R.
Sahai & J. Trauger (JPL).
Over time the ouflow of gas from the star will stop and the gas in the Planetary
Nebula will cool down and get lost amongst the rest of the gas floating around in
space. A nifty thing about this material is that it is often enriched with heavy elements
- remember, this phase comes after a star has finished doing all of those various fusion
processes, so that there is an excess of heavy elements in the material. This is one way
that heavy elements (stuff other than hydrogen and helium) can get deposited into
space. This is an important point that we'll bring up again later.
172 | P a g e

All that is left of the star after the Planetary Nebula stage is the hot little core. What is
the core doing? It isn't burning, but it still has to deal with the influence of gravity,
which will compress the core down to the point where the material is again electron
degenerate. Once it gets to that point, the compression will stop. The star is now a hot,
dense but stable object. That's pretty boring, but also pretty bizarre in a way - it's still
electron degenerate, which makes it rather abnormal.
The core that is left over will have a mass that is about 1/2 the mass of the Sun, but it
will have a radius comparable to the that of the Earth (around 6000 km). The surface
temperature will be up to 100,000 K, but it will cool off in time. The density of this
object is about 1 million grams per cubic cm. This is about the same as the density of
a Volkswagen - not a regular Volkswagen, but one that has been crushed down to the
size of a sugar cube. Pretty dense, eh?
Figure 7. A typical white dwarf (right)
compared to the Earth (left). Even though it
has a radius similar to the Earth, the mass of
a white dwarf is much closer to that of the
Sun. This makes it a very dense object. Earth
image courtesy of NASA.
What do we call this hot, dense, small sized
object? A White Dwarf - actually, this name kind of makes sense; it's small and it's
hot - wow, a name of something in astronomy that actually makes sense.
Now there is a really nifty thing about white dwarfs. They are electron degenerate
objects, so they don't obey the same laws of physics that normal stuff does. Someone
noted a rather interesting consequence of this feature. A young student who was
thinking about these things, Subrahmanya Chandrasekhar, came to a rather startling
realization - that if you added more mass to a white dwarf, it got smaller in size
(radius). More mass, more shrinking - this doesn't make sense, but this is the way
degenerate material acts. Chandrasekhar eventually figured out that if you added
enough mass, the white dwarf would be shrunk to a size of 0! Which means it could
not exist - you can't have things that have no size, after all (later we'll sort of break
this rule). Chandrasekhar determined that if an object is electron degenerate it cannot
have a mass greater than 1.4 solar masses; any bigger than this and it would shrink out
of existence - it just could not hold itself up anymore. All of the white dwarfs that we
know of are stable, so they must all be less than 1.4 solar masses. Actually, where we
can measure their masses, we always find that they are less than 1.4 solar masses. This
mass limit is known as the Chandrasekhar Limit. Perhaps the best way of thinking
of the Chandrasekhar limit is just like a wrestler trying to get under their weight limit,
173 | P a g e

however in the case of the wrestler, he would still exist if he were overweight, unlike
the white dwarf.
For a star to eventually end up being a white dwarf, it must get below the
Chandrasekhar limit (1.4 solar masses). Some astronomers think that stars that started
out their Main Sequence lives with about 8 times the Sun's mass can end up as white
dwarfs. For this to happen, they must lose at least 6.6 solar masses on the way to
being a white dwarf - that's sure a lot of material, and it can lose a large amount of
that material in the Planetary Nebula stage, though it can also lose it in other ways,
such as with a steady, strong stellar wind. Regardless of how it does it, it still has to
do it! If it doesn't, the star won't end up as a white dwarf, but something else will
happen (as you'll see later).
Figure 8. The entire evolution of a star
like the sun, from the Main Sequence,
through helium fusion and the
planetary nebula stage, down to the
white dwarf and ultimately black dwarf
stage. The line from the end of the
helium fusion stage to the planetary
nebula stage is not drawn in since this
path is not well known.
In Figure 8, the evolution of the Sun is
diagrammed from the ZAMS all the
way to the white dwarf stage. You may
notice that there is no solid line
between the end of the helium fusion
stage and the Planetary Nebula stage
since this part of the star's evolution is kind of difficult to plot. At the start of this
span, the star is a very cool red giant with a hot, compressed core. Then, when the
Planetary Nebula stage starts, it basically starts to peel off the cool outer layers,
revealing the hot core. In a way it sort of jumps from the far right side of the H-R
diagram all the way to the left side in a very short time.
When the Sun eventually goes through the Planetary Nebula stage, it should lose
about 0.4 solar masses and end up as a 0.6 solar mass white dwarf. That's not very
exciting, but what else can it do? - nothing. About the only thing that a white dwarf
can do is cool down. Eventually it will get cooler and less bright (remember how
luminosity depends on temperature) until it gets too cool to see. Once a white dwarf
completely cools down, it will become a black dwarf. However, that takes such a
long, long time that there are no black dwarfs currently in the Universe (the Universe
174 | P a g e

isn't old enough for any to exist). A rather nifty animation showing the size of a white
dwarf and its ultimate fate can be seen here.
What about stars more massive than the Sun? More mass means more gravity, more
gravity means more heat in the core, and more heat in the core means more fusion
cycles can occur. These stars have the chance to start burning other elements, like
carbon or oxygen. The more mass a star has, the more fusion cycles it can go through.
Remember, it does this very quickly - big stars use up their fuel rapidly. Even though
they have more fuel, they are not very economical in its usage. Even though they can
burn more stuff then the Sun, they are still not very careful in how they use it. Big
stars will eventually end up as white dwarfs but they tend to go through a few more
burning cycles. On the H-R diagram, when these stars are burning other elements,
they would just sort of wander around the area that red giants or red supergiants are
found in. They'll stay as red giants or supergiants so long as the fuels lasts, but when it
does run out, they'll end up down in the white dwarf location (after going through a
Planetary Nebula stage), just like the Sun.

Not all white dwarfs are the same - they can have very different compositions.
Remember, a white dwarf is the remains of the core, so whatever the core of the star is
made of ends up being what the white dwarf is made up of. For a very low mass star,
it will only go through the hydrogen fusion cycle where it produces helium. That star's
white dwarf should be made mainly of helium. A star like the Sun will burn helium,
producing carbon and oxygen. The Sun's white dwarf will be made mainly of carbon
and oxygen. There are white dwarfs out there made of various things like mixes of
oxygen, neon and magnesium, to name a few. The larger the star was on the main
sequence, the heavier the final fusion product will be - and this will determine the
final composition of the white dwarf. Now let's look at an interesting fate for some
white dwarfs.
Nova
If a white dwarf is sitting all by itself in space like the Sun's white dwarf will be, not
much else will happen with it. If it is a binary system, especially a close binary system
where things are really tight, it can get very interesting indeed.

It is possible for mass to transfer from one star to another in a close binary system.
This can happen in certain stages of the binary star system's evolution. One example
of what can happen is the following - let's say you have two stars of slightly different
mass. The more massive star will die first (remember, mass determines fate and big
mass means short life), the big star will go through its various stages of evolution and
end up as a white dwarf before the less massive star has a chance to even do much of
anything. In fact, the lower mass star will still be on the Main Sequence long after the
175 | P a g e

other star goes through its entire life. You now have a white dwarf star and a Main
Sequence star. Eventually the remaining MS star will start to die and will enter the
Red Giant stage - nothing unusual about that. As it puffs up it will get larger and
larger and will fill up a gravitational boundary, the Roche Lobe, around the binary
system.
Figure 9. The ingredients needed for a nova. Two stars, one
a white dwarf, the other a star becoming a red giant, are in
orbit about one another. The Roche Lobe marks the
gravitational limit for each star. As the red giant expands,
the material in its outer layers don't expand out in any
direction due to the proximity of the nearby white dwarf, so
the material is funneled toward it. This is due to the
constraints of the Roche Lobes. Eventually the material on
the white dwarf will ignite as a nova.
The Roche Lobe isn't a physical barrier like a wall, but it
just defines how the gravitational pull of the binary system
causes material to move in certain ways. As the red giant
fills up its side of the Roche Lobe, the material will not just
expand outwards, but will instead be funneled towards the
other star - that's all due to our electron degenerate friend's
high gravity. Now the material from the red giant is being
transferred over to the other star in the system, which in this
case is a white dwarf. This cannot be good! The material
transferring from the red giant will not just dump directly
onto the white dwarf since the whole system is moving and things are going around in
orbits, so the material sort of spirals in. It will gradually build up a disk around the
white dwarf. The material builds up around the white dwarf into what is called
an accretion disk. The white dwarf is rather hot and the material gets heated up in the
process of spiraling in, so it will tend to give off ultraviolet (UV) radiation. This is
good, since it provides a way for astronomers to identify such binary systems.
Remember, this is in a binary system and the only thing that might be visible in a
regular telescope is the red giant, which is too cool to produce any large quantities of
UV radiation. The presence of the large amounts of UV radiation points to an unseen
accretion disk with a white dwarf buried inside it. Gradually material will get to the
white dwarf and build up on its surface. What type of material is it? It is what the
outer layers of stars are made up of, which is the regular mix of mainly hydrogen with
some helium in it.
176 | P a g e

Figure 10. The set up for a nova - one
star getting material ripped off, another
star (white dwarf) pulling the material in.
The material forms an accretion disk
which will heat up to the point where it
gives off UV light.
Now we have a white dwarf getting
dumped on. After some time (years, or
perhaps even decades), the material that
accumulates on the surface of the white dwarf will ignite in an explosive blast. This
explosion has a brightness about 100,000 times greater than the Sun's luminosity.
What you have here is a Nova. Novae (that's the plural form) will stay bright for
weeks, but are not as bright as supernovae (which we'll get to later). While the
explosion is pretty powerful, energetic and bright, it doesn't destroy the stars involved.
The nova is produced by mass being transferred in a binary system, so it is possible
for it to happen again and again in the same star system. This is known as a recurrent
novae. There are 10 known recurrent novae in our galaxy. The current record holder
for most frequent eruptions is U Scorpii (in the constellation of Scorpius). This system
went nova in 1863, 1906, 1917, 1936, 1945, 1969, 1979, 1987, 1999 and 2010.
Click here to see the light variation of U Scorpii during the 2010 nova. I guess a
recurrent nova is sort of like a repeat offender - they just can't help themselves and
keep doing the same thing over and over. In the case of U Scorpii, even though it has
gone off at least 10 times, it is possible that it could go off again!
Figure 11. Nova Cygni seen at two different
times, 1992 and 1993 (left and right
respectively). The expansion of the blast shell
over time is obvious in the two images. The two
stars that are the source of the nova appear
like one star in the middle. Image credit: F.
Paresce, R. Jedrzejewski (STScI) NASA/ESA.
Novae are so bright that they can at times be seen in other galaxies millions of light-
years away. Eventually the explosion will fade away, though it did cause a nice light
show while it lasted. There can be dozens or hundreds of novae occurring in a galaxy
each year, since there are so many white dwarfs out there. A system that produces a
nova will not be able to repeat the process forever, since eventually the red giant will
become a white dwarf, so you'll end up having a system containing two white dwarfs.
What will these white dwarfs do? - nothing but cool off slowly, eventually becoming
a pair of black dwarfs, but that takes a very long time. They end up being pretty dull
and boring in the end.
177 | P a g e


The Sun will not become a nova (remember, it takes two stars to have a nova), so it
will have a pretty boring end. Sorry, folks, this is pretty much the end of the line for
low mass stars like the Sun.

Now that you've read this section, you should be able to answer these questions....
How does the interior of the Sun change over time? What effect does this have
on the rest of the Sun?
What happens when the hydrogen runs out?
What happens to the planets when the Sun goes through its Red Giant phase?
What are the characteristics of a Red Giant?
What causes helium fusion to start?
What happens once helium fusion ceases?
What happens during the Planetary Nebula stage?
What are the characteristics of a White Dwarf?
What is the Chandrasekhar limit?
How can the evolution of a star like the Sun be depicted on an H-R Diagram?
What is needed to produce a nova?
Large Mass Stellar Death

What's covered here:
How do massive stars end their lives?
What happens when a star becomes a supernova?
What are the characteristics of a supernova?
What are some famous supernovae of the past?

The fates of large mass stars are quite different from those of the low mass ones. At
first glance you might think that with more mass they will live longer - but no, they
are just fuel guzzlers. As such, massive stars (O and B types on the Main Sequence)
will have very short lives. To get a good idea of how a large mass star dies, let's look
at what a really big one goes through.
Very massive stars are also very luminous, so they tend to have very strong stellar
winds (since this is linked to their energy outputs). Due to this they may have a lot of
gas around them that was blown off long ago. Sometimes it is easy to see these gas
178 | P a g e

clouds, sometimes not. Like other mass ejection things we have already seen, the
material can be spewed out in several ways, with the most common being bipolar
outflow and rings. Figure 1 shows several different stages in the life cycle of stars, all
in one convenient location.
Figure 1. A picture from the Hubble
Space telescope showing stars in various
stages of their lives. The lower right
region (1) is a gas cloud from which stars
form. Near the center (2) is a group of
main sequence stars. Up and to the left of
those is a large star (3) in the process of
dying. The massive star has already
ejected a ring of material as well as
material out from it in a bipolar direction.
In a way this image is just like a family
portrait showing the oldest and youngest
members of a family. Image credit:
Wolfgang Brandner (JPL/IPAC), Eva K.
Grebel (Univ. Washington), You-Hua Chu
(Univ. Illinois Urbana-Champaign), and
NASA.
Sometimes the outflow is even more energetic and at times confusing to astronomers.
Some very massive O-type stars have such strong winds that they can completely
screw up their evolutions - they lose so much mass that you have to take that into
account in your calculations when trying to figure out how these stars will live their
lives. Due to this uncertainty, we have a rather hard time predicting what a very
massive star will do or even trying to figure out what one of them may have done in
the past. Images of two such objects are shown in Figure 2. A recently observed
massive star (with the exciting name of WOH G64) has lost so much material that a
thick, dusty ring has formed around it. This star is very far away, in another galaxy in
fact, so it is not possible to easily see the ring, but its presence is revealed by spectra
and other instruments. You can learn about this massive star here. Even though it has
lost a bunch of material, WOH G64 is still about 1500 times wider than the Sun.
179 | P a g e


Figure 2. Two very massive stars with very extreme mass loss episodes. On the left is the star
WR124, a massive star ejecting material out at speeds of 100,000 miles/hr. Each blob of gas that
it ejects out has a mass of more than 30 times that of the Earth's. To the right is the unusual star
Eta Carina. Actually, the star is buried in the center of the two bubbles, which are thought to be
from an eruption that occurred in 1847. Follow this link to see how the bubbles formed. During
this outburst the two bubbles were ejected as well as a disk of material that can be seen between
the bubbles. The speed of the material in this case is on the order of 1.5 million miles/hr. Image
credits: Yves Grosdidier (University of Montreal and Observatoire de Strasbourg), Anthony
Moffat (Universitie de Montreal), Gilles Joncas (Universite Laval), Agnes Acker (Observatoire
de Strasbourg), Jon Morse (University of Colorado), and NASA.
All of these big time mass loss episodes are just a sneak preview of what is to
eventually happen to these stars. These stars are more massive than the cut off for
going into Planetary Nebula stages, so the death scenarios of the massive stars go
down very different paths. Just how big can a star be? We're not exactly sure, but
there is a limit to how much material can come together to form a single star, and
conservative estimates put that at around 150 solar masses. However in 2010
astronomers from the Very Large Telescope in Chile announced the discovery of a
star that may have a mass that is currently 265 times the mass of the Sun. What is
really amazing about this star, with the cute name R136a1, is that it started out with a
mass of about 320 solar masses! So it probably had some serious mass loss episodes
in its life. In case you were wondering, this star is not even in our galaxy, but is in a
neighboring galaxy.
We'll look at the life cycle of a 25 Solar Mass star to see what happens to one of
these big beasts. It is more massive, so it can go through more burning stages than a
low mass star. It can ignite the more massive elements due to the greater gravitational
180 | P a g e

heating in the core (more mass means more gravity). Each burning stage takes less
time. The data in the accompanying table is from Chieffi, Limongi, and Straniero
(1998) based upon their computer model for such a star.
Fusion Process Main Fusion Products Duration of Fusion Process
H He 6 million years
He C, O 700,000 years
C Ne, O 1000 years
Ne O 9 Months
O S, Si, Ar 4 Months
Si Fe, Cr 1 day
When the star is burning hydrogen, it is on the Main Sequence. Every other burning
stage after that has the star off the Main sequence and in the area of the H-R diagram
populated by Supergiants. The stars just evolve through these various burning stages
while they are Supergiants, sometimes blue supergiants, sometimes yellow
supergiants, and sometimes red supergiants, so they will wander back and forth across
the top of the H-R diagram during these later stages. If you click here you can see
some evolutionary paths of such stars. The region that the supergiants inhabit isn't a
clearly defined location on the H-R diagram, but is basically just the top part where
the luminosity is very high.
Take a look at the numbers in the rightmost column. Why is each stage shorter than
the previous stage? There are two reasons
There is less fuel available in each stage. After all, stars are originally made up
of mainly hydrogen and helium, so all of the other material comes for the most
part from previous burning stages. Only a fraction of the original star is made
up of the elements heavier than hydrogen and helium that are used in the later
burning stages.
There is lower efficiency in the later burning processes. Less energy is released
when heavier elements are undergoing fusion, so the star needs to burn that
material at a greater rate to produce enough energy to sustain the structure of
the star (to maintain stuff like Hydrostatic and Thermal Equilibrium).
181 | P a g e



Figure 2.The structure of the
interior of a supergiant star when it
finishes silicon fusion. The very
center has the inert iron core, and
it is surrounded by thin layers of
fusion shells. The core size is
actually exaggerated a bit here. It
is actually only 1/1000 the radius
of the star (about the radius of our
Sun), but it contains about 1/3 of
the star's mass (around 8 solar
masses for a 25 solar mass star).
Note how the radius of the star
stretches out to the size of Jupiter's
orbit. Obviously we wouldn't want
a star like this in our
neighborhood.
Figure 3. Betelgeuse - a truly giant supergiant. The
shoulder star of Orion is so large that its size can actually
be seen in the Hubble Space Telescope (most stars are so
far away that they always look like dots unless they have a
really huge diameter, like this one). Image credit: Andrea
Dupree (Harvard-Smithsonian CfA), Ronald Gilliland
(STScI), NASA and ESA.
So now these dying massive stars will be seen as either Red Supergiants or Blue
Supergiants, depending upon how hot or cool they are. Due to their very large radius
they also tend to be extremely luminous. How large can the radii get? In 2005,
astronomers discovered several red supergiants with radii that were much larger than
that of Betelgeuse (shown above) - these stars had radii that are about 1500 times that
of the Sun! To see how big these stars actually are, just take a look here. Obviously
with the outer layers so stretched out, the fusion going on deep in the core is not going
to be visible to anyone. Even though stars like this look like they are in the last stages
182 | P a g e

of their lives, we can only see what is going on at the surface, not what is actually
happening in the core. Of course the core is the interesting part!
Eventually the core of the massive star will look like a giant onion, with the densest
material in the middle and the lowest density stuff on the top. Each shell of the onion
will have some small amount of fusion still going on, but the energy that is being
produced at this point is pretty pathetic.
All sorts of elements have been burning, and now we come to the last element to burn,
iron (Fe). Does it burn? It sure does, but at a cost. Whereas previous fusion processes
released energy, iron burning consumes energy. The energy that should go into
holding up the star instead goes into burning the iron. Is this a problem? You bet your
buttonhole it is! The iron fusion consumes energy, so there isn't enough energy to help
support the star. Without support, gravity comes along and squeezes down the core.
What happens when you squeeze stuff? It gets hotter. The iron core gets hotter and
starts burning faster, which causes more energy to be sucked away, which removes
more support against gravity, which causes the core to compress more and heat up
more, which causes the iron to burn faster, which... I think you get the picture.
The core of the star collapses during the iron burning stage (since nothing is fighting
the gravity), which takes only about 1/4 second. The collapsing process shrinks the
core down to the size of the Earth. It gets very dense, up to the point of electron
degeneracy (remember, that's what a white dwarf is like). Will the collapse stop? No,
since the core is more massive than the Chandrasekhar limit (the iron core is about 1.5
solar masses in this case). The mass of the core is too much for the Chandrasekhar
limit, so electron degeneracy will not stop the collapse - it will keep going. Gravity
keeps crushing the star down until it reaches the point where the pieces of atoms are
crushed together. This is not an easy thing to do, but as the protons (p
+
) and electrons
(e
-
) are slammed together they form neutrons (n) and neutrinos ( ), which can be
written out as
p
+
+ e
-
n +
This reaction makes sense, because you combine a positive and a negative together
and just get neutral stuff out in the end. All of this compression and atomic mushing
results in the core of the star ending up as a big ball of neutrons. At this point the
collapse can be stopped by neutron degeneracy (10
14
g/cc is the density of such
material). Neutron degeneracy is much more extreme than electron degeneracy -
greater density, more extreme rules of physics and so forth. Because of the
degeneracy, the star will not get any denser so long as the neutron degeneracy can
183 | P a g e

hold it up. In the case where the degeneracy can't hold it up, you end up with a black
hole - more on this later.
The core that ends up as a ball of neutron degenerate material is called a Neutron
Star. This is a star so small and compact that a 1.5 solar mass neutron star would be
only about 20 km in size. Think of that - an object more massive than the Sun only the
size of a large city! It would be incredibly dense, 10
14
gm/cc. This is sort like the
density you'd get if you took 100 aircraft carriers and crush them down to the size of a
sugar cube. Try putting that in a cup of tea!
In about 1/4 of a second the core has been crushed down, resulting in a neutron star,
which are in most cases only a few solar masses in size. What happens to the rest of
the star? Remember, the neutron star core at this point is only a small part of the total
mass, so you still have quite a few solar masses to watch out for that is located beyond
the star's core.
The core collapsed very quickly from a size close to that of the Sun's to only about 20
km, so there is a gap in the support of the rest of the star. What support? - there is NO
SUPPORT! Nothing is holding up the rest of the star. This is sort of like when the
Coyote runs off a cliff and doesn't immediately fall down - at least not until he
realizes that he is off the cliff. The outer layers of the star don't really know that they
have had their legs cut out from under them for a moment, but once they do - watch
out. The upper layers will fall onto the ultra dense neutron degenerate core and the
material will heat up to about 5 billion degrees. This high temperature and the
corresponding high pressure will generate an incredible amount of energy. The energy
that is generated by the slamming of the outer layers on the core is huge. This energy
that is produced here in this small interval of time is the same amount as that given off
by the Sun over its entire lifetime (10 billion years). This huge bottled up energy is
released in a massive explosion that will blow off the outer layers - basically, the star
explodes. That's how you produce a Supernova. This little animation shows a blue
supergiant quickly collapsing down and then exploding as a supernova.

What is a supernova like? Here are some of its main characteristics -

184 | P a g e

The explosion blows away almost all of the mass of the star. What is left
behind may be only a few solar masses in size, though in the case of a black
hole it can be larger.
The core that is left over is in the form of a neutron star or a black hole.
There is the release of a large amount of energy, so these things are very bright.
A supernova produces the equivalent amount of energy as an entire galaxy
(billions of stars), and it can stay bright for quite some time - weeks or months
depending upon the distance. There have even been supernovae that were
visible in the daytime.
The iron fusion and the huge amount of energy from the collapse of the star are
so great that the fusion of even heavier elements occurs. All elements more
massive than iron require huge amounts of energy to form, since like iron, their
fusion processes consume energy. A supernova is the only thing that has energy
to spare for the fusion of the heavy elements, so this is the only way that these
things can be made. All of the copper, zinc, nickel, gold, silver, mercury, and
other elements up to uranium are produced in supernovae (these include
elements numbers 27 to 92 on the periodic table).
During the formation of the neutron star there is the release of neutrinos, and at
times these can be detected. Remember, there are neutrino detectors that are
currently working to detect neutrinos from the Sun and these detectors are able
to pick up supernova neutrinos as well.
A large shock wave is produced by the explosion. The shock wave can travel
through space and can compress gas clouds, which will lead to new star
formation. This can help explain why large scale star formation can continue
on, since large stars die relatively quickly, usually near to the location where
they were born. If one dies in a supernova near the location of its formation
(near a Giant Molecular Cloud), the shock wave from the explosion can ignite
new episodes of star formation in the GMC. The death of one star can lead to
the birth of many more. Recent observations by the Spitzer telescope appear to
support this scenario, with a region of star formation found near a likely
supernova - you can see a schematic of the event here. Some people have even
linked the explosion of supernovae to things on the Earth, such as climate
changes or various large extinction events - but those are only theories.
Massive stars are pretty rare, so on average there is only one supernova occurring in a
galaxy every century.
Now I'm going to complicate matters a bit. There are actually two main types of
supernovae. One is the type that I just described; the other occurs when a white dwarf
is near its mass limit (the good old Chandrasekhar limit=1.4 solar masses) and is
pushed over this limit when too much mass is dumped on it, usually in a binary
185 | P a g e

system. You may want to refresh your memory on the stuff about novae in the
previous set of notes. If the white dwarf star in the binary system is really big (close to
1.4 solar masses), rather than just going nova when mass is dumped on it, it will be
too massive to hold itself up and may instead become a supernova. There is also a
theory that if you had two white dwarfs in a binary system and they collide, it will
become a type I supernova. Either way, the white dwarf ends up being too massive
and collapses in on itself. Since there are two vastly different kinds of supernovae -
and they have to be distinguishable, and they are, mainly because the object that
explodes in each case is very different (massive star versus white dwarf). To
distinguish between the two types, the following designations are given
Type Ia Supernova - a white dwarf going over the Chandrasekhar Limit
Type II Supernova - a massive star collapsing and then exploding when iron
fusion starts

Figure 4. The light curves of the
different types of supernovae are
shown - note that the Type Ia
supernovae are brighter. Also, the rate
at which they brighten and fade away
is different. This helps astronomers
distinguish the two types. This graph
uses absolute magnitude as the
brightness scale since the
corresponding luminosity values are
around a few billion solar
luminosities.

Figure 5. The spectra of the different types of
supernovae are shown. The Type Ia supernova has
absorption (the valleys) and emission features (the
peaks) associated with heavy elements, while the Type II
has only hydrogen (H and H ) showing up
prominently in its spectrum. Spectra were obtained from
the spectra archive at the University of Oklahoma.
In general Type Ia supernovae are brighter by about two magnitudes or so than the
Type II. You would have to know the distance to the supernova if you want to use the
186 | P a g e

brightness as a way of categorizing it, so that isn't a good method. Fortunately it is
possible to tell the two supernovae apart by looking at their spectra. The object that
goes supernova is quite different in each case, so the spectrum from each type of
supernova is distinct. A white dwarf is the burned out core of a dead star, so it is made
mainly of stuff like carbon, oxygen, nitrogen, etc., but not a lot of hydrogen. A
massive star, on the other hand, is still mainly made of hydrogen, so when it explodes
its spectrum will be full of hydrogen. It is pretty easy to distinguish the two types of
supernovae without knowing their distances. You'll notice that if you do know which
type of supernova you have, you can then use its typical maximum brightness (from
Figure 4 above) and the apparent magnitude that you observe to get its distance.
One very unusual Ia supernova is 2006gz. This has a spectrum just like a 'regular' Ia,
but it was brighter than normal. It also had too much carbon and silicon, which led
some astronomers to speculate that this was actually a collision of two white dwarfs.
It has also been proposed that 2006gz came from a super-Chandrasekhar limit white
dwarf - an abnormally large white dwarf. So far we do not know the answer to this
unusual question, but it does show that sometimes very strange things do happen.
There was also a recent study by the Chandra observatory that seemed to indicate that
most type Ia supernovae are actually produced by the merger of two white dwarfs,
rather than a single star. So until more evidence is found (because that's how science
is done), we'll just have to use the general phrase "white dwarf going over the mass
limit" - which could occur due to very different reasons.
You should be asking yourself Why is it 'Ia'? Why not just call it a 'I'?. Good question.
A type I supernova has very little hydrogen in the spectra, but sometimes it is not
because the star that exploded was a white dwarf. There are two other type I
supernova, with the amazingly original names of Ib and Ic. Both are thought to come
from large mass stars that have blown away most of their outer layers so there is not
much hydrogen left in their spectra. The main difference between Ib and Ic is whether
there is any helium left - a Ic has no helium in their spectrum, while a Ib still has some
helium. Both the Ib and Ic are fainter than a Ia, and a bit rare since they only come
from very massive stars. Unlike the massive stars involved in the type II supernova,
these have lost too much of their mass to have a big blast. Generally when we talk
about type I supernovae, we are referring to type Ia.
Recently astronomers have found that there are some supernovae that could be better
described as super-supernovae - but that's sort of a silly name. The
term hypernovae has been proposed for these extreme explosions. There is a bit of
debate as to exactly what happens during a hypernova, since this is a relatively new
idea and most of them have only been observed to occur at very great distances. One
option is that a very massive star (more than 30 solar masses, possibly up to 150 solar
masses) that had been previously losing mass eventually collapses in on itself, causes
187 | P a g e

a massive explosion and eventually forms a black hole. Another option is the merger
of unusual stars, such as two neutron stars, which results in a massive explosion
(you'll learn more about neutron stars in the next set of notes). One side effect of such
an event is the emission of a large amount of gamma-rays, which aren't normally
observed from stars. Now a days such gamma-ray bursts can be spotted with
the Swift telescope, and several possible hypernova have been discovered in that way -
though it wasn't until after we look at them with other telescopes, light visible light or
x-ray telescopes are we certain about their overall energy output. The overall energy
of a hypernovae is generally 100 times greater than the energy given off by "normal"
supernovae. Currently the record holder for the most powerful stellar explosion is the
object known as SN2006gy (yes, that is a lame name), which was observed in 2006
and had an energy output that was greater than any other supernova. Here is a
comparison of this objects brightness over time compared to other "normal"
supernova. This is similar to Figure 4 shown above. Here is an animation showing the
explosion based upon the x-ray and gamma-ray data from the region following the
explosion. It is interesting that there are two large bubbles of material, which were
given off by SN2006gy before the explosion occurred. This is very similar to what we
see today in Eta Carina (Figure 2), an object in our own galaxy which might someday
erupt as a hypernova. Several other stars are thought to have also had smaller
outbursts before they blew themselves to pieces as supernova - sort of a hiccup before
a really big belch! And in case you were wondering how you missed such an amazing
explosion in 2006, don't worry, not too many people noticed it since SN2006gy
occurred in a galaxy that is about 240 million light years away, so it was barely
visible, except with the largest telescopes.
While most gamma-ray bursts are invisible to us (since gamma-rays cannot reach the
surface of the Earth), it is still possible to "see" them, since the total energy given off
is huge. Gamma-ray bursts give off light at all wavelengths, not just gamma-rays
(remember hot objects give off light at many wavelengths even though they have only
one peak for their emission). So when the Swift satellite observes a burst, the location
is transmitted to regular ground based telescopes, both visible light and radio, in an
effort to measure the light output in as many wavelengths as possible. In March 2008,
there was a gamma-ray burst that was so powerful and concentrated that it could have
been seen by the naked eye for about 15 seconds. That's pretty impressive considering
that the object that produced the burst was about 7.5 billion light years away!
Supernovae are sort of rare, so astronomers are only able to observe them occurring in
distant galaxies. Currently about 200 or more supernovae are observed each year in
other galaxies - in 2003, nearly 330 supernovae were discovered, some nearly a
billion light-years away. While this may seem okay, the problem is that since these are
very distant supernova, there isn't a lot of detail visible. For the most part, astronomers
188 | P a g e

can figure out the type of supernova in a distant galaxy by obtaining a spectrum, but
that's about all. Astronomers over the centuries have seen supernovae, some in our
own galaxy, though in the old days they didn't know what they were. By looking at
the records of various cultures, astronomers have figured out that some unusual
astronomical events that mystified people in the past were actually supernovae. Here
are some of the more famous ones -
1054 A.D. - Chinese and Arabic astronomers observed a new star appear in the
constellation of Taurus the bull - this is a very important supernova and we'll
run across it later.
1572 A. D. - Tycho Brahe made careful observations of one in the constellation
of Cassiopeia
1604 A. D. - Johannes Kepler (and many others) observed one in the
constellation of Ophiuchus
Kepler's supernova (as it is sometimes called) was the last supernova that was
observed to go off in our galaxy. We haven't seen a star become a supernova in our
galaxy in about 400 years - we're long overdue for one! Of course, it is possible that
other supernovae have occurred in our galaxy since Kepler's, but they may have
happened in distant parts of our galaxy, and we couldn't see them. As you'll see there
is good suspicion that this is the case.
While it is difficult to see a supernova in great detail today (since most are so far
away), it is sometimes easy to find the material left over from the explosion. This is
because there is not only a lot of material, but it stays relatively hot for quite a long
time. The gas cloud left over from the supernova explosion is known as a Supernova
Remnant (SNR). Like an explosion in a fireworks display, it takes a long time for the
cloud to fade away, though in the case of a SNR, the cloud can hang around for
thousands of years.



Figure 6. Various Supernova Remnants are shown. To the left is the Crab Nebula, the remnant
189 | P a g e

from the Supernova observed by the Eastern astronomers in 1054 A. D. This visible light image
is from the Very Large Telescope. In the center is the Cygnus loop, as seen by an x-ray telescope.
This is a very old remnant, and it is also very large. The line in the lower right part of the picture
is the width of the Full Moon. On the right is an x-ray image of the remnant that was Tycho's
supernova (observed in 1572 A. D.). If you click on the Crab or Tycho images, you'll be able to
see the Chandra x-ray image of those remnants. (Image credits: VLT, ROSAT, MPE, NASA,
NASA/CXC/ASU/J. Hester et al., NASA/CXC/Rutgers/J. Warren & J. Hughes et al.).
These aren't just neat things to observe; they can provide useful information about
supernovae. This is done by measuring the velocity of the gas and combining that
with the size of the object. Doing this provides astronomers with the age of the
supernova (since velocity = size / time since explosion). Even though we might not
have seen the explosion, we can estimate when it occurred. Many SNRs seen today
are still rather hot even though they may have "gone off" hundreds of years ago. They
often have x-ray emissions - which tells us that they are still really hot. The
supernovae described earlier, like the one in observed in 1054 and those seen by
Kepler and Tycho, all left behind large expanding gas clouds. The one seen in 1054
can be observed today as an object called the Crab Nebula (or, by its catalog name,
M1).


Figure 7. Cas A, a supernova
remnant. The image to the left is
from the Very Large Array radio
telescope (NRAO). The image to
the right is from the Chandra x-
ray telescope
(NASA/CXC/MIT/UMass
Amherst/M.D.Stage et al). Click
on either image to see a larger
view.
One rather confusing SNR is Cas A, a SNR in the constellation of Cassiopeia (that's
where the "Cas" comes from in its name). This is a very strong radio source, as well
as a strong x-ray source (see Figure 7). Obviously it is still rather hot. It is also
relatively small. Combining this information with velocity and size data tells
astronomers that this object blew up about 300 years ago. We haven't seen a
supernova in our galaxy for about 400 years! Somehow, this thing blew up and no one
noticed it! It also looks like we missed another supernova that was even more recent -
only 150 or so years ago! Unfortunately it was in a rather messy part of the galaxy, so
it would have been difficult to see even if we knew it was going on.
190 | P a g e

Another neat SNR is the Gum Nebula. This is a large gas cloud, and it is really large
because it is really old. It is estimated that this object blew up sometime around
10,000 BC. Not only is it rather old, but it is also relatively nearby. Taking this into
account, when this thing blew up, it would have been as bright as the Full Moon!
Let's see what we have got so far. Astronomers can study supernovae in other
galaxies, but they are so far and faint that only the largest telescopes we have can see
most of them. They can study supernova remnants to try to figure out what happened
in the past when these things blew up. That's sort of boring, eh? It isn't all of the time.
Actually, things in the astronomical community got rather crazy not too long ago, on
February 23, 1987, to be exact. On that night there was the event of a lifetime -
Supernova 1987A, or SN 1987A for short. Supernovae are named for the year (1987
in this case) and a letter for the order of their occurrence in the year. The first one of
the year is A, the second is B, etc. In 1987, the first supernova seen was labeled SN
1987A.
Figure 8. Supernovae 1987A - image of the supernova at
its brightest in the part of the sky it appeared in. In the
upper left of the image is the Tarantula nebula, a large H
II region. These objects are actually thousands of light-
years away - around 160,000 light-years in fact. Image
Anglo-Australian Observatory, Photograph by David
Malin.
What was the big deal about SN 1987A? This was the first naked eye visibility
supernova observed since 1604 - you didn't need a telescope to see it; that's how
bright it was! While it did not occur in our galaxy, but in one of our neighboring
galaxies (the Large Magellanic Cloud is the galaxy it happened in), it was still close
enough that detailed studies of it could be carried out. The supernova was discovered,
or perhaps a better word is "noticed," by Ian Shelton on the night of Feb 23, 1987. It is
quite likely that others saw it but they did not recognize it for what it was.
When SN1987A went off, virtually every telescope that could observe it was used to
study it. This included telescopes located in countries such as Chile, South Africa, and
Australia, mainly because the Large Magellanic Cloud and the supernova are very
southern objects. Satellites were also used - including the ultraviolet satellite, IUE
(which made over 600 observations) and a variety of rockets that obtained x-ray
images. The Hubble Space telescope wasn't in orbit at the time, but it has been
looking at it since.
191 | P a g e

Figure 9. The brightness
variation of SN1987A over time.
This chart shows thousands of
observations by both
professional and amateur
astronomers. The time scale at
the bottom is in days, but uses a
rather strange system. The
entire graph covers a time
corresponding to about 3.3
years. The magnitude scale is on
the sides. Remember, the faintest
that you can see with your eye is
magnitude 6, so SN1987A was
visible to the naked eye for about a year. Chart is courtesy of the American
Association of Variable Star Observers (AAVSO).


Another added advantage to this supernova over others that are studied is that we
knew which object blew up. This area of the sky is fairly well studied, so many stellar
surveys were done and the star in question, Sanduleak -69 202, was cataloged and its
characteristics (color or temperature and luminosity) were known before it became a
supernova. Why was that important? We knew what type of star exploded
(temperature, luminosity, likely mass, composition, etc.), so information about it
helped astronomers refine computer models of supernova events. The supernova was
bright for a long time (visible to the naked eye for about a year), so its evolution was
followed closely (and still is being followed) to see if our theories about supernova
remnants are correct. The rate at which the supernova brightened and is fading away
can also provide information for various computer models and can be compared to
other supernovae.
Figure 10. SN1987A and the
star that eventually went
supernova, Sanduleak -69
202. It was a relatively
bright star to begin with, but
when it went supernova (as
is seen on the left), it got
very bright. Image Anglo-
192 | P a g e

Australian Observatory, Photograph by David Malin.
The spectrum of the explosion confirmed the heavy element production (elements
heavier than iron). This was seen in the way that various elements would appear and
then decay over time. This was exactly in line with the theories about the heavy
element production - a nice example where the theories that have been around for a
long time finally were supported by observations. Also, by measuring the expansion
velocity, we can determine the distance to the Large Magellanic Cloud galaxy in a
very accurate manner - something that is very difficult to do. Probably one of the
more exciting discoveries was of the neutrinos from the supernova. All of those
neutrino detectors around the world were set up to look for neutrinos from the Sun.
On the day of the supernova, these detectors were practically flooded with neutrinos.
By "flooded" I mean they detected maybe six or eight neutrinos, which is considered a
flood when compared to the normal number of neutrinos that are detected. This was
another case of theory and observations coming together!
When you look at the supernova today, you can see several rings of material around
its location. This material is not part of the supernova explosion, but was blown off by
the star years before (you may want to go back up to the top of the notes to look at
those big stars that are currently blowing off mass, especially Eta Carina). It wasn't
until the supernova went off that the ring actually did something - in this case the
rings lit up, because the light (energy) from the supernova traveled through them and
heated up the gases.
Figure 11. How the area around SN 1987A looks today.
The supernova is in the center of the multi-ring
structure. The inner ring is less than a light-year from
the supernova, while the other rings are a few light-
years away. The rings are not part of the supernova, but
are made of material that was blown out from the star
before it went supernova. To see why the rings have the
orientation that they do, just click on this link. Image
credit: P. Challis (CfA).
Today the supernova is too faint to see with the eye, but its evolution is still being
followed, particularly what is going on with the rings around it. The distance from the
supernova to the nearest ring is less than a light-year. The shock wave from the
supernova started hitting the ring material in 1998, which is causing it to light up
again. It will take some time for the shock wave to travel through the ring, and you
can expect more neat pictures from theHubble Telescope showing that. Here is the
latest movie showing the ring lighting up due to the shock wave. Even though the
193 | P a g e

supernova did happen quite some time ago, it is still one of the major events in
modern astronomy.
Neutron Stars and Black Holes

What's covered here:
What is a neutron star?
What is a pulsar?
What is the Special Theory of Relativity about?
What is the General Theory of Relativity about?
What are the characteristics of a black hole?
What would happen to you if you fell into a black hole?
What other exotic objects exist in the sky?

What does a supernova leave behind? That sort of depends upon the size of the
original star. Remember, mass determines everything about the life of a star - so here's
another example of that rule. In the case of massive stars (those that die via the Type
II supernova mechanism), there are two likely possibilities - a neutron star or a black
hole. We'll look at the less extreme case of the neutron star first.
Neutron Stars and Pulsars
If you recall from the discussion of Type II supernovae, the core of the star collapses
to form a big ball of neutrons, what we would call a neutron star. This ball of
neutron degenerate material is packed in so tightly that a teaspoon of neutron star
material would weigh as much as a mountain. The cores of some massive stars end up
as neutron stars, and this is where things get tricky. A neutron star is so small (only
about 20 km in diameter), so it was thought unlikely that such an object would be
visible even using the most sophisticated telescope. People had theorized about the
existence of neutron stars since the 1930s, but because of their small sizes, people
thought they would only be theoretical objects that could never be detected.
Let's turn the clock back to 1967. Over in England, the Cambridge University Radio
telescope project was in progress. While this did not involve what we would call a
really sophisticated set up, it was able to do the job. Basically the telescope was a
large array of wires spread out over 4 1/2 acres, kind of like a great big wire
clothesline strung out over a field. It was just a bunch of wires on posts, so the
194 | P a g e

telescope could not be aimed or moved. How did it look at the sky? The rotation of
the Earth allowed the telescope to look in different parts of the sky over the course of
a day - though there weren't many places it actually could look - it just pretty much
scanned the same strip in the sky day after day.
A graduate student, Jocelyn Bell, was looking at the radio signals coming from the
telescope and she noticed that a strong radio signal occurred when the telescope was
aimed at certain regions in the sky. The telescope pointed in that same direction each
day, and she noticed that the signal reappeared not 24 hours later but 23 hours and 56
minutes later. The signal was detected four minutes earlier each day. What does that
mean? Remember, the time for the rotation of the Earth (the sidereal period) is 23
hours, 56 minutes, so sources in the sky will reappear four minutes earlier from the
previous day's observation. This told her that the object giving off the signal was up in
the sky. It took her a while to convince her boss of the source of the signals - he
thought they were from an Earth based source, like telephone lines or radio
transmitters. Eventually he was convinced and they started to examine the signals
more closely.
Figure 1. A typical pulsar signal,
as picked up by a radio telescope.
They noted that the signals were
not a continuous radio signal (like
a steady humming noise) but actually a bunch of individual radio "pulses" (like a heart
beat). The pulses were very rapid, so they often had the appearance of a steady,
sustained signal. The timing of the pulses was pretty precise - very little deviation
from one pulse to the next. Due to the pulsed nature of the signals, the objects were
initially called Pulsars.
What exactly are pulsars? Believe it or not, at first they did think the signals were
from extraterrestrials due to the very precise nature of the pulses, but later trashed this
idea when they discovered a wide range of pulse periods (the time between each
pulse) located in many different directions in the sky. If these were alien beacons, they
certainly didn't appear to be very useful.
A sound file of a typical pulsar
(1.4 pulses/second)
A sound file of the Crab Pulsar
(30 pulses/second)
A sound file of a fast pulsar
(642 pulses/second)
What type of object gives off periodic signals, or changes its energy output in a
periodic manner? Eclipsing stars can do that. You could get periodic signals from an
eclipsing system, but since the signals from the pulsars had periods on the order of
195 | P a g e

one second, that would mean stars would have to be moving around in orbits once
each second - that is just a little too fast to be possible, so that idea was scarped.
What about stars that actually pulsate? Remember, the Sun has periodic vibrations,
and there are many types of stars that do pulsate. We run into that annoying problem
of the period of variation being pretty fast. If a star were to pulsate at the rate of one
pulse each second, it would be wobbling so much it would rip itself apart. Again, the
pulses are just too fast to be this option.
What else do we have that is periodic in nature? You probably remember that stars
also rotate. If a star had an unusual feature on its surface then that feature would
appear each time the star rotated toward you, sort of like a lighthouse beacon. This
scenario is possible. What type of star could rotate around at a rate of about one
rotation per second?
Could a Main Sequence star do it? No, it would tear itself apart. What about a white
dwarf? No, it would still have to rotate too fast for it to survive. We need a star that is
very small and very dense so that its self-gravity (what's holding it together) could
withstand the strong rotation (what's ripping it apart). What sort of object would be
appropriate to be a pulsar? What could it be; I wonder, could it be...a neutron star?
It seems that the only object that has enough self-gravity to hold itself together while
it rotates at very high speeds would have to be a neutron star - very small, very dense,
but able to handle the difficult job. It looks like that pretty much answers the question
- or does it? Just because we think that a neutron star is the source of the radio pulses
doesn't actually prove that it is cause. As with anything in science you need to get
evidence to support the idea. Science involves doing tests and finding evidence - sort
of like detective work. Now comes the difficult part, since it is unlikely that we can
actually "see" a neutron star directly with a telescope to test this theory (remember
how small its diameter is). How do we find the evidence?
If you remember, we thought that neutron stars were produced in a supernovae
explosions. We need to link a supernova with a pulsar. If we take a closer look at one
of the signals that was discovered very early, you will notice that it is originating from
a gas cloud known as the Crab Nebula, which is located in the constellation of Taurus
the Bull. Observations of this cloud show that it is changing at a fast rate - it is
expanding pretty quickly. It is also a strong x-ray source as well as a strong radio light
source. Astronomers thought that it was a supernova remnant. Why do you suppose
they thought that?
196 | P a g e

Figure 2. An up close image of the Crab
Nebula obtained by the Hubble Space
Telescope. The pulsar is indicated.
Credit: NASA and The Hubble Heritage
Team (STScI/AURA).
Astronomers took a look at historical
records that tell us that something
happened there many years ago.
Observations made in July of 1054 AD
indicated that a bright object appeared in
the sky. Just how bright was this object?
It was brighter than Venus! It was so
bright that it was even visible in the
daytime for quite some time.
Observations of the object were made by
the astronomers in the Far East (China
and Korea) as well as Arabic
astronomers. It seems that the Europeans were too busy with other things to notice the
object, or perhaps they thought that since nothing can change in the sky, it must just
be a local event, like a meteor shower. The Far Eastern astronomers were pretty
accurate in mapping the object's location, so we know that the bright object appeared
in the same spot of the sky where we now see the Crab Nebula. Also, the brightness
measurements that were made follow the brightness variations of a supernova (a Type
II supernova, in fact).
There's the smoking gun. A supernova was observed in 1054 AD. Today we look at
this same part of the sky and see not only the Crab Nebula (which was discovered
long ago) but also a pulsar right in the middle of the mess. The pulsar at the center is a
pretty fast spinner, rotating at a rate of 30 times/second. That seems to confirm the
theory - a pulsar that is directly related to a supernova, an event that we thought would
produce a neutron star, and appears to have done just that. The Crab Nebula pulsar is
one of the few that can be seen with a regular visible light telescope. Long ago
astronomers surveyed that part of the sky and thought that the star that we now
identify with the Crab pulsar was just a chemically peculiar star. It is spinning so fast
it is difficult to see the flashes with the naked eye. There are several other supernova
remnants with pulsars in them. One, with the exciting name of G11.2-0.3, was
recently studied by the Chandra space telescope and might possibly be a supernovae
that was observed by Chinese astronomers in 386 AD. You can read about it here.
That's the basic upshot (at least once the mystery was solved). Pulsars are rotating
neutron stars! Following the standing rule in astronomy, the name is misleading -
197 | P a g e

pulsars rotate, but they don't pulsate! This is probably a good rule about naming things
too early - the name stuck before they knew what they were dealing with.
With that mystery cleared up, there are still some questions that need to be answered -
namely, why are pulsars rotating so fast, and what is actually causing the pulses?
We'll tackle the easy one first. The fast rotation is due to a rule concerning spinning
objects - the conservation of angular momentum. This is a relation between the rate
of rotation (how fast it spins) and the distribution of mass in the spinning object (how
spread out it is). As the mass distribution changes, the rate of rotation changes. By
decreasing the mass distribution (bringing it in closer), you increase the rate of
rotation. You may have seen this in the way that an ice skater will spin faster when
they bring their arms in. A change of mass leads to a change in the rotation rate. What
does this mean for a pulsar?
First of all, pretty much all stars spin to some degree. Remember, the Sun spins
around once every 30 days or so. During the supernova event, the massive core
collapses in an extreme way. As it collapses (as the mass is brought in), the rotation
rate increases. The reduction in size is so great that the rotation rate increases by a
drastic amount - thousands of times faster than the previous rate. A star that was
spinning around at a rather leisurely pace will end up spinning around very, very fast
due to the extreme collapse of the core. I suppose this would be like having an ice
skater with 1000 pound arms that are brought in suddenly during their rotation. The
skater would probably rotate so fast they'd burn a hole in the ice.
Currently the fastest spinning pulsar that we have detected is XTE J1739-285 (what a
lovely name), and is spinning around at 1122 rotations per second. Yes, that's right,
1122 rotations every second! That would make anyone dizzy.
That's one mystery solved, so now we have to figure out what is causing the pulses.
Another thing that stars have are magnetic fields. A weird feature of magnetic fields
is that they get very strong when they are compressed. Guess what happens to the
magnetic field during the collapse of the star's core in the supernova event? You
guessed it, the magnetic field gets compressed. In the case of the formation of a
neutron star, the compression is so great that the increase in the magnetic field
strength is also incredibly huge.
A neutron star would tend to have a really strong magnetic field - how does that make
the pulses? In the neighborhood around pulsars there are charged particles (mainly
electrons) that tend to travel along the magnetic field lines. This has to do with the
way magnetic fields and electrically charged things interact. These electrons are
198 | P a g e

zooming along the magnetic field lines. As they move along them, they get
accelerated (sped up) by the magnetic field, especially near the magnetic poles of the
pulsar. It is a strange feature of electrons that when you accelerate them, they give off
radiation - synchrotron radiation. I should mention an important aspect of this
radiation. Previous to this all of the light (radiation) that we have come across had its
origin in objects that were giving off light because of their temperatures. Synchrotron
radiation is not produced by heat and therefore is rather unique. We refer to this type
of light as non-thermal. We'll run across more non-thermal radiation sources later.
Any ways, synchrotron radiation is mainly found in the radio part of the spectrum,
which is exactly why Jocelyn Bell and the Cambridge radio telescope detected it.
Very few pulsars can be seen with a visible light telescope (only those that are
relatively nearby, usually).
Figure 3. The model for pulsars. The
magnetic field of the rotating neutron star
gives off synchrotron energy. As the magnetic
pole sweeps across the sky, radio light is then
sent toward observers in various directions.
This is the classic lighthouse model. You can
view a little animation of a pulsar here. Only
when the light source is directed toward the
Earth is the pulse detected. Picture Credit:
NASA.
Now let's see what we have - a pulsar, spinning very fast, has a whole bunch of
radiation coming out of its magnetic poles - so why would that make a pulsed signal?
This depends upon the way that the magnetic fields are arranged. The magnetic poles
may be arranged in any sort of orientation relative to the rotation of the star. They are
not necessarily aligned with the axis of rotation. As the star rotates, the magnetic poles
will sweep by, sort of like the way a lighthouse beam sweeps past. These sweeps
produce the pulse of light that we detect. Often two pulses are detected from the two
magnetic poles, since they cover a good fraction of the star's surface. The high
rotation rate of the pulsar makes the pulse very short in duration - now you see it; now
you don't.
199 | P a g e

Figure
4. Light
pulses of
several
different
pulsars in
different
wavelengt
hs. In
some cases
pulsars
are visible
at other
wavelengt
hs than radio, the most common type of signal. Geminga is one of the few pulsars that
does not give off a detectable radio signal, but it does emit x-rays and gamma rays.
The Crab pulsar is one of the few that can be seen in visible wavelengths. Picture
credit: D. J. Thompson, NASA/MSFC.
That pretty much explains pulsars (which, you remember, don't pulse). There are over
1000 pulsars that have been discovered in our galaxy so far. Most pulse (well, they
really don't pulsate, but you know what I mean) at a rate of about 1 pulse/second,
though there are some beasts that pulse at rates of over 500 pulses/second. That is a
very fast pulse! Pulsars do slow down very slowly, and over time they lose their
energy, so eventually they won't be detectable.
There is another interesting aspect about neutron stars - they are neutron degenerate.
What does this mean, apart from the fact that they are very dense? It means that if you
add mass to a neutron star, it will decrease in radius, just like a white dwarf will
shrink in size if you increase its mass. If you have a neutron star in a binary star
system and mass is transferring to it, then it would decrease in radius. What happens
to a rotating object that decreases its mass distribution (shrinks in size)? You got it - it
spins faster. There are actually a couple of pulsars that are spinning faster, and we
think this is due to their gaining mass - rather strange, but true. If you thought that
pulsars and neutron stars were strange, then just keep reading and learn about some
other, even stranger things.
Black holes
Now if you remember about white dwarfs, they are electron degenerate and they have
an upper mass limit (the highest mass that they can have without collapsing in on
themselves). Now we just finished off neutron stars, which are neutron degenerate.
200 | P a g e

Does the same sort of rule apply? If so, what would happen to a neutron star if it went
over the neutron star mass limit?
The only problem is that the laws of physics that operate on neutron stars are more
complicated than those for white dwarfs, so we really don't have a definite answer to
this question. Based upon our best ideas about neutron degeneracy we think that there
might be a mass limit, like 4 or 5 solar masses. We have seen quite a few neutron stars
in binary systems and the masses that we find for them tend to be in the neighborhood
of 1.5 solar masses. It doesn't appear that there are very many massive neutron stars
out there, which is in line with our theories.
Let's say that there is a massive star out there that is dying and its core mass is pretty
large, maybe 6-10 solar masses or so. What does happen if the core of a star is too big
to be a neutron star? If it is too big, it has too much gravity, and the collapse will keep
going on. Don't forget about what I said about gravity - it rules! If the gravity is too
big, then nothing, and I mean nothing, will stop the collapse. We've got something
that is getting smaller in size (radius), smaller, smaller and smaller. It's just getting
smaller, right? That's no big deal, right? - wrong, wrong, wrong, wrong! Remember
the rules for gravity? Remember that the force of gravity depends upon the size of the
object - that it goes up strongly for a small decrease in radius. Therefore a big
decrease in radius, like that of a collapsing core, would cause a really, really, really
huge increase in the surface gravity.
How big is really, really, really huge for surface gravity? What does gravity do any
ways? It prevents things from getting away. Let's say you wanted to go to another
planet. First of all, you would need to have a rocket that would go fast enough to
escape from the the Earth's surface. If the rocket goes just a little bit too slow, it will
fall back to the surface of the Earth, which is not a good thing. The velocity that is
needed to get away is called the escape velocity, and all it depends upon is the surface
gravity of the object that you are trying to get away from. The formula for the escape
velocity is

That's a pretty scary formula, but it's not too much different from the gravity formula -
the bigger the mass of the object you are getting away from, the higher the escape
velocity. The same thing happens if the radius goes down - the escape velocity goes
up, so you need to go faster to get away. Just how fast do you have to go? Let's try a
few examples. Let's say that you want to leave the Earth (and who doesn't?). How fast
would you have to go? - only a measly 11.2 km/s. Does that sound fast to you? Maybe
201 | P a g e

you didn't read that carefully - 11.2 km per second. That's like going across town in
one second. You could get a speeding ticket for that, assuming someone could catch
you.
We were looking at a collapsing massive star, remember? What happens to the escape
velocity when the core is collapsing? It gets larger, right? How much larger does it
get? Here are some more examples.
The Sun - it's pretty big, right? To get off of the surface of the Sun (which would
probably be a good thing since it is a tad too hot) you would need to move at a speed
of 436 km/s. This is like going from Waterloo to Chicago in one second - pretty fast!
We can do better - just by decreasing the radius. Let's say that there is a white dwarf.
White dwarfs are similar to the mass of the Sun, so the only thing that makes a
difference to the escape velocity is their small radius (remember, they're only about
the radius of the Earth). To get off of the surface of a white dwarf, you would have to
go at about 4700 km/s. How fast is that? That's like traveling across the country in one
second.
We can do better than that - go to the next level of extremely small objects with pretty
good masses - neutron stars. These little beasts have masses similar to the Sun's, but
radii of only about 20 km. The escape velocity gets really, really huge - in this case
around 100,000 km/s. This is pretty fast, so fast that you can even describe the
velocity in terms of the speed of light - the escape velocity from a neutron star is 1/3
the speed of light. Now that will certainly get you a speeding ticket on the freeway.
Just how small can we make an object (and how high can we push up the escape
velocity)? If we make the object smaller in size than the neutron star (but having the
same mass, similar to the Sun's), the surface gravity will get bigger and bigger,
until...what? The velocity gets up so high that the escape velocity reaches the highest
possible speed - the speed of light. When you get to the point where the escape
velocity is 300,000 km/s (which is the speed of light), then you have that amazing
object known as a Black hole.
Figure 5. Albert Einstein - the fellow with the messy hair who
gave us the two theories of relativity. His hair probably had
nothing to do with it, but who knows?
Now a lot of people have really weird ideas about what black
holes are and how they act. Don't believe it - they aren't that
bizarre. To understand black holes and how they interact with
the rest of the Universe, we need to understand some of the
rules that are in effect around them, how they influence things (gravity related stuff)
202 | P a g e

and how light operates. These are explained in some ofAlbert Einstein's theories.
Now don't panic - these aren't all that bad, and I'll just go over the basics so that you
understand what is going on with them.
First we'll look at the easier of the two theories - the Special Theory of Relativity.
I'm not joking about this one being easier - it really is. You've already run into this
theory in part, since this is where we get that famous formula E=mc
2
. That is not the
most important part of the theory; it's not even a major component. The main feature
of the theory can be stated in the following concept - the speed of light is a constant
no matter what is going on, no matter what you are doing, and there is nothing (I
really mean nothing) that can travel faster than the speed of light. Now, we already
ran over the first part of this when we talked about the properties of light - and
actually without that concept the rest of the Special Theory of Relativity wouldn't
exist. The second part is sort of like a cosmic speed limit. Nothing goes faster than the
speed of light, not even when the cops aren't looking. You might not think that these
rules are a big deal, but if you relate it to how things move on the Earth, then it is a
big deal.
Figure 6. An example of non-relativistic (low
velocity) motion. The final velocity of the paper
towel roll (or football) would be 100
miles/hour (assuming no air resistance or
friction).
Here is an example of what I mean. Let's say
you have a truck going down the road at 60
mph. In the back of the truck is former UNI quarterback Kurt Warner. He can throw a
roll of paper towels with a speed of about 40 mph. Let's say that the truck is going the
same direction that he is throwing. Just how fast will the roll of paper towels be
moving when it is caught (ignoring the effects of wind resistance and such)? You
basically add up the velocities, so the paper towels would be going at a speed of 100
mph. Even though they are only paper towels, I would not want to be on the receiving
end of the toss in this case.
The example illustrated above is pretty straightforward - you just add up the velocities
(if they are going in the same direction). What if Kurt is throwing in the direction
opposite of the direction that the truck is going? Then you would subtract the
velocities. This is what happens to things going at non-relativistic speeds (speeds
much less than the speed of light). Let's do another experiment, but this time in space.
Let's say you have a spaceship going at 1/2 the speed of light (v =1/2 c) - which is
within the rules; remember, things can go at speeds less than that of light, so 1/2 the
speed of light may seem pretty fast, but it is entirely legal. Let's say that hanging on
203 | P a g e

for dear life outside of this spaceship is an astronaut who has a flashlight. Let's say
that the astronaut aims the light in the same direction that the spacecraft is moving. Of
course, light from the flashlight is traveling at the speed of light (v=c). Now let's say
we have a detector aimed at the rocket and the flashlight, and this detector is
measuring the speed of the light coming from the flashlight. How fast would the light
be traveling as we measure it? If you went by the paper-towel-throwing example
above, you might think that we would measure a speed of 1.5 c (the sum of the two
velocities). Unfortunately, that isn't correct. What does the special theory of relativity
say? It says that the speed of light would always be c (300,000 km/s) no matter what
is going on. Even though the ship is moving at 1/2 the speed of light, the speed we
measure for the light will always be c - ALWAYS!!! It doesn't matter if the ship is
coming toward us or going away from us; we'd always measure the same speed for the
light.
Figure 7. An example of relativistic velocities. Here the
ship is traveling at 1/2 the speed of light (1/2 c), while the
light from it travels at the speed of light (c). Unlike the
example shown in Figure 6, you can't just add the velocities together, based upon
Einstein's Special Theory of relativity. The light coming from the spaceship will
always travel at the speed of light (v=c).
Isn't that just weird? Did we somehow or other lose 1/2 c of velocity? No, we don't
lose it, because we can't gain it in the first place. Light is just a rather strange thing.
No matter what you do, you can't get a velocity for light greater than c. Don't feel
disappointed, because even though we don't get the expected result of adding
velocities (like we got with Kurt Warner), we get some other nifty stuff happening. If
we were outside of the spaceship watching it go by, and we were to measure the
length of the spaceship, we would measure a different length than would be measured
by someone in the spaceship or if the spaceship were not moving. If we were to
measure the mass of the spaceship from our location, it would be different than if we
were in the spaceship or if the spaceship were not moving. If we were to look inside
the spaceship at its clocks, we would think they were broken since they would be
going at a different rate than clocks not on the spaceship. This is sort of a way that the
Special Theory of Relativity makes up for the fact that you don't get the addition of
velocities like in the previous example. This also sort of explains why we call it the
Special Theory of Relativity - what you measure depends upon your
situation relative to the thing you're measuring. If you were on the spaceship and you
were measuring things that aren't moving along with you (stuff outside the spaceship
for example), you would measure very bizarre things, sort of like the same way
someone who isn't moving would be getting strange measurements for the
characteristics of your spaceship. You could be moving or the spaceship could be
204 | P a g e

moving - it doesn't matter; the high velocity screws up measurable quantities like
length, color, mass, time, etc. The higher the velocity (the closer to the speed of light),
the more screwed up the measurements.
Now you are probably thinking that I have to be making all of this stuff up, right?
This stuff can't really be happening, can it? Does time really pass more slowly on a
fast moving rocket compared to the passage of time on the Earth? Actually, this does
happen - and people have checked it out. If you were to put a clock on the space
shuttle and let it go around the Earth for a few days, it would not keep time as well as
a clock that stayed on the Earth. Even though the space shuttle doesn't go anywhere
near the speed of light, it does go fast enough that the effect can be measured on very
accurate clocks. These things have been measured - people have put clocks on jets and
checked them out. Also, when particles (like electrons) are accelerated in physics labs,
their masses appear to change. This isn't science fiction - this stuff does happen.
For the Special Theory of Relativity, just remember, only light goes at the speed of
light, it always goes at the speed of light, and if an object goes faster and faster, its
characteristics as we measure them will appear to be bizarre. If someone were on an
object that is moving very fast, they would think that our characteristics or that of
anything not on the fast moving object were screwed up - it's all relative.
Believe it or not, that was the easy theory - now it is time to tackle the much more
complex General Theory of Relativity. Actually it won't be so bad, since we're
skipping all of the math that goes with the general theory - that's what makes it really
nasty. Let's start off with what the General Theory of Relativity does - it picks up
where Newton's theory of gravity left off. I'll bet you didn't think that the law of
gravity needed any picking up. For the most part it doesn't - the law of gravity as
Newton formulated it works just fine for most things. It is in situations where it
doesn't work well that Einstein's formula needs to pick up the slack. How does the
General Theory do that? It sort of re-defines gravity. You might think of gravity as a
kind of rubber band or a glue that makes things stick together (have you ever thought
about what actually makes you stick to the surface of the Earth?). Einstein's General
Theory re-did gravity by looking at how space (and time) are distorted by mass. Also,
if you have distorted space, then the way that mass travels through it will be effected.
Basically, mass distorts space, and space effects mass (and anything else within it,
like light).
Wait a second, what is this I am talking about? Is distorting space like warping space?
It is! How do you distort space? You are familiar with three dimensional space. Any
location can be referenced by three coordinates - you can call them x, y, and z, or
perhaps latitude, longitude and elevation. No matter how you do it, any location in the
Universe can be designated by three dimensional coordinates. Space is three
205 | P a g e

dimensional. Let's distort it - warp it, bend it, and stretch it like silly putty. When
space gets distorted, it gets warped into another dimension that we can't see. Let's call
this the fourth dimension - this is a dimension of space; it isn't time, that's something
else. Where is this fourth dimension? I don't know; I can't show you where it is since
we (humans) are only able to exist in and visually comprehend three-dimensional
space. You can't point to it, since you can only point in 3-D space, but it is does exist.
Let's see if we can simply things. Think how a flat surface (a two-dimensional
surface) can get warped in such a way that it bends into a third spatial dimension.
That's how you can have a flat tortilla (two-dimensional object) curved up to form a
three-dimensional object (a taco). If we were tiny two-dimensional creatures living on
a tortilla, and we were only aware of two dimensions (we could only experience our
world in two dimensions), we would not be able to see that the tortilla has been
curved into a third dimension, since our senses wouldn't be able to perceive three
dimensions. Even if we could "see" the three dimensions of the taco shell, we would
be so small compared to the curvature that it would be difficult to detect. This is
similar to why you can't easily see that the Earth is a sphere, but that it looks fairly flat
to your eyes (after all, we are in Iowa). Going back to your Universe, three-
dimensional creatures, like you, me and Einstein, can't see the fourth dimension that
space is being distorted into. No matter how badly distorted the space near you is, you
can never actually see its curvature. It is also difficult to draw four-dimensional space
- frankly, we can't do it. In order to describe warped space we like to simplify things
and just use two-dimensional analogies, which we can
warp into three dimensions.
Figure 8. A 2-D analogy to warped 3-D space. The top
view is undistorted 2-D space. The two lines (red and
blue) are straight, and objects moving on them will travel
the same distance. Placing a mass in this space (middle
picture) causes the space to distort around the mass. The
larger the mass, the greater the distortion. The red and
blue paths are now no longer equal in length. The blue
path is slightly longer. In the last case, the distortion is
very extreme due to a very massive compact object. Now
the differences in the lengths between the red and blue
lines are very extreme. It is even possible that the blue
path has no end but continues down the hole, never
ending.
Let's start out with a 2-D universe which doesn't have any
distortion, so that it is flat - pretty dull and boring, eh?
The rules of General Relativity say that if we were to add
206 | P a g e

some mass to this space, it would get distorted - more mass, more distortion. Let's do
that.
Here's the nifty stuff - far away from the mass you do not detect any distortion. The
space is just as flat as if you didn't have any mass. As you get closer to the mass, you
will experience unusual effects due to the warping of space. Let's say you have two
ants in a race in this 2-D universe. They are going to race across the space and travel
along straight lines. One will be traveling closer to the mass than the other (one will
be along the red line, the other along the blue line in Figure 8). How does the race
end? The ant that traveled near the mass had to actually travel along a greater path,
since that ant's space was distorted. He lost the race. The weird thing is that he didn't
even know that his path was distorted - being a 2-D creature, he could not see the
curvature and the line looked straight and normal to him. If the ant had a flashlight
(and what well equipped ant doesn't have a flashlight?) then the light from this
flashlight would travel along the curved space, but to the ant the beam would be
shining in a straight line, as it should. Light, like the ant, must travel in the space that
comprises this universe, so it will travel along what is really a curved path. Even
though light travels in a curved path, it will always appear to be straight as seen by the
inhabitants of the space.
If you really want to see major distortions you'll need to add a large amount of mass.
You could also concentrate the mass into a really small space - make it very dense.
This is such a strange concept, this warping space stuff, that it might be hard to
believe. Believe it or not, any mass will distort space - you are distorting space, the
computer is distorting space, the building you are in is distorting space, and even your
nose is distorting space (not a great deal, but it is still distorting space).
This sounds like a bunch of science fiction; does this stuff really happen? To test this
out, you need to have a large mass. The largest mass object in our neighborhood is the
Sun, and we know that it distorts space. One way that this is apparent is in the way
that Mercury's orbit about the Sun is gradually altered (orbital precession is the
term). When Mercury gets really close to the Sun (when it is at perihelion), it is in
slightly distorted space. Its orbit is shifted a little bit. Over the years this alteration to
the orbit of Mercury was noticed by astronomers. Newton's laws (and Kepler's laws)
could not explain this apparent aberration in the motion of Mercury, and this caused
some people to suspect that there was another planet close to the Sun that was
messing up Mercury's orbit. This isn't actually the case, but that didn't stop people
from trying to find the mythical planet Vulcan (yes, they really did refer to it as
Vulcan). Of course, when Einstein came along with his General Relativity, this pretty
much explained the problem with Mercury.
207 | P a g e

Figure 9. The way that the space
around the Sun is warped. In the top
image there is no warped space. In
this case, the stars located behind the
Sun could not be seen on the Earth,
since the light from these stars will
never reach the Earth. This is not the
case, however. The lower image
shows how the space actually alters
the path of light, so that the stars are
visible from the Earth. However,
since we can't "see" the warped space, we think the stars are actually in the wrong
positions (the dark blue spots). The positions of the stars that are lined up with the
Sun appear to be off their proper locations by a small amount, due to their light
traveling along warped paths.
Is that not enough for you? Here's another instance. Let's say that you have some light
traveling near the Sun (light which came from other stars). The space near the Sun is
distorted, so the path that the starlight takes will be influenced by this distortion. This
was demonstrated soon after Einstein published his theory by astronomers viewing the
Sun during a total eclipse. During the eclipse, the locations of stars around the Sun
were noted and compared to their normal locations. The star locations were off
slightly due to the distortion of space near the Sun. Of course, when we see these stars
during an eclipse, we presume they are in the direction in the sky that we see them in,
in other words, straight away from us. But those direction are the wrong directions for
the star's true location. If space were not distorted by mass, we would always see the
stars in their normal locations and Einstein would probably have never gotten that
really cushy job at Princeton.
The General Theory of Relatively relates the masses of objects to the effect they have
on space. An object like the Sun has measureable effects as described (and tested
many, many times) above, but the Earth also warps space. A probe studied the effect
of the warping of space by the Earth recently provided evidence that supports the
theory. You can learn about the probe and its measurements here.
Einstein's General Theory helped to fill in the gaps that Newton's law of gravity
couldn't cover, especially the cases involving extreme masses. It can also be thought
of as a way to actually cause gravity. The Earth orbits the Sun due to gravity, right?
What if you consider the fact that the Sun is warping space and the planets move in
that warped space in the only way they can, like a marble rolling around the inside of
a bowl? - the orbits of the planets can be explained by having the Sun's warping of
space hold the planets in their own areas of curved space. Near the Sun the "bowl" is
208 | P a g e

steeper, so the orbits are smaller and the planets have to move faster so they don't fall
in. Further away, the "bowl" is less steep, so the planets don't have to move so fast.
Newton's and Kepler's laws can describe the motion, but Einstein's curved space can
give us the cause of it. Whether this helps you in understanding gravity or not is up to
you; I'm just telling you how it works. I should also mention that one of the other
aspects of the General Theory of Relativity is what it does to time - yes, you got it,
time passes more slowly near massive objects than it does further away from them. It's
another victim of curved space!
That's enough wandering down the confusing path that Einstein gave us - let's get
back to black holes. What can become a black hole? The only ingredient that you need
is mass, so anything with mass could become a black hole. This includes stuff like the
Earth, a pencil, John Goodman and so forth. The only reason that these things don't
normally become black holes is that you have to make them dense enough (compact
in size) so that their escape velocities are equal to the speed of light. To do that to the
Earth you have to crush it down to a size of less than a centimeter. I suppose a
chocolate eclair could push John Goodman over the edge, but I could be mistaken.
That's one characteristic of a black hole - it has mass. Does it have anything else?
How about size or shape? Technically it can't have any size, since if it did have a size,
that would mean it isn't entirely collapsed down to a black hole. That's sort of a
confusing argument, but if you don't think about it too much, it does make sense. For
something to become a black hole it must collapse down to such a small size that it
can't hold itself up, so it has no size. This means that the mass has to be crushed down
to an infinitely small point - a singularity, which has no size but does have a
measurable mass. If a black hole didn't collapse down to a singularity, then there must
be something preventing the collapse - but there is no known force that can overcome
the huge gravitational collapse that we run into in these extreme conditions. Black
holes have no measurable radii, but do have measurable masses. Of course, if an
object has mass, then it has a gravitational pull on anything around it. Now contrary to
what you may think, black holes are not magical vacuum cleaners that travel though
the Universe sucking everything in sight down their gullets. They have to obey the
laws of physics just like everything else in the Universe, so you can determine how
much gravitational pull a black hole has on objects near it. For objects further away,
the gravitational pull is much less (just like it is for anything in the Universe). At large
distances there would be no problem in escaping from a black hole, so if you ever do
run across one in your travels, don't go in for a closer look - you need to stay far away,
where the gravitational pull is low. As you'll see, if you do get too close, there will be
nothing to save you.
A black hole has to obey the laws of physics, so there is that rule about escape
velocity - the speed needed to escape from an object - that you have to watch out for.
209 | P a g e

If you are too close to an object, you need to travel at a very great speed to escape. If
you start out at a point further away, the velocity you need to escape is much less. In
the case of the black hole you can use this to define a "boundary" for the black hole.
There is a distance from the black hole (the singularity) where the escape velocity
equals the speed of light. This distance from the singularity is known as
the Schwarzschild radius (R
s
). This is often referred to as the "radius" of the black
hole, since its true physical radius has no size. You can also think of this radius as the
"point of no return." If you were closer to the singularity than R
s
you would have to
travel at a speed greater than that of light to escape, and since nothing can travel faster
than the speed of light, nothing can escape from a location within the Schwarzschild
radius, not even light (that sort of explains the name "black" hole, doesn't it?). At
distances greater than R
s
, the gravity is less, so the escape velocity is less than the
speed of light, so things like rockets could escape. To give you an idea of how big this
"point of no return" is, if you were to make the Sun into a black hole, the R
s
would be
about 3 km.
The term event horizon is also used to describe the Schwarzschild radius, since that is
the closest you can see an "event" happening near a black hole. Anything that happens
within the event horizon is not visible since no information can get out of the black
hole at that distance. So you could think of it as the last "event" location. In many
ways black holes are like Las Vegas - whatever happens in a black hole, stays in the
black hole.
The value of the R
s
is proportional to the mass of the black hole. A 10 solar mass
black hole has a R
s
of 30 km, a 1000 solar mass black hole has a R
s
of 3000 km, etc.
The formula for a black hole's Schwarzschild radius is

Now I suppose you have been wondering what it would be like to fall into a black
hole? Admit it; you've wondered that for a long time. Let's find out.
First of all, we'll start at a great distance from the black hole, where the effect of it on
the space around it (the strong spatial distortion as defined by the laws of General
Relativity) is pretty low. Far away from the black hole, space and time are not
distorted, so things would not appear unusual. We'll leave someone (perhaps a friend
of yours) at this location so they can watch you approach the black hole (what a swell
friend they are - actually, if they were really a good friend, they'd volunteer to go into
the black hole). For your journey into the black hole, you'll have the best equipment -
a flashlight and a watch. Obviously no expense was spared for your expedition. You
have the flashlight so that you can signal your friend in the spaceship that you are
210 | P a g e

okay, and you agreed that you would signal every 10 seconds (that's why you need the
watch). Off you go toward the black hole. Now one of the aspects of General
Relativity (and Special Relativity) is that different people in different physical
environments (moving and non-moving, and warped space and non-warped space)
will measure things differently. There will actually be two different views as to what
happens to you as you fall into the black hole. First we'll look at what your friend (or
so-called friend who's safely in the spaceship) sees from their position far from the
black hole. As you get closer and closer to the black hole, your friend starts noticing
some weird things. One is that the signals from the flashlight will become less and
less frequent. Perhaps your watch is broken and is running slower than it should - but
when the signals start coming at intervals of once a minute, once every 5 minutes,
once every half hour, once every hour, once every day...obviously this is not just a
case of a broken watch - somehow time is running much more slowly for you (or at
least that is what your friend sees). Why? This is one of the connections of space with
time - that the space distortions around the black hole also affect the passage of time
for different observers. Another thing they'll notice is that the light from the flashlight
is getting redder and redder as you approach the black hole. Why? Light is feeling the
pull of gravity, but since it can't go more slowly, as other things would, it instead loses
energy. Lower energy light has a longer wavelength - so it gets redder and redder as
you get closer and closer to the black hole. Eventually the light will be at such a long
wavelength that it will not be visible to the eye, since it will be in infrared, microwave
and eventually radio wavelengths. In spite of these hardships, your friend (who is
probably having milk and cookies in their cozy spaceship) is still able to observe you
as you approach the black hole. The time delay gets so extreme that over their
lifetime, they don't see you moving any closer to the black hole. Actually, over many
generations of people sitting in the spaceship, watching you head toward the black
hole, it will look like you are basically frozen near the Schwarzschild radius. To
everyone in the Universe, it looks like you don't even get to go across the
Schwarzschild radius. As perceived by the rest of the Universe, you're "frozen" on the
edge of the black hole.
What will you see? In a way, you see sort of the opposite of what your friend sees. It
looks like their clocks are running too fast - and they are moving too fast. Eventually,
in what seems like a moment to you, someone on the spaceship will be born, grow old
and die right before your eyes. You'll be able to watch the entire life cycle of stars and
see them become supernovae. Not only will it look like the passage of time in the
Universe is strange, but also the Universe itself will. You will start to see things that
were originally not in your line of sight (in front of you). Stuff that was over to the
right or left of you would appear to be in front of you - a kind of funneling of light
near the black hole. You won't have to turn around to view the entire sky, since it will
all be compressed down into a size that you can see. You won't be able to enjoy the
211 | P a g e

cosmic light show for long, since you will cross the Schwarzschild radius, fall into the
singularity, and be crushed down to nothingness - ouch.
Figure 10. An astronaut heads toward a black hole. At a
great distance there is no problem. As the astronaut gets
closer, it gets very uncomfortable, and the tidal forces of
the black hole will stretch him out of shape. It's not the
most pleasant way to go, but what can you do? It should
be noted that the diagram isn't to scale, since the
Schwarzschild radius of the 10 solar mass black hole is
30 km (diameter of 60 km), and the stretching effects
don't get really bad until the astronaut is a few thousand
kilometers from the black hole.
That's sort of a sad end, eh? Actually, the end would
probably come well before you were anywhere near the
Schwarzschild radius. Depending upon the mass of the
black hole, you could be ripped apart by tidal forces.
What do I mean by this? Let's just put it this way: if you
were ever to travel into a black hole or into the general
direction of a black hole, you would want to travel
sideways, not head-first or feet-first. What difference
would that make? Just remember the laws of gravity;
anything that is closer to a massive object (be it a black hole, a planet or whatever)
feels a stronger gravitational pull than something further away. In the case of a black
hole the difference in the pull that your feet would feel and the pull that your head
would feel would increase until the part that was closest to the black hole would want
to separate from the part that was further away, since it is getting pulled more. You
would be basically ripped apart. This is the effect of tidal forces. The less difference
in distance between the various parts of your body from the black hole, the less the
difference in the pulls those parts would feel. Remember, never go head- or feet-first
into a black hole, though in some cases the pulls would be so extreme it wouldn't
really matter how you went in, you'd still be stretched like silly putty.
Is a black hole really black? No, not really, because we don't know what color it is.
Remember, what your eye perceives as color is just a certain wavelength of light.
Light can't get out of a black hole, so the color of a black hole remains unknown. The
absence of light from an object would result in the object looking black.
Could a black hole be a passage to another Universe, or just a shortcut into another
part of our own Universe? This is sort of the realm of science fiction, but not entirely.
We know that black holes severely distort space, so it is possible for it to stretch space
212 | P a g e

from one location to another in our Universe. In a way, a tunnel could be created
between different parts of the Universe, though I wouldn't want to travel in that
tunnel. Remember, there is a black hole in that tunnel, so it would be easy to get in but
not to get out. Such passageways have been proposed by theoreticians and are
possible in theory, but whether they exist in reality has yet to be determined. Probably
the best kind of tunnel is one that doesn't contain a black hole - though how the tunnel
can exist without the black hole is a bit tricky. Just having a nice tube of warped space
would be useful for traveling great distances in a short period of time.
Figure 11. The set up for a black hole
binary. A star, generally a large supergiant,
is near an accretion disk and a black hole.
The size of the black hole (maybe only a few
10s of kilometers) is much smaller than that
of the accretion disk (a few 1000s or
100,000s of kilometers in size), so it would
be very difficult to see, even if it were
visible to telescopes. The accretion disk is
made up of material sucked off from the
normal star (mainly hydrogen) and gives off x-rays. Credit: Space Telescope Science
Institute, NASA.
Are there any black holes out there? How can you tell? Remember, space is black and
black holes are black, so how do you find one? You can't see it directly, but you can
see it indirectly. A black hole has mass, so it is able to influence objects near it. You
would want to look for an unusual binary system. How can you tell if a binary system
is unusual? What difference would it make if one of the stars in the binary system
were a black hole? First of all, you have to remember about the extreme gravity. Due
to this, the black hole may have the tendency to pull material in from the other star.
As this material falls into the black hole it will form into an accretion disk(remember
how this happens around a white dwarf before it goes nova?). The disk material will
be compressed, heated up and basically crunched up even at a great distance from the
black hole's Schwarzschild radius. This material (mainly hydrogen and helium from
the other star) is heated up to such a high temperature that it will start to give off x-
rays. These x-rays are what make the binary system unique. If you were to look at a
black hole binary system with a regular visible light telescope, you would probably
only see the other star - the normal star. The fact that there is another star in the
system, a star that gives off x-rays (actually, the accretion disk around the star is what
is giving off x-rays), will make the binary system something that astronomers would
notice.
213 | P a g e

You should be careful about saying whether you have a black hole or not, since x-rays
can also come from a binary system that contains a neutron star. The only way that
you can figure out which it is (black hole or neutron star) is by determining the masses
of the stars in the binary system. And how do you determine the masses of stars in a
binary system? With Kepler's 3rd Law, of course (assuming you can determine the
size of the orbit and the period of the orbit)! If the mass of the star in the accretion
disk is much greater than about 5 solar masses, odds are it is not a neutron star but is
instead a black hole. In 2002, there were reports of two very low mass black holes
found in separate binary systems; one was only about 4.25 solar masses, and the other
was around 5.25 solar masses. These were both discovered in x-ray binary systems
with orbital periods of only a few hours!
Another thing that astronomers can look for is how material falls in towards the
suspected black hole. When material falls in and lands on a neutron star, it will give
off a particular x-ray signal that indicates that it hit something. On the other hand,
when material falls past the Schwarzschild radius of a black hole, it just "disappears" -
it does not give off a signal indicating any sort of surface or barrier, since there is not
one present. Astronomers have carefully looked at the x-rays coming from stuff
falling onto neutron stars and stuff going towards a suspected black holes and found
very distinct differences between the two - and these matched the models that are
based upon the physics involved in these
situations.
Figure 12. Another view of a black hole in a
binary system. If you click on the image you'll
see a small movie showing material flowing
into a black hole. There is the accretion disk
around the black hole and also bipolar
outflow - material being ejected from above
and below the black hole's location. This
material isn't coming from the black hole
itself but is from the accretion disk. Image
and movie from NASA/Honeywell Max-Q digital group/Dana Berry.
There are actually quite a few interesting neutron star and black hole binary systems
out there. This is only a small sampling of some of the systems -
PSR 1913+16 - This is a pulsar binary system containing two neutron stars - only one
of the neutron stars is observed to be a pulsar. The neutron stars are in a very close
orbit, so they are moving very fast and this causes the pulsar signals to be distorted.
Actually, this type of distortion would happen even if the pulsar were in a regular
binary system. What is different about this system is that the distortion of the
214 | P a g e

pulsation periods is also effected by the distortion of space near the neutron stars
(General Relativity again rears it ugly head). There is so much distortion of space near
this system that the two neutron stars warp each other's space, causing their orbits to
continually change. The discovery of this system and its motion over the years
provided more support for Einstein's General Relativity (and earned the discoverers
the Nobel prize).
PSR J0737-3039A - Like the previous example, this is another binary neutron star
system, and in this case both of the objects are pulsars. The orbital period is only 2.4
hours - yes, that's what I said, 2.4 hours. And of course there are also the periods of
the pulsars (due to their rotations). One of the pulsars is a milli-second pulsar,
spinning around 40 times every second. This is common for binary systems, since any
mass transfer will speed up the rotation of degenerate objects. The scientific
community was very excited about these two since like the previous example they can
be used to test Einstein's theory of General Relativity. There are also all of the neat
possible interactions of the two stars' magnetic fields. These stars will
eventually merge together and likely end up as a black hole, but that wont happen for
another 85 million years.
Cygnus X-1 - This is a very strong x-ray source. However, when you observe it with
a regular telescope you see an O type star (named HDE 226868 - isn't that romantic?).
The O star is not a source for x-rays, so where are those coming from? You can also
see, by looking at the spectrum of the O star, that it is in orbit about something else -
something with a mass of about 10 solar masses. If this object were a regular star it
should be easily visible. However, it is not - all that is seen is the regular O star. What
is 10 times the mass of the Sun but appears to be invisible? - most likely a black hole
(it is thought to be too massive to be a neutron star). Cygnus X-1 is the best example
we have of a black hole candidate - remember, we only think that it is a black hole; we
can't see it directly.
215 | P a g e


Figure
13. To the
left, a view
of HDE
226868, a
normal star
that is also
a x-ray
source. The
x-rays aren't
really from
the visible
star but
from an
accretion
disk around
a possible
black hole.
This is one
of the best
cases for the
existence of
a black hole
in our
galaxy.
Images from
NASA.

Remember, it isn't the black hole that is giving off the energy, but the accretion disk
around it. The disk is usually much larger than the black hole itself - a 10 solar mass
black hole has a Schwarzschild radius of 30 km, but an accretion disk around it can be
over a million km in size. Even if the object itself is not visible, its influence on its
surroundings gives it away. There are quite a few objects out there that we think are
black holes, but we can't really be 100% sure, since there are always possibilities that
these objects are something we haven't yet thought of. Until we know for sure, the
really big ones in x-ray binary systems are classified as probable black holes, while
those that are less massive are likely neutron stars.
XTE J0929-314 - This is another pulsar, this time in a binary system where it is
destroying its companion star. In this case, the pulsar is both sucking material off of
the companion star and the strong radiation from the accretion disk is blowing
material off of the companion star. These objects are so close that it takes less than
one hour for them to orbit one another. The end result is a very rapidly spinning pulsar
216 | P a g e

(185 rotations/second) and a companion star that has evaporated down to the size of a
planet (the companion's mass is only about 10 times Jupiter's).
Figure 14. A diagram showing what is
happening in XTE J0929-314. The pulsar
buried in the accretion disk is pulling material
off of the companion star and also blowing
material away. The companion star's mass is
now too small for it to be much of a
star. Ultimately there will only be a neutron
star left. Image courtesy NASA.
Some astronomers have been looking for some rather amazing effects caused by black
holes, including the triggering of a supernova in a white dwarf. Computer simulations
of what happens to a white dwarf that gets too close to a black hole did not just result
in the destruction of the white dwarf (which is what was expected) but the ignition of
the degenerate material within it due to the strong tidal forces. Just check this article
to see results of the study. And even having a normal star getting too close to a black
hole can also result in violent supernova-like explosions. It is possible that some of
these types of events may have already been observed, as is described in here.
Other Bizarre Objects
As with many things in astronomy, this is not the last word about how the most
massive stars die. Neutron stars and black holes are pretty tricky to observe, so it
should not be surprising that there are some other things out there that have only
recently been discovered (due to advanced telescopes and observing techniques). One
of these weird things is called a Magnetar. This is basically a neutron star with a
souped up magnetic field. Now you'll recall that the magnetic field of a neutron star is
pretty intense, so having something even more powerful is pretty amazing. To give
you an idea of what's involved, here are some facts. The magnetic field of the Earth is
around 0.5 Gauss (that's the unit of measure for magnetic fields). A typical
refrigerator magnet has a strength of about 100 Gauss. The Sun's magnetic field has a
strength of around a few thousand Gauss (remember all those sunspots). A neutron
star has a magnetic field with a strength of about 10
12
Gauss. A magnetar has a
magnetic field strength about 100 times greater than that, around 10
14
Gauss. In June
2002, the strongest magnetic field for one of these things was announced - being over
10
15
times as strong as the Earth's magnetic field. The magnetic field on a magnetar is
so strong that it can cause the crust of the star to buckle and pop every once in a while.
This would produce what is basically a starquake, which can be a source of gamma
rays. At this point magnetars are pretty rare, since they appear to slow down relatively
217 | P a g e

quickly and lose energy quickly as well. Only a handful of these beasts have been
discovered up to this time.
The most recent discovery of a strange object is something that can be called
a Strange Star - really, that's what it's called. You can also call them Quark Stars.
What's that? You remember that white dwarfs are stars where the atoms are packed as
tightly as possible and neutron stars are where the neutrons (a piece of an atom) are
packed as tightly as possible. If you break a neutron (or a proton) down, you'd find it
is made up ofquarks. What if you have an object that is so compressed that the pieces
of atoms (like neutrons and protons) are broken down into their smaller bits and those
pieces are what make up the star? In that case you would have a Quark Star. Two
groups of astronomers in 2002 announced the discovery of possible quark stars -
objects that are observed to be too small and cool to be neutron stars. While a typical
neutron star should be about 20 km in diameter, these were even smaller.
Magnetars and Quark stars are both extreme cases of neutron stars, though it appears
that they are pretty rare at this point. There is some debate as to whether they actually
exist; that perhaps the astronomers got their numbers wrong. You'll just have to wait
and see.
Figure 15. A size comparison
of the theoretical size range
for neutron stars (17-23
miles), the theoretical size
range for quark stars (10-17
miles) and the recently
discovered possible quark
star, RX J1856-3734. The
width of this object (around
12 miles) is similar to the
distance from Cedar Falls to
Waverly. Its mass, though, is
likely in the range of 1.5 solar
masses.
Objects like magnetars can be incredibly violent, as witnessed in the spectacular
eruption that occured on December 27, 2004. You remember that, don't you? The
energy from the magnetar SGR 1806-20, which is a mere 50,000 lightyears away was
observed on that day, and this energy release, which lasted about 10 seconds, was
brighter than the Full Moon! Surely you remember seeing that? Well, no you didn't
see it, since the energy was mainly in the form of gamma-rays. But even though the
eruption wasn't visible to our eyes, it was detected by the Swift satellite as well as
218 | P a g e

other telescopes. The energy was equivalent to the energy released by our Sun over
150,000 years. Only recently have astronomers detected some bursts that are even
more powerful than this one (see the article about it here). Most likely this will be one
of those records that will be broken over and over again.
Gamma-ray outbursts are so energetic that they can be observed from great distances,
and as was described previously, we are still not certain how they are produced. In
September of 2005 a burst was recorded which was later discovered to have occured
from an object that was 13 billion light years away - and in case you were wondering,
that is about one of the most distant objects ever observed. This particular outburst,
called GRB 050904, was detected by the Swift satellite, and was a very long duration
burst. Generally we think that this gamma-ray burst happened when one of the very
massive, early stars died and formed a black hole - possibly one of the earliest black
holes created in the Universe!
There are some scientists that think that explosions from magnetars or supernovae
may have altered or in some cases wiped out life on the Earth in the past. Here is a
recent study by some folks that link a mass-extinction that happened 450 million years
to a large gamma-ray burst. Such events can severely damage the ozone layer and lead
to strong disruptions to the food chain. Of course, this is only speculation, but it does
make you wonder what would happen if one of these things went off nearby today.
The area of gamma-ray astronomy is relatively new and astronomers are not exactly
sure what the cause or causes of these gamma-ray bursts actually are. As mentioned
previously there are several possible causes, including a massive hypernovae
explosion, or the collision of objects like neutron stars, or black holes and neutron
stars (click here to see an animation showing a black hole shredding and gobbling up
a neutron star). Generally it is thought that the end result of gamma-ray bursts are
black holes, however this is something that astronomers are still working on, so stay
tuned.

Now that you've read this section, you should be able to answer these questions....
What are the characteristics of a neutron star and a pulsar?
What makes a pulsar produce "pulses"?
What evidence is there that connects pulsars to neutron stars and supernovae?
What are the basic principles for the Special Theory of Relativity?
How do these principles appear to go against conventional physics?
What are the basic principles for the General Theory of Relativity?
How does General Relativity explain gravity?
219 | P a g e

The Milky Way Galaxy

What's covered here:
What are the characteristics of our galaxy, the Milky Way?
How do we measure the various features of the Milky Way?
What is the center of the galaxy like?
How far are we from the center?
How does the galaxy move?

Time to move onto the next level - galaxies. Galaxies are the objects that contain
stars, the gas clouds out of which stars form and all the other stuff in our
neighborhood. In a way, galaxies are like states, while the stars in them are cities
within the states. Perhaps hog confinement lots can be thought of as the gas between
the stars? Any ways, before we look at galaxies in general, we need to look at the
galaxy that we know best, our own, theMilky Way Galaxy. Yes, that is really its
name - and not just the name of a candy bar.

Figure 1. A view of the Milky
Way stretching across the
sky. This picture was taken
using a fish eye lens on a
camera. A comet is seen in
the lower part of the picture.
Image by G. Garradd of
Australia.
First of all, what does it look
like? How big is it? Where
are we located in it? Are we
on the edge of it, in the
middle, or somewhere in
between? How can we tell?
What do you see when you
look at the Milky Way?
When you go outside on a
clear, moonless night (and
you are in a site far away
220 | P a g e

from the city), you can clearly see the Milky Way. It looks like a band of fuzziness
that appears to stretch across the sky, though it isn't always easily visible. The best
times of year to see it are the winter and summer months. This is when it is high
above the horizon. Why does it look like that? That is where more stars and other
material are located, so that gives it that fuzzy appearance. The fact that it is located in
a band across the sky also gives us a clue about the Milky Way's shape - obviously it
isn't a sphere - at least not where we're located, but is probably closer in shape to a flat
object. In the old days, people thought that if they could measure the depth of the stars
in the band they could get an idea about how big the galaxy is and possibly also our
location in it. To your eyes, the distribution of stars in the sky looks pretty uniform.
Could you point to any specific site and say, there is the center of the galaxy?
Probably not, but if you were to assume that stars are fairly evenly spread out, you
could count the stars in various directions and use that information to try to estimate
our location in the Milky Way. In directions where you see many stars, you would
probably believe that the galaxy stretched out a long way in that direction. In those
directions where there are very few stars, you might think that the extent of the galaxy
there is not so big - seems pretty
straightforward, eh?
Figure 2. An early image of the galaxy
based upon counting stars. This shows
the Earth (and Sun) located near the
center of a flattened distribution of stars
- not a good model of the galaxy.
What sort of a picture of the galaxy do we get when we use this method? We get a
galaxy with us located near the center of a small, flat disk (only a few hundred light
years in size). Is this correct? Are we located near the center of the galaxy? That
seems pretty unlikely - why would we be at the center of everything? I mean, we are
pretty important, right, but are we really that important? NO! If this isn't the correct
answer, what went wrong? Do we need to count the stars again? We could, but no
matter how many times this experiment was done, the result was the same - we end up
being in the middle of a relatively small galaxy. Why do we keep getting this result
for the shape of the galaxy? What prevents us from seeing the actual shape? Just a
whole bunch of gas and dust, that's what! It is sort of like a fog that limits how far we
can see. It is a bit of a paradox - to see how far the disk of the galaxy extends, we have
to look in the disk of the galaxy, and this is where all of the stuff (gas and dust) is that
is blocking our view. We don't actually see very far when we look in the main part of
the Milky Way, so when we map out our position using the stuff we can see with our
eyes, it looks like we are near the center. Obviously someone needs to figure out a
better way to do this.
221 | P a g e

If we want to try to locate our position in the galaxy, we need to use a method that
will not be affected by the gas and dust in the plane of the galaxy. We need to look for
things that are located away from the plane (disk) of the galaxy and that we can get a
reliable distance from. What sort of objects are out there that allow us to do this?
We're going to have to make a little detour back into the realm of stars, in this case
some rather interesting Giant and Supergiant stars. You thought we were all finished
with stars - no, not yet.
There is a region on the H-R Diagram where a star's structure (temperature, density,
and pressure) will be slightly bizarre. These stars tend to not be in Hydrostatic
Equilibrium. What does that mean? Do you remember what Hydrostatic Equilibrium
is? It is the balance of gravity (pulling inward) and gas pressure (pushing outward).
When a star is in balance, then the star is stable - if you kicked it, it would wobble a
bit but eventually settle down. When a star is NOT in Hydrostatic Equilibrium, it
won't settle down. In fact, if you were to kick such a star, it would not settle down but
would expand and contract, expand and contract, and so on. A star that is not in
Hydrostatic Equilibrium is like a faulty pressure valve. It will keep over compensating
and under compensating, allowing in too much or too little air, or whatever the faulty
valve is supposed to regulate. In the case of a star in this situation, the outer layers
will expand and contract in a periodic manner as it blocks up and then releases energy.
If you're changing the radius of the star during this expansion/contraction process,
then the observed characteristics of the star will also change. Things such as
luminosity (brightness) and temperature will be affected by these pulsations. The neat
thing is that these variations are very well organized (periodic) - they occur in
generally very well behaved cycles. All this occurs while the star is located in a region
of the H-R diagram known as the Instability Strip.
The Instability Strip is located in the upper part of the H-R diagram, so the stars that
tend to pass into it are either Red Giants or Supergiants. During a star's Red Giant
phase, a star could cross the Instability Strip once, twice or more times depending on
the type of evolutionary path it takes, which of course depends upon the mass. Once a
star does enter the Instability Strip and starts to pulsate, we call it a variable star.
There are actually many types of variable stars out there, but we'll only look at two of
the main types.
Cepheids and RR Lyrae Stars.
Figure 3. Light variation of a Cepheid
variable. It is plotted so that two cycles of
pulsation are shown.
222 | P a g e

Cepheids are high mass pulsators and are often classified as Supergiants. They are so
big that they tend to have long pulsation periods; after all, it takes a long time to move
all of that material. They tend to have pulsation periods between 2 and 40 days,
though some can have periods that are much longer. The masses of these stars are
typically between 3 and 10 solar masses. To make matters a bit complicated, there are
actually two types of Cepheids, which are easily distinguished according to the
amounts of heavy elements (metals) in their compositions. The not so original names
are
Type I Cepheids, also called Classical Cepheids, are metal rich - in other
words, chemically similar to the Sun.
Type II Cepheids, also called W Vir stars, are metal poor (deficient), which
means they have fewer of the heavy elements in them, like iron, than found in
the Sun.
Type I Cepheids also are a bit more luminous compared to the Type II Cepheids, and
some Type II Cepheids can also have really short pulsation periods (less than a day).
Cepheids are very nifty stars and provide us with some useful information. This is
because of two aspects -
They are very bright, remember these are Supergiants, so it possible to see
them in other galaxies over very great distances
Their average luminosities (Absolute Magnitudes) are directly correlated to
their pulsation periods.
What does that mean? It means that if a Cepheid takes a long time to vary its
brightness (has a long pulsation period), it is more luminous than a Cepheid with a
shorter period. This concept is known as the Leavitt Law, named after Henrietta
Leavitt and this law is possibly the best thing since sliced bread. Why? I'll get to that
in a minute.
Figure 4. The Leavitt Law for Cepheids is shown. The
longer the pulsation period, the brighter the star is on
average. We have to say "on average" since these stars
vary their brightnesses, so the average brightness is
along the y-axis.
For these reasons, Cepheids are amongst the most
important stars in the Universe. Why? You can get
distances to the Cepheids and the galaxies they are
located in - fairly accurate distances, in fact. How? First
223 | P a g e

you need to determine the pulsation period of the Cepheid. This is pretty easy since all
you have to do is watch it for a long time. Once you have that, you go to the Leavitt
Law and determine the expected luminosity (absolute magnitude) of the Cepheid with
that period. By comparing the expected luminosity (absolute magnitude) to the
observed brightness (apparent magnitude), you can determine the distance that the
Cepheid is away from you. Basically, you know how bright it should be, and you
compare it to how bright it appears and the difference of those two is related to its
distance. This is a very easy process and the results are very reliable. Cepheids are
observed in other galaxies, so we can use them to find the distances to those other
galaxies. The only complication is that there are two types of Cepheids, so you need
to know which type you are observing
since one type is more luminous than the
other.
Figure 5. Light variation of an RR Lyra
star (Plural: RR Lyrae). Notice how much
shorter the period is than for the Cepheid.
This is one of the ways that they are
distinguished from Cepheids.
The other type of pulsating star is the RR
Lyra (plural: RR Lyrae) type. These are
small mass variables, generally much less
massive than Cepheids, and even less
massive than the Sun typically, but they pretty much operate the same ways as
Cepheids in most respects. They are less massive, so they tend to have much shorter
pulsation periods, typically less than one day for an entire pulsation cycle. They are
also less luminous then the Cepheids (again, because they are much less massive).
Unlike Cepheids there is no strong period-luminosity relation like the Leavitt Law,
but they all tend to have about the same average luminosity (absolute magnitude).
Basically, they all have about the same wattage (if you want to think of them as light
bulbs). If you find an RR Lyra star, you know how luminous it is - about the same as
all other RR Lyrae out there. As with Cepheids, you can compare the RR Lyrae's
observed brightnesses to their expected brightnesses and use the differences to
determine the distances to these objects. The only problem is that they are only useful
for determining short distances since they are not as bright as Cepheids. While
Cepheids can be both metal rich and metal poor, RR Lyrae stars are pretty much all
metal poor.
Let's get back to the problem at hand, mapping out the Milky Way. How will these
stars help us map out the size of the Milky Way and our location in it? We can use
224 | P a g e

them because they are located in Clusters. You haven't heard about clusters yet? It
looks like we'll have to make another side track.
If you find a group of stars that is physically associated (gravitationally bound), it is
likely that these stars were born at the same time (remember how large scale star
formation works?). This group is what is called a cluster. The really neat thing is that
when these stars were formed, they would all have had different masses (again,
remember how large scale star formation works - if you don't you might want to go
back to that section). You know that different mass stars live their lives at different
rates (high mass, short life; low mass, long life). Even though these stars were all
formed at the same time, they would be in different phases of their lives depending
upon how old the cluster is and how diverse the masses are. By looking at one group
of stars, you could actually see stars in a wide range of stellar life stages. Actually,
clusters of stars are fairly common and there are quite a few star clusters out there.
These clusters fall into two main types.

Figure 6. A typical Open Cluster. Notice
how widely spread the stars are and how
blue many of the stars are. Image
Anglo-Australian
Observatory, Photograph by David Malin.

The Pleiades, also known as the Seven Sisters,
located in the constellation of Taurus. Image
Anglo-Australian Observatory, Photograph from
UK Schmidt plates by David Malin.
225 | P a g e

Open Clusters are also called Galactic Clusters and contain about 1000 stars or so.
There are probably around 20,000 open clusters in our galaxy, and most are found in
the main body of the galaxy (the plane). The extent of a typical open cluster is at most
30 pc. They are called open clusters since that sort of describes how they look - very
wide open. The stars in them are all spread out and in no particular arrangement. The
spacing between the stars is pretty big, with there being less than one star in a cubic
parsec of space. An aspect of open clusters that makes them good objects to observe is
the type of stars that are found in them - mainly young and hot stars that are very
similar to the Sun in terms of composition. Many open clusters have rather bright,
blue stars that make them easy to see. These clusters are chemically similar to the Sun
(metal rich), so if they contain any variable stars, they would have Type I Cepheids.
You have probably seen an open cluster and not even known it - there is a group
known as the Pleiades (also called the Seven Sisters or Subaru) that is visible in the
Winter and early Spring skys. Some people think that it is the little dipper (it isn't),
and it is obviously a grouping. If you were to use some binoculars to look at it, you
would see about 20-30 stars, all of which are part of the cluster. Other open clusters
are also out there, but they
are not bright and obvious
as the Pleiades.
Figure 7. A typical
Globular Cluster, this one
is known as M80 . Notice
how tightly packed in the
stars are, as well as how
yellowish-orangish many
of the stars appear to be.
Image Credit: Hubble
Heritage Team (AURA/
STScI/ NASA).
The other type of cluster is
the Globular Cluster.
Why are they called this?
If you looked at them you
would see a glob of stars
all bunched up in the
middle. Globular clusters contain many more stars than open clusters, typically about
1,000,000 stars packed within an area about 25 pc wide. This is the same amount of
space that an open cluster occupies, but with the globular clusters, there are about
1000 times more stars. That is why these clusters look like globs; all the stars are
226 | P a g e

bunched in so close together that they look like fuzzy cotton balls in a telescope. The
spacing between stars in a globular cluster is much less than in an open cluster - stars
may be only a small fraction of a parsec apart from one another in a globular cluster.
The stars are so tightly packed together that they are in a very well organized system
and appear in a dense spherical arrangement. Globular clusters are found out of the
plane of the galaxy in general and are arranged in a spherical distribution about the
center of the galaxy (this is a very important point, which we'll get to in a bit). These
clusters contain old stars, which are generally deficient in heavy elements. Variables
found in globular clusters include Type II Cepheids and RR Lyrae stars.
How do we know that globular clusters are made up of old stars, while open clusters
have young stars in them? It is actually pretty easy to see. All you have to do is plot up
all the stars in a cluster on an H-R diagram and see where they are located. In an open
cluster, many of the stars are still on the main sequence, and only a small number have
died and can be found in the red giant or supergiant region. You can actually see how
the ages of clusters vary by how many or few stars are still on the main sequence - the
basic upshot is that the shorter the main sequence, the older the cluster. Globular
clusters are so old that they have very few main sequence stars remaining, while their
easily visible stars are all hanging out in the red giant or supergiant area of an H-R
diagram. It is pretty easy to see the relative ages of clusters. To get actual values for
ages, we need to use computer models that can tell us how long it takes for the various
stars of a cluster to die (which, as you remember from the previous section of the
course, is dependent upon mass).

Figure 8. The evolution of a star cluster is shown. In the first H-R diagram (far left),
the stars have just been formed. The high mass stars evolve faster, so they are on the
Main Sequence (red line) before the low mass stars. In the second image, all of the
low mass stars have made it to the Main Sequence, but by this time, the high mass
stars have already started to die and are evolving over to the right. This trend
continues in the next two panels, and as the cluster ages the Main Sequence gets
227 | P a g e

shorter and shorter. By plotting up stars in a cluster it is possible to compare their
locations on the H-R diagram in order to determine the age of the cluster.
You'll notice that both types of clusters have variable stars in them. Clusters can have
either Cepheids or RR Lyrae in them, so we can determine the distances to them. How
does this help? At first we weren't sure, but someone thought that they could use
clusters to determine how far the center of the galaxy is from us. Which clusters
would be the most useful? Open clusters are found in the plane of the galaxy and our
view of them tends to be obstructed by all the gas and dust in the galactic plane.
Obviously we don't want to use them. That leaves us with globular clusters. If you
look at the globular clusters, you can map out their distribution around the galaxy by
determining how far away they are, and we can do that since they contain Type II
Cepheids and RR Lyrae. Someone did just that; an astronomer by the name
of Harlow Shapley (who did the work around 1915) mapped out the locations of
globular clusters and determined their distances using Type II Cepheids and RR
Lyrae. What did he find?
Figure 9. The distribution of globular
clusters as seen by Harlow Shapley. There
is a strong concentration of clusters in the
direction of the constellation of Sagittarius,
and they are quite a ways from the Earth.
When you look in one particular direction
(toward the constellation of Sagittarius),
you'll see more globular clusters in that
direction than in any other. This distribution
indicated to Shapley that the center of the
galaxy was in the direction of the
constellation of Sagittarius. He also saw
how the clusters were spread out in a rather
spherical distribution that radiated from the
direction of Sagittarius, which indicated that the center was not anywhere near us but
somewhere far away toward the direction of Sagittarius. By looking at the typical
distances of the globular clusters in the direction of Sagittarius, he determined how far
away the center was from us (our solar system) and it was much greater than anyone
ever thought possible.
It took quite some time for the rest of the astronomical community to accept Shapley's
model of the galaxy, but now we know that his method was basically correct. Later
observations confirmed his result that we (our solar system) are nowhere near the
center of the galaxy. Nowadays, astronomers use radio telescopes, which can see
228 | P a g e

through the gas and dust at certain wavelengths, to map out the size of the galaxy and
our location in it. Now we can see through the "fog" of the disk and get a pretty
accurate view of the galaxy. It is pretty amazing that Shapley was able to do pretty
much the same thing by using the tools he had and getting pretty much the same result
we get today. We can also compare our own galaxy to others to give us more clues
about the size, shape and content of our galaxy. We now have a pretty good idea of
the arrangement of the stuff in our galaxy, and this is shown below.

Figure 10. Diagrams showing the major parts of the Milky Way. The Sun's location is
also indicated, and the dot that represents the Sun is not to scale. The locations of the
spiral arms are not precise and are open to debate. The distance of the Sun from the
center of the galaxy is about 8000 parsecs (26,000 light years). In the view on the
right, the globular clusters are represented by fuzzy red blobs.
There are several major parts, including the disk, the bulge, and the halo. The sizes
and extents of these areas are shown in the diagram above. We'll look at the various
parts of the Milky Way in the next sections. Just to make life interesting, we'll start
from the center and work our way out.
The Bulge
Figure 11. An infrared
telescope view of the entire sky
showing the locations of the
disk of the galaxy and the bulge.
This view has the center of the
galaxy in the middle. The extent
229 | P a g e

of the bulge show how well named it is. Image is courtesy E. L. Wright (UCLA), The
COBE Project, DIRBE, NASA.
The center of the galaxy is located within the bulge, which, as the name implies,
bulges out from the flat disk shape. Not only is it bulging out, but it is made up of
distinct stars, in this case those that appear reddish or yellowish (relatively cool stars).
We see this kind of coloring in other galaxies also. There is so much junk between us
and the center that we don't have a very good view of the bulge with visible light
telescopes. However, those clever astronomers can use other types of telescopes, and
when they do, they notice that the bulge doesn't really have a spherical shape but is
sort of stretched out, kind of like a cigar, so that the center actually has more of a bar-
like structure. This is kind of important, since it helps us to classify the type of galaxy
we live in (we'll get to galaxy types in the next set of notes). The size of the bulge isn't
exactly known, but it appears to be about 2000 parsecs (or 2 kiloparsecs) in radius,
though it could be a bit larger or smaller - it's hard to tell, especially since there is so
much junk in that direction preventing us from getting a good view. Now here's a
rather strange feature - even though there is a bunch of dust and gas between
ourselves and the bulge, there isn't a whole lot of gas and dust within the bulge - a
very important aspect as to what does and does not occur in the bulge (as you'll see in
a little bit).
You may have noticed that there is
now a new unit of measure given
above, the kiloparsec (kpc). This is
just the same as a 1000 parsecs and
is a pretty convenient unit of
measure for sizes within the galaxy.
You could also use kilo light-years,
but no one is ever sure how to
abbreviate that.
Figure 12. A view looking toward
the direction the Milky Way's center
using a visible light telescope. Some
bright clusters and star formation
regions are indicated as well as the
location of some of the stars of
Sagittarius and Scorpius. The point
in the sky that is the direction
toward the center of the galaxy is
230 | P a g e

also shown. The line shows the orientation of the disk of the galaxy.
Let's get back to the bulge. What is in the center of this bulge? It isn't really easy to
tell, again because of all the gas and dust blocking our view. There are of course
views with other telescopes, and radio, infrared and x-ray telescopes have given us a
lot of information about what's happening there. One of the first clues was that in the
center of the galaxy is a very strong radio source. In the wonderfully creative way that
astronomers name things, they decided to call this strong radio source Sagittarius A,
or Sgr A for short. Isn't that clever? As radio telescopes got better and better,
astronomers were able to see more detailed structures and see that there were more
objects in the middle than just one radio source. In fact, there were a whole bunch of
other objects giving off all sorts of radio signals. After a while, there was Sgr B, Sgr
B1, Sgr B2, Sgr C, Sgr D, Sgr E, ... pretty sloppy, eh? Any ways, astronomers have
finally weeded out the sources and have come upon one that is stronger and more
centrally located than all of the others, which is called the Sgr A*, and this is the thing
that is at the center of the galaxy. They were also able to identify many objects that
aren't visible with a visible light telescope that we have yet to understand. Some of
these are shown in Figure 13.
Figure 13. Click on the image to see the full picture. A radio
telescope view of the center of the galaxy. Many features are
seen including supernova remnants (SNR), H II regions and
other objects whose nature has yet to be understood. The view
seen here and that in Figure 12 are in the same orientation. You
can see this by how the stuff is arranged along a diagonal line,
which is the plane of the galaxy. Image credit: N. E. Kassim, D.
S. Briggs, T. J. W. Lazio, T. N. LaRosa, J. Imamura (NRL/RSD).
We've determined that Sgr A* is what's at the center of the
galaxy. What's it like? First off, it isn't very big. It may only be a few A. U.s in size
(about 10 A.U. or so). Observations of stuff around it show that Sgr A* is not moving,
but that there is a bunch of stuff going around it. By looking at the speeds at which
this stuff moves and their distances from Sgr A*, and by using Kepler's third law, you
can figure out the mass of the stuff in the center - the mass of Sgr A*. Hold on to your
hats, campers, since the mass of Sgr A* comes out to be about 3.7 million times the
mass of the Sun! Egad, that's huge! You have something much more massive than the
Sun in a region of space smaller than our solar system. I wonder if it is a black hole?
This seems to be a popular idea, especially when we look at other galaxies and see
sort of the same thing there. In case you were wondering, a black hole that is 3.7
million times the Sun's mass would have a Schwarzschild radius of about 8 million
km. This radius is about 10 times the radius of our Sun. Such a black hole could easily
fit inside our solar system.
231 | P a g e

Figure 14. The motions of stars near the center of
the galaxy are mapped out and this helps
astronomers determine the mass of the object
located at Sgr A*. More information about this
research can be found here. Image from Andrea
Ghez et al, 2005
Sgr A*, or at least what we see, is itself not a black
hole, since you can't see a black hole. It is likely that
it is a disk of material around a massive black hole
(an accretion disk). That is one of the reasons it is
such a strong source of radiation. Also, the type of
radiation (in this case synchrotron radiation) tells us that there is a strong magnetic
field around the monster in the middle. What's feeding this monster? Actually, not a
great deal of stuff. For the amount of energy that is coming out of Sgr A*, you need
only drop in about 0.1 solar masses of material each year - so it doesn't suck in a huge
amount of stuff, but just needs a steady stream of material.
If you look at all of the stuff around Sgr A* (all those other Sgr things), some of the
objects are recognizable as supernova remnants. Actually, there are quite a few in that
direction of the sky (remember, you have to look through a lot of the disk to see the
center). There are also a bunch of unusual features streaming out from the center -
things known as threads. As usual, we're still working on figuring out what these
things are, so stay tuned; perhaps one day we'll figure them out (but don't hold your
breath).
You're probably wondering how you could have objects near a massive black hole
without their destruction, like the stars shown in Figure 14. That's a tricky problem
since many of the objects that are observed are young, massive stars, so they could not
have formed elsewhere and just happened to pass close to Sgr A*. A recent study
showed that it is possible for stars to actually form in the neighborhood of Sgr A*, in
particular in the disk of material that is likely to exist around the black hole. The
density of the disk "protects" the newly forming stars so they aren't shredded by the
black holes strong gravity - and they may be forming at a distance from the black hole
where it is relatively safe (though how you can ever be safe near a black hole is
beyond me). Check out this animation that shows how stars might currently be
forming in this dangerous environment.
Outside of the region housing Sgr A*, the bulge of the galaxy isn't very exciting. In
fact, the bulge sort of looks like an overgrown globular cluster, since many of the stars
in it are rather old, cool and yellowish looking. There is some debate, but there doesn't
appear to be a huge amount of star formation going on in the bulge of our galaxy. If
232 | P a g e

you want to look where stars are actually forming in a big way, you need to go to the
disk of the galaxy. One aspect about the bulge of our galaxy that is interesting is its
shape. The bulge isn't really spherical in shape like a globular cluster, but is actually
stretched out into a bar shape, which is a fairly common type of structure seen in
galaxies like ours (what are called spiral galaxies). Astronomers in Wisconsin found
that the bar extends quite a ways out, about 4 kpc from the center. This is about 1/2 of
the distance from the center of the galaxy to the solar system, so that's a pretty big bar.
And in case you were wondering, no they don't serve any tequila at this bar.
The Disk
The disk is where most of the action occurs in the galaxy, where most of the visible
material is, and where the planets, gas clouds, stars and other things hang around.
Also, because of all of that gas and dust, this is the region where star formation is
really active - it's pretty much the only part of our galaxy where stars are constantly
being formed. The disk is quite large, about 30 kpc from end to end (or, if you prefer,
it is 100 kilo-light years from end to end). It isn't very thick; it's only about 300 pc
thick (1000 light-years). As mentioned previously, we aren't near the center or the
edge but sort of midway between the center and the edge, about 8.0 kpc (26 kltyr)
from the center, and we're not even in the mid-plane of the disk but about 20 pc from
the mid-plane - definitely in the suburbs. There is still some debate about how far we
are from the center of the galaxy. Some astronomers think it may be as large as 8.4 -
8.5 kpc, while other favor a value closer to 8.0 kpc. As you'll see, this part of the
course is an area where the words "about" and "approximtely" will pop up quite a bit,
so keep in mind that these numbers may be revised in the future.
Why does the galaxy have a disk shape? It is easiest to explain by looking at how the
galaxy formed. It likely was rotating to some degree when it formed and this motion
would have caused material, as it fell inward, to form into a flat disk. The same
process can explain why the planets in the solar system are located in pretty much the
same plane around the Sun (the ecliptic).
We know from our early attempts to map out the size of the galaxy that the structure
of the disk can not be seen with optical telescopes due to the gas and dust. We instead
need to use radio wavelengths to measure the amount of material there. A very good
radio telescope wavelength to use is 21 cm, which indicates the presence of H I (this
is regular, normal hydrogen). Using this wavelength, it is possible to detect where the
gas is located, especially if it is in a concentrated form. Not only is H I detectable, but
so are a wide range of other molecules in the disk. Actually, things like alcohol and
sugar molecules have been found in space. I don't know if you can make a margarita
out of them, but it's out there.
233 | P a g e

If you want to see what the disk of the galaxy looks like at various wavelengths, just
check out this link.
There's all this gas in the disk (as well as dust and giant margaritas - no, I'm making
that last bit up). What is this gas doing? Some of it (if you remember a few weeks ago
in the course) was involved in the formation of stars. The disk is where this action
occurs - not just anywhere in the disk, but in the spiral arms. That is the pretty pin-
wheel structure you see when you look at pictures of galaxies. Why does the star
formation only occur in the spiral arms? Why doesn't it happen everywhere in the disk
- isn't there gas everywhere in the disk? What's up with that? Why can't you ever open
those little packets of ketchup that you get with your fries? I don't know about the last
question, but I'll try to tackle the
stuff concerning the galaxy.
Figure 15. A possible map of
the spiral structure of our
galaxy. The map is based upon
observations of various
segments of spiral structure and
theory of how the structure
should behave. In this model,
there are 4 spiral arms and the
Sun is located about 7.9 kpc
from the center of the galaxy
(marked by a +). While this
appears to show where the arms
are, other maps by other
astronomers show different
features or a different number
of arms. Based upon data from
J. P. Vallee (2005).
First of all, the spiral arms or spiral structure that you can see is due to the star
formation - these things stand out like a great big zit on prom night. Actually, they are
a bit bigger. The arms are bright because of large scale star formation. If you
remember what happens with large scale star formation, you'll recall that there are all
sorts of bright big things in these areas. If you were to look at the locations of the
spiral arms, what would you see? You'd see things like Giant Molecular Clouds and
the most massive stars that are formed (OB types). You'd also see the things around
such objects like the H II regions. Even though spiral arms are associated with stellar
birth, since the big stars die quickly, you can find their remains in the forms of
supernova remnants right next to the locations of their births. All of these big or bright
234 | P a g e

things are used to trace out where the spiral arms are, and to help us determine how
many spiral arms there are. That's why we call them spiral arm tracers. Astronomers
take a look at these things, map out their positions in the galaxy (as best they can), and
try to determine what the shape of the spiral structure in our galaxy is like. What is the
answer? We aren't sure. That's the annoying phrase you're going to see quite a bit over
the next few weeks. Even though we can see the stuff pretty easily and try to figure
out their locations, we're never in agreement. Some astronomers think that there are
two spiral arms, some think there are four, some think there are two but they are very
sloppy, some are completely clueless, and some could care less. How many arms are
there? I have no idea. There are probably two arms, since that is the most common
number, and it is likely that they are a bit sloppy so it looks like there are more, but
that's about the best we can do right now. You know, mapping out the spiral structure
of a galaxy when you are in the galaxy is like trying to make a map of an entire city
when you are stuck on just one street corner and you can't move around at all. It isn't
easy.
In spite of the difficulties astronomers keep working on the problem. One feature that
extends out from the bar at the center of the galaxy is known as the 3-Kpc arm (since
it is about 3 Kpc from the center). It is possible that this is the "base" of one of the
Milky Way's spiral arms. Recently it was determined that there is another 3 Kpc on
the other side of the galaxy (see the press release about this from Harvard here). This
would lead some to think that the Milky Way has 2 spiral arms as is shown in
the press release images. Is that correct? Well the Harvard study was announced in
June 2008. 6 months later, an astronomer at Iowa State University announced the
measurement of the 4 spiral arms! So which is it? Expect more discussion in this area
in the future....
If we can't tell what the structure of the spirals in the disk is like, can we at least figure
out what the disk material is doing? Remember, there is more stuff to the disk than
just the spirals - those are just the most obvious things. There is a lot more stuff out
there that isn't so obvious, and it has mass, so it must obey some sort of law as to how
it goes around the galaxy, but how does it do this? If things went around the galaxy
the same way that planets go around the Sun, you would have the stuff far from the
center moving very slowly (that's what Kepler's 3rd law tells you). It could also rotate
like a solid - the way a record or a cd spins around. In that case, the stuff far from the
center goes the fastest. What does the disk of the galaxy do? How does its material
move?
235 | P a g e

Figure 16. The rotation
curve for the Milky Way.
The curve varies at different
distances from the center.
Our location is at roughly 8
kpc from the center. Graph
is from a paper by Clemens,
1985.
The best way to figure out
how the disk moves is to
follow the motions of the
most common stuff in the
disk - those big clouds of
gas (H I) that are all over the place. These clouds and other clouds like CO or
CO
2
give us a pretty good idea of how the disk moves, and that is shown in Figure 16.
We don't really have a good picture of what is going on in the middle (in the bulge),
but for most of the galaxy, we have a good view of the action. We (the solar system)
go about the galaxy at a speed of around 220 km/s and it takes us about 225 million
years for one trip around.
Here is a nifty idea. What if we pretend that we are orbiting the center of the galaxy?
If we do that (and that isn't such a wild idea), then we can use Kepler's 3rd law to
figure out the mass of what we're going around. Why is that? If you remember, and
I'm sure you do, Kepler's third law can be rewritten as
M
1
+ M
2
= a
3
/P
2

where a is the distance from the center (in AU) and P is the orbital period in years.
The values for M
1
and M
2
are the mass of what your going around (the galaxy's mass
within our orbit) and your mass. The masses of the Earth, Sun, Moon, and all the rest
of the stuff in our solar system is really dinky compared to the mass of the galaxy, so
we can simplify the formula to
M
galaxy
= a
3
/P
2

where we'll just ignore our own mass. Here's a really nifty thing - we can determine
the mass of our galaxy, or at least part of it (the amount within our orbit), just by
looking at how long it takes us to go around the galaxy and how far we are from the
center. When we use the numbers we know and convert them into the appropriate
units, we get a mass of about 10
11
solar masses - 100 billion times the mass of the sun.
Wow, that's a lot! Actually, that really isn't since we are only measuring part of the
236 | P a g e

mass of the galaxy - this is the amount of material we measure going out only about
half way from the center. If we look at other parts further from the center and do the
same calculations, we get some really large masses - sometimes values like 10
12
solar
masses. That's a lot. Are we sure about this? Yes, because if you look at how the
galaxy rotates in the other parts, we see that the velocity stays pretty high or even
increases. The only way to sustain this high velocity is to have a large amount of mass
in the outer galaxy. Without the mass, the velocity would drop off. It's not like our
galaxy is a weirdo; we see the same thing with other galaxies similar to our own.
There is only one problem with this large amount of material. We detect a large
amount of mass in the outer part of the galaxy based upon the way stuff in our galaxy
moves, but we can't actually see all this stuff. The outer parts of galaxies (not just
ours, but others like it) appear dimmer than the parts closer in. This material that
makes up a large fraction of the mass of the galaxy just doesn't show up very well - it
is very hard to see. Astronomers call this dark unseen matter (oh, you'll love this
name) dark matter. How's that for a very clever name? Actually, it sort of makes
sense, because there is quite a large amount of matter, but it doesn't show up in our
telescopes - and I just don't mean the telescopes that detect visible light, but also other
types of telescopes. We'll run into more dark matter later on, so keep it in mind.
You may have noticed that in the outer part of the galaxy the rotation curve is fairly
uniform - that stuff is moving at roughly the same speed. Isn't that special? Actually, it
may be a disaster. If the velocity is the same value or nearly the same value for stuff at
all of those different distances from the center, then things at those distances will
travel roughly the same distance around the galaxy in the same amount of time. Say
you have a straight line drawn from the edge to the center of the galaxy. If you let it
go for a while, the inner part of this line will get around the galaxy before the outer
parts of the line, since it has a shorter path to travel. After a relatively short time, what
started out as a straight line would wind up completely. If a straight line would wind
up, wouldn't all other structures also "wind up"? How come we have spiral arms then?
Why aren't they all wound up? The solution is rather simple: The spiral pattern (spiral
arms) do NOT travel at the same rate as the material in the disk. They travel through
the disk but they don't move in the same way that the disk material does (which is
shown in Figure 16). That's about as confusing as it gets. How can the spiral arms
travel at a different reate than the stuff in the disk? Aren't they in the disk? Yes, they
are, but they are not solid objects. The spiral pattern is like a sound wave traveling
through air. The sound doesn't cause material to move around a great deal, it just
passes through the air and compresses it as it goes along. This is sort of what is going
on in the spiral arms. The concept that explains this is known as the spiral density
wave theory, but I like to call it the naked jogger model. Let me explain it first to
show you what I mean.
237 | P a g e

Let's say you have a bunch of cars going down the freeway. They are moving along at
a steady pace and are well spaced. Now, let's put along this freeway a nude jogger.
You get to choose the gender of the jogger. Now we have a nude jogger going down
the road, and what is going to happen to the traffic? Obviously, the traffic is going to
get bottlenecked in the area of the nude jogger. It won't stop; it will just slow down
and get congested for a while. The location of the freeway that experiences this
congestion will gradually change - it will move along as the jogger moves along. Far
from the jogger, things are pretty normal, whereever the jogger is, things are
conges
ted.
Figur
e
17. An
analog
y of
how
the
spiral
arms
move
throug
h the
disk.
The cars are moving down the highway toward the right. At first they are well spaced.
A nude jogger is represented by the green dot (you thought I would include a picture
of a nude jogger?!) and the jogger is moving toward the left. As the jogger moves
down the side of the road, the traffic near the jogger slows down and gets congested.
The area of congestion moves down the highway at a pace that is independent of the
stuff moving into the congestion area. The cars represent disk material, the jogger
represents the density wave and the congestion is the star forming regions.
What the heck does a nude jogger have to do with the spiral structure of our galaxy? It
is just like it! The jogger is a density wave (just like the spiral pattern); in the area of
the jogger, traffic is congested (just like a GMC gets compressed in a spiral arm). The
jogger is moving independently of the stuff he/she compresses (just like the spiral
pattern moving through the disk material). So, in a galaxy, you have a really large
nude jogger - no, not really. You have a region of compression that travels through the
galaxy, and it goes independently of the stuff it compresses (the gas clouds). We know
where it is, because the compression causes star formation and that shows up as the
238 | P a g e

things we call the spiral arms. I guess this would be like the nude jogger causing a few
fender benders.
Let me re-cap: the spiral density pattern moves through the disk at its own speed,
compressing the disk material as it goes along, causing star formation - which is what
we see as the spiral arms. The stuff in the disk - the gas, the stars, all that stuff -
moves according to the rotation curves we measure (like in Figure 16), while the
pattern moves in its own sweet way.
Why is there a spiral density wave moving through our galaxy? We're not really sure
where it came from or what keeps it going, but there are some ideas about how
galaxies evolve that we're still working on. Perhaps you should re-take this course in
about 10 years; we might have the answers by then.
Halo
This one is pretty simple - this is the region around the galaxy where you find globular
clusters. It is actually a large, spherical region that surrounds the galaxy's disk. We
can estimate its size by looking for RR Lyrae stars in distant globular clusters. When
we look at the velocities of these distant objects, we know they are moving fast due to
the presence of a large amount of dark matter in the halo. Again, we're not quite sure
what the dark matter is, but there is a whole bunch of it out there.
That pretty much wraps up what we know about our galaxy - pretty spiral, unknown
number of arms, probable black hole in the core, spiral density wave (what makes the
spiral pattern) helping in the formation of stars, and a large fraction of its mass
remaings unseen (all that dark matter).
Beyond the Halo
While some people think of galaxies as pretty well defined objects, they really are not.
They don't have abrupt edges but tend to thin out in terms of the amount of material
they have, over large distances. When we say the size of the Milky Way is around
100,000 light years across, that is clearly an approximation, since material can be
found at even greater distances from our galaxy. We also have some streams of
material that is associated with our galaxy and extends well outside of the normal
"edges" of our galaxy. Astronomers generally believe that these streams are material
that was once part of our galaxy or another galaxy, that has been gravitationally
stretched. You can think of this as sort of cosmic breadcrumbs left behind as a small
galaxy or globular cluster gets shredded by the Milky Way. As you can see here, the
streams extend quite a ways from the disk of the galaxy.
239 | P a g e

Stellar Populations
Before we start looking at other galaxies we have to talk about stars again. You have
to remember that when you look at distant objects, you can't always see things as
individual objects. You usually can't see individual stars at great distances; instead
you see all the stars in a region (or a galaxy) as one glowing fuzzy blob. This is
exactly what Walter Baade saw when he was looking at other galaxies. It was during
World War II, and Baade was a German national who was working in the US. He
could not do war research like the other scientists were doing at the time, so he spent
all of his time up at the Mount Wilson Observatory in southern California. Normally,
this is a really lousy place to look at the sky due to all of the light pollution from Los
Angeles, but during the war the city was under a black out, so the night skies were
really beautiful. Baade took advantage of these conditions to observe the Andromeda
Galaxy, a large nearby galaxy similar to our own in its structure. He noticed that the
inner and outer parts of the galaxy were different. He also noticed that the stars in the
center of the Andromeda Galaxy and the stars in the direction of our own galaxy's
bulge looked very similar. He decided that the stars in the inner part (bulge) and the
outer part (spiral arms) were different. In the fine tradition of really original
astronomical names, he decided to name these two groups of stars Population
I and Population II stars.
Population I stars are found in the disk around the locations where you find young
stars. These stars have compositions like the Sun (particularly in terms of how much
iron and other heavy elements are in them). You find this population in areas where
you have current star formation occurring - this is a big clue as to where they are and
are not seen - you look for star formation! The Sun is a Population I star - hooray,
we're #1, we're #1! Pop I stars are found in regions of star formation, so they tend to
contain very hot stars and give the regions they are found in a rather blue-ish
appearance. Population I stars tend to have rather boring orbits, since they are found
in the disk of the galaxy - they pretty much stay close to the disk.
Population II stars are found in actually two places - away from the disk, in the halo
where the globular clusters are, and in the bulge. These are places where there is no
current star formation, so you are only looking at old stars in these locations. These
old stars are metal deficient (low in the amount of iron in them) when compared to
Population I stars and tend to look rather reddish. Population II stars have very wild
orbits; they are all over the place with respect to the disk. While they tend to orbit
about the center of the galaxy, some go up, some go down, some go this way and that.
They aren't as orderly as the Population I stars.
240 | P a g e

It should be noted that stars, gas, and dust in the disk are continually being recycled
with each new generation of stars. The first stars that were formed would have had
only hydrogen and helium in them. When they died and blew their guts out in
supernova explosions or planetary nebula phases, they scattered the heavier elements
(which were produced during fusion or supernovae/explosive events) out into space,
along with hydrogen and helium, of course. Later generations of stars are then made
from these elements (along with any other stuff in space, which was still mainly
hydrogen and helium). Population I stars are the way they are since material from
many previous generations of stars was incorporated into them when they were
formed. As time goes on, with more stellar birth (and death), the metal content of the
galaxy's disk will increase. As long as stars are continually being formed, they will be
made of more and more metal rich material (carbon, oxygen, iron, nitrogen, etc).
Without continuous stellar birth and death, the metal content of the galaxy would not
change but would stay at the same level. This is what we see in the halo and bulge of
our galaxy (remember, they don't have current star formation, so their metal content
hasn't changed much). This metal content-stellar population-star formation
interconnectedness will be obvious when we look at other galaxies.

Now that you've read this section, you should be able to answer these questions...
What prevents astronomers from easily mapping out the size of the galaxy, and
our location within it?
What are the properties of Cepheids and RR Lyrae stars and how can they be
used to obtain distances?
What are the properties of the different types of clusters?
What is the center of the galaxy like?
What evidence is there for a massive black hole in the center?
How do astronomers measure the mass of our galaxy?
How do astronomers map out the structure of our galaxy?
How does the galaxy rotate, and what exactly is moving in this rotation?
How do the spiral arms move relative to the disk material?
What is the halo?
Why do astronomers suspect that our galaxy contains a large amount of dark
matter?
What are the characteristics of the different stellar populations and why do they
have those characteristics?
What evidence exists that supports the Special or General Theories of
Relativity?
What are the measurable characteristics of a black hole?
241 | P a g e

What happens as you fall into a black hole, and what would others see
happening to you?
How can we search for black holes?
What are magnetars, and quark stars?


Figure 12. The inner ring around SN 1987A is
starting to light up due to the collision of the
shock wave from the supernova. The yellow areas
in the picture on the right are where the
collisions are currently occurring. Image credit:
P. Challis and R. Kirshner (Harvard-Smithsonian
Center for Astrophysics), P. Garnavich
(University of Notre Dame) and Z. Levay
(STScI).

Here is a nifty animation showing the way that SN 1987A and the rings have changed
over time. First the supernova happens, and after a while its energy lights up the ring
as the light passes through it. Then the ring gradually fades away. Then the ring starts
lighting up again, this time due to the collision of the shock wave with the ring
material. The ring and the shock wave are not nice and neatly organized objects, so
some parts of the ring will light up before others. It will take some time for the
collision to stop and the ring to again fade away, but until then, it looks like a rather
nice show!

Now that you've read this section, you should be able to answer these questions....
What are the various fusion processes that massive stars can go through?
Why is each fusion process shorter than the previous one?
What causes the core of a star to collapse?
What happens during the core collapse?
What are the important characteristics of supernovae?
What are the characteristics of the two main types of supernovae that allow
astronomers to tell them apart?
What are hypernovae and how are they related to gamma-ray bursts?
What information can be obtained from old supernova remnants?
Why was SN1987A so important?
What is currently happening to the environment around SN1987A?
Other Galaxies
242 | P a g e


What's covered here:
How do astronomers distinguish galaxies from other fuzzy objects?
What did Edwin Hubble do that made him so famous?
What different types of galaxies are there?

We know quite a bit about other galaxies by using information about the Milky Way
and applying it to them as well as using our own observations of other galaxies to
figure out what is going on out there. The reverse is also true - characteristics of other
galaxies can be applied to the Milky Way. Remember, we have a hard time seeing
various parts of our own galaxy, so checking out other galaxies gives us an idea about
what distant parts of our galaxy look like or what is probably happening in those
places.
243 | P a g e


An image showing a wide array of galaxies. The "spikey" objects are stars in our own
galaxy. Can you tell which object is closer and which is further away?
What do other galaxies look like? To the eye they are fuzzy patches in the sky. With
long exposures on film, various features such as spiral structure and star clusters are
visible. Though it is possible to see trends in the general shapes of galaxies, they are
all unique - sort of like snowflakes.
But when you go to a telescope and look out in the sky and you see these fuzzy things,
how can you tell it isa galaxy? All little bits of fuzz look sort of similar through a
telescope, though there are certain pattern to some of them. And just by looking at
them with your eyes through a telescope doesn't really help. Even the invention of
photography which helped astronomers get more detailed images of galaxies, wasn't
enough to tell them what these things were and how far away they were. That was a
major problem in astronomy around the beginning of the 20th century. There were all
244 | P a g e

of these fuzzy things out there (which we callednebula, plural is nebulae) located all
over the place. But what were they? We thought some of them were in our galaxy, but
we were not sure which ones were and which ones weren't. Perhaps none of the fuzzy
things were outside of our galaxy? Perhaps our galaxy is the only galaxy - who says
there really are any other objects out there any ways? Or perhaps our galaxy is just
like one of those fuzzy things - but like which type of fuzzy things? Does it have a
spiral shape like some of the spiral nebulae? Or an oval shape like some of the nebula
have? Or maybe some other kind of shape? According to the best data at the time, we
didn't really know where we were located even in our own galaxy, so how was it
possible to figure out what those darn fuzzy things were?
This is really a conundrum. How will the astronomers be able to solve this problem
and save the day (well, at least solve this problem)? They didn't really solve the
problem at first. Initially there was just a lot of talking about the problem and trying to
show who was right, who was wrong, and who had the worst breath. One of the big
discussions about the whole nebula problem took place in Washington D.C. in 1920 at
the National Academy of Science. The two people involved in the debate were Heber
Curtis and Harlow Shapley, so this debate is known as the Curtis-Shapley Debate.
Now you've already run into Harlow Shapley, who was the fellow that figured out
the distance to the center of our galaxy back in 1915, but some people had their
doubts about his work. The other fellow, Heber Curtis (1872-1942), studied the
spiral nebulae quite a bit, so he was really the expert when it came to those things.
You have two guys - one an expert about the size of our galaxy, the other an expert on
the spiral nebulae - debating each other about various things like what are those spiral
nebulae? What is it really like out there? Just how big are the spiral nebulae? How far
away are they? How big is our galaxy? What is the Universe like?
It's sort of a draw as to who won the debate. Since Shapley was an expert on the
Milky Way, he was able to put forth his idea about the distance we are from the center
- which was pretty big. This was correct. (Shapley actually thought that the Milky
Way was about 100 kpc wide, which isn't really correct, but he was along the right
track). Then he proposed that since the Milky Way was so huge (at least, bigger than
anyone ever thought it could be), those spiral nebulae were probably not very far
away objects. After all, our galaxy is so huge - or so he thought.
Curtis believed that the spiral nebulae were actually distant objects, not part of our
Milky Way, and not even very close to it. He also thought that our galaxy is just like
them in terms of size, shape and structure. He was correct on those points. Then he
put forth that the Milky Way was not very large (he thought it was only about 10 kpc
in size, which is much too small), since it was not large enough to contain the spiral
nebulae. He sort of missed that one.
245 | P a g e

In a way, since each of them was an expert in one aspect of the problem, they each got
one thing right. They didn't know that at the time, since just debating about something
doesn't help figure out what the actual answer is. Shapley was generally considered
the winner of the debate due to his charisma (imagine that, an astronomer with
charisma!). The debate really didn't solve the problem. What did solve the problem
was good, solid science, in this case when Edwin Hubble provided real evidence for
the distances to galaxies in 1924. Hubble had an advantage over Curtis and Shapley,
since he had the use of the new, big telescope on Mount Wilson in southern
California. At that time, the big scope was the 100" (2.5 meter) telescope located
there.
Hubble and the telescope operator, Milton Humason used this telescope to find and
analyze Cepheid stars in the Andromeda Galaxy (of course, at that time, it was still
known as the Andromeda Nebula). What good are Cepheids? - plenty good! If you
remember from the previous section, Cepheids are those Red Giant stars that pulsate
and you can use them to determine distance. The longer the period, the greater the
average brightness (the Leavitt Law) so if you can find a Cepheid with a period
similar to one in our own galaxy, you can compare their apparent magnitudes (how
bright they look to your eye), and the difference in brightness is directly related to
their different distances. If you can find a Cepheid in a galaxy, you can find the
distance to that galaxy. That is exactly what Hubble and Humason did. By using the
Cepheid Leavitt Law to determine the distance to the Andromeda Galaxy (one of our
closer neighbors), they found that it was 900,000 light years away! This distance was
much greater than anyone even suspected, and this is one of the close ones! Actually,
they were a bit off - it is actually about 2,250,000 light years away - our current
distance estimate methods are a little bit better now.
Not only did Hubble figure out that those fuzzy spiral nebulae (like Andromeda) were
actually very distant, separate objects, he started a whole new field of astronomy, the
search for objects that could reveal distances to remote galaxies. To find the distances
to far away galaxies, it is necessary to use objects whose brightnesses we know fairly
well (or some other property that is well defined) and also objects that are bright
enough to be seen at a great distance. Objects that fit both these criteria are referred to
as Standard Candles. That's just sort of a cute nickname for bright, well behaved
objects. What sort of things are Standard Candles? Here are some examples of them -
Cepheids - The best thing to use, very bright, fairly common, reliable thanks to
the Leavitt Law
RR Lyrae - Like Cepheids, but not as useful, since not as bright. Generally used
only for nearby galaxies
Supernovae - very bright, pretty useful, but can be tricky since there are two
types of supernovae, and some can be rather abnormal
246 | P a g e

Planetary Nebulae - these are pretty hot and produce a lot of UV light, so they
can be seen distinctly from other things in the galaxy; pretty reliable as distance
indicators
Novae - pretty bright, but not all novae are alike, so not as reliable for accurate
distances
OB stars - by looking at the brightest stars you can sometimes get good
distances, but these are only in certain types of galaxies and are often in regions
of star formation that have quite a bit of dust
H II Regions - not so great, since they can have different sizes
Globular Clusters - if you can't see individual stars, use a whole cluster, but not
too good, since clusters are not exactly alike - different sizes, brightnesses
Brightness of the entire galaxy - assuming that all galaxies have the same
brightness is not too good, since they come in a range of brightnesses - often
just the brightest galaxy in a group is used
and other methods...
Once an astronomer can determine how far away a fuzzy blob in the sky is using a
Standard Candle, they'll know if the fuzzy blob is in our galaxy (a few thousand
parsecs away) or is a distant galaxy (millions or more parsecs away).
Astronomers are able to get distances for galaxies within about 1 billion light-years
fairly reliably, but there tend to be greater and greater uncertainties in the values for
greater and greater distances. If you were to look up all of the distances to even
nearby galaxies (like Andromeda or the Large Magellanic Cloud), you'd see a range
of values, not just one single value. For very distant galaxies (say, more than 1 billion
light years away) the distances that we derive are much more imprecise and can
always be improved. This is one of the reasons that bigger and better telescopes are
being built all of the time, to measure more and better distances. We need to know
how far away things are so we can figure out what the Universe is like.
Galaxy Characteristics
Once it was determined that many of those fuzzy things were actually quite distant
galaxies, astronomers had to classify them. Why? We are sort of compulsive about
doing things like this, but really it is because we could learn more about them if we
could group them together in a way that was scientifically meaningful. This is sort of
like how we can group stars into bins like Main Sequence, Red Giant, and so forth.
We know that objects in such groups share common characteristics, and we can use
that information to learn more about galaxies that we don't see very well or that are
too distant to measure all of their characteristics with much certainty.
247 | P a g e

Figure 1. The
Hubble Tuning
fork diagram
showing the
different forms that
galaxies come in.
The main groups
are the ellipticals,
the spirals and the
barred spirals. A
transitional form
is the Lenticular
type (labeled S0).
Anything that can't
be placed on the
Tuning fork due to
unusual structures
is simply labeled
as Irregular.
What do they look like? Do they all look the same? No, of course not; that would be
too easy. Most often we classify galaxies based upon their appearance, since that is
the most easily observed feature. The basic classification scheme that is used is known
as the Hubble Tuning Fork Diagram (I wonder what clever astronomer thought that
up). Yes, good old Eddy Hubble set down the framework for the primary
classification scheme. There are some other schemes used, and there have been slight
alterations to the guidelines that Hubble used, but it is pretty much still the same thing
used today. It should be noted that this scheme is based only on appearance - the
shapes of galaxies. It doesn't account for how they got into those shapes or the
differences in sizes that exist.
Not only do the galaxy shapes vary, but also the content of the galaxies varies -
different types of galaxies can have quite different types of stars in them and different
environments. This can result in galaxies having different colors, different things
happening (or not happening) in them, different ages, different evolutions, and so on.
Remember the different stellar populations -
Population I - hot, young stars present, which are chemically like the Sun; their
presence indicates that current star formation is going on
Population II - old stars dominate, metal deficient compositions, no new or
significant star formation currently occurring
248 | P a g e

Remember, galaxies are very far away, so you generally can't see individual stars, but
you can see large groups of stars. That is why we talk about stellar populations, since
the characteristics of groups of stars is what we are able to measure. Let's start
checking out the different types of galaxies that are out there.
Ellipticals

Figure 2. Several different elliptical galaxies are shown. Copyright Association of
Universities for Research in Astronomy Inc. (AURA), all rights reserved.
As the name implies, these are elliptical in shape, though some are not very elliptical
at all but look like circles. To distinguish the different shapes we use a numerical
designation along the lines of E0, E1, E2...all the way up to E7. The "E" is for
elliptical, while the number describes the degree of ovalness. The number is found by
measuring the long (a) and the short (b) axis, and taking those values and putting them
into the following formula
10(a-b)/a
Figure 3. The method used for defining the different
elliptical galaxies is illustrated here. The longer axis
length is compared to the shorter length and a number
based upon this value is used to distinguish the range of
elongation. In the first case both axes are the same length
so the type is E0, while in second the value of 4.7 is found
using the formula, which becomes 5, making that
elliptical an E5.
Ellipticals tend to look rather yellowish or orangish. This
indicates that they are made up of mainly Population II stars. Observations of them
show that there is no new (or significant) star formation occurring. There is not much
star formation occurring, which means that there must not be a lot of gas and dust in
249 | P a g e

them, since this is what stars are made from. This also gives us a clue concerning how
they were made - but I'll get to that later. If you were to look at how the stars in
elliptical galaxies move, you'd tend to see rather random motions (sort of how
globular clusters move around our galaxy).
Figure 4. A
group of
galaxies with
a large cD
(Giant
Elliptical) in
the center of
the group.
Image from
the Hubble
Space
Telescope.
The biggest of
the ellipticals
are often just
called Giant
Ellipticals an
d these are the
largest of all
galaxies. They
get a special
designation
rather than the E designation; they are labeled as cDgalaxies - don't ask me why
they're called that, they just are. These tend to be very spherical in shape, so I guess
they don't need the "E" designation scheme - but there are other features that make
them distinct. They can have masses of up to 10 trillion solar masses (10
13
M
solar
).
They are so big that they tend to be found in the center of groups or clusters of
galaxies. It is likely that these big brutes weren't always that big but have gotten
bigger over time by eating up little galaxies that got too close to them (what we
call Galactic Cannibalism- really, we do).
On the other end of the scale, one finds the Dwarf Ellipticals and Dwarf
Spheroidals. These are among the smallest of all galaxies, typically with masses
around a few million solar masses (10
6
M
solar
). Dwarf Ellipticals and Spheroidals can
be best described as galaxy groupies, since they tend to hang around much larger
galaxies. If you look at a picture of the Andromeda Galaxy, you'll see two little dwarf
250 | P a g e

galaxies (one is an elliptical the other spheroidal) around it. Due to their wide range of
masses, ellipticals are sort of hard to figure out. Sometimes it is hard to determine if
you are looking at a nearby dwarf elliptical or a distant larger elliptical.
Spirals

Figure 5. Several different spiral galaxies. Copyright Association of Universities for
Research in Astronomy Inc. (AURA), all rights reserved.
Spirals show a much greater range of structure than ellipticals, so their classification
is a bit more complex. First there is the letter "S" designating the galaxy as a spiral.
Then there are the cases where there is a Bar going through the center of the galaxy.
If so, you need to add a "B" to the designation. Then there are the other characteristics
- how big the bulge is compared to the entire galaxy, and how tightly wound up the
arms are. There is a tendency that when the bulge is large the arms are wound up
pretty tightly, and when the bulge is really small the arms are really spread out. The
letters a, b, c and d are used to categorize this characteristic. The various designations
for spirals are Sa, Sb, Sc, Sd, SBa, SBb, SBc and SBd. Some people are a bit
indecisive about a galaxy being in a particular group, so sometimes a spiral can be
designated as a Sab or Sbc, since they're not sure which group it belongs in.
Spirals are easy to identify since they have a spiral structure or flat disk shape (if seen
edge on). Of course, they have the spiral arms due to the star formation that is
occurring there, but remember, there is material between the arms; it is just not as
exciting or as easy to see as the arms. The arms stand out so well because they have
all of those hot, big stars to light them up as well as the H II regions in the area. The
masses of spirals are typically a few billions to a trillion solar masses. There is the
added complication that they aren't made of the same stellar populations. The
populations of stars vary depending upon where you are looking - in the disk you find
Population I stars and in the bulge and halo you find Population II. This is why in
251 | P a g e

color pictures of some spirals you see the disk looking bluish while the bulge looks
yellowish-orangish.

Figure 6. Classic examples of Barred spirals. Copyright Association of Universities
for Research in Astronomy Inc. (AURA), all rights reserved.
Barred Spirals share pretty much the same characteristics as spirals except for that
extended bulge. It is sort of like someone has taken the normally circular bulge shape
and stretched it out. The arms then start up on the ends of the bar. Due to this added
structure, the arms in barred spirals tend to be wound up a little bit more tightly than
in regular spirals. It is now thought that the Milky Way Galaxy is a barred spiral;
perhaps it could be classified as a SBb or maybe even
an SBc.
Figure 7. A barred spiral, NGC 6744. It is possible
that this is what our Milky Way galaxy looks like.
Notice how the arms are not very distinct and poorly
defined. It is also thought that the Milky Way galaxy
has a bar similar to this one. Image Anglo-
Australian Observatory, Photograph by David Malin.
Lenticular Galaxies
S0 and SB0 types, also known as Lenticular
Galaxies, are sort of crosses between a spiral and and
elliptical. They are best described as having a flying saucer shape, since they have a
disk and a bulge like a spiral galaxy but no spiral arms. The bulge is often pretty big!
They don't have any spiral structure, so they don't have much star formation going on
(remember, that's why we have spiral structure). The lack of star formation indicates a
lack of gas and dust out of which to make stars. Thus, S0 galaxies have mainly
Population II stars in them. If they have a bar, then the SB0 designation is used (make
252 | P a g e

sure you don't get the letters out of order for this designation!) In general S0 galaxies
are pretty rare.

Figure 8. Lenticular (or S0) galaxies. These look like a cross between an elliptical
and a spiral. You might think of them as a type of spiral galaxy without any spiral
structure. The one on the left has a dusty plane, but that is unusual for these galaxies.
The one on the right is in the same orientation but shows no dusty structure.
Copyright Association of Universities for Research in Astronomy Inc. (AURA), all
rights reserved.
Irregulars
As with any classification system, there have to be a bunch of objects that don't fit in.
For galaxies, these are the Irregulars. Amongst the more famous irregular galaxies
are the two neighboring galaxies to the Milky Way, the Large and Small Magellanic
Clouds (LMC and SMC). They sort of have a bar-like structure, but there isn't
anything else there - no spiral structure, no defined bulge, nothing.

Figure 9. Some typical irregular galaxies. The Large Magellanic Cloud is in the
middle, and the Small Magellanic Cloud is on the right. Copyright Association of
Universities for Research in Astronomy Inc. (AURA), all rights reserved.
253 | P a g e

It is thought that if there was some structure to an irregular galaxy at some time, like
spiral arms or bars, then those parts of the galaxy could have been stripped off due to
collisions or other gravitational interactions with larger galaxies. Galaxies don't have
to actually get too close for there to be tidal disruptions of the galactic structure. As
previously mentioned, it is possible for one large galaxy to strip off the gas and dust
from a small, nearby galaxy and to suck it up. The bigger galaxy is basically eating
away the star forming material (gas and dust) from its hapless victim. This is known
as Galactic Cannibalism (and is best served with fava beans and a nice chianti).
Irregular galaxies tend to be associated with rather tumultuous events, so they tend to
have a lot of star formation going on in them, but this isn't true for all of them, since
there can be a wide range of stellar populations (I and II) in different irregulars.
Generally, Population I is what is seen. Irregular galaxies form from previously
normal galaxies, so they tend to have a wide range of masses. They're just irregular!
It is worth mentioning that the two irregular galaxies that are best known are the LMC
and SMC (Large and Small Magellanic Clouds). These are not visible from Iowa, but
only from fairly far south of the equator. It is possible to see these two nearby galaxies
with the naked eye. Since they are so close, they have been studied very extensively -
though you have to go to Chile, Australia, or South Africa to study them at all.
Because these galaxies appear to be interacting with our own, and it is likely that the
Milky Way collided with them in the past, they show a great deal of star formation
currently going on. New images showing the hot gas in these galaxies are visible, and
can be seen here - for the LMC and the SMC. Features such as hot gas from star
formation, novae and old supernovae is clearly visible. The part of the galaxies that
we tend to see with our eyes is generally much smaller.

Now that you've read this section, you should be able to answer these questions....
What methods are used to determine the distances to galaxies?
What is the general scheme that is used used to classify galaxies?
What are the criteria used to classify elliptical galaxies?
What are the criteria used to classify spiral galaxies?
Why do irregular galaxies have no specific form?
Properties of galaxies

What's covered here:
How do astronomers measure the various physical characteristics of galaxies?
What influences the evolution of galaxies?
254 | P a g e

What are galaxy clusters?
What other galaxy groupings exist?
What is dark matter?

Before people knew that galaxies were separate objects outside our Milky Way, many
of the brighter or more prominent ones were cataloged in with the groups of various
fuzzy things in the sky, which includes stuff like planetary nebulae, star forming
regions, globular clusters and other non-galaxy things. When you look at the names of
some of these objects today, you can see a galaxy listed right next to a globular cluster
or a planetary nebula. Later catalogs were compiled which tried to include only
galaxies or were put together by specialized telescope surveys, so some names are
really screwy. In general you'll only run across galaxies that are listed in the common
catalogs. There are of course galaxies that have proper or cute names like Andromeda,
the Large Magellanic Cloud, the Antenna galaxies, etc., but these are pretty rare. More
often you'll see a name from one of the catalogs like the following -
Messier catalog (M1, M2, etc.), which has listed for the Andromeda Galaxy
M31. There are about 110 objects in this catalog.
New General Catalog (NGC 1, NGC 2, etc.), which all start with NGC, and in
this catalog the Andromeda Galaxy is known as NGC 224. There are nearly
8000 objects in this catalog.
Index Catalog (IC 1, IC 2, etc.), sort of an extension of the NGC catalog, and a
galaxy in this system could be called something as exotic as IC 3242. There are
about 5500 objects in this catalog.
From this little discussion, you can see how original and creative astronomers can be
with naming galaxies. Actually, with there being thousands of galaxies visible to most
large telescopes, giving them all individual names like "Bob" and "Becky" would be a
bit too difficult. You must remember that the catalogs mentioned above are just the
more commonly used ones, and these catalogs include a lot of non-galaxy objects.
The Andromeda Galaxy is so commonly surveyed and studied that it has more than 20
different catalog names or designations - talk about overkill! Now once astronomers
started figuring out which objects were and were not galaxies, they started to look for
other features that they could compare. Two things that we like to measure with stars
are mass and luminosity, and this is also true with galaxies. Let's see how that is done,
shall we? Unfortunately, with galaxies being so far away, it isn't always possible to
determine their masses accurately. There are a few special situations where the mass
can be determined, though, and these are done by using methods you've already seen
before - by looking at how quickly a galaxy rotates and by looking at the motion of
binary galaxies.
255 | P a g e

The first method, using the rotation of the galaxy, is the same way that we determine
the mass of the Milky Way. If you measure the speed of rotation (from the galaxy's
spectrum) of a distant part of a galaxy, you can then use Kepler's laws to figure out
how much mass is located within the orbit of that distant part. There are a few
problems with this method, mainly that it really only works well for spiral galaxies, so
it can't be used for all types of galaxies. You also want to have the galaxy tilted so that
the velocity is directly toward/away from us. If the galaxy is tilted at a random angle
relative to our view, we have to take that into account when figuring out the velocity
and therefore the mass of the galaxy, so it is always better to have direct motion
toward or away from us. When we do this experiment we see that spiral galaxies
rotate pretty much like the Milky Way does - and this supports our observations of our
own galaxy.
Figure 1. Several galaxies in
a group. The motions of these
galaxies about one another
can help determine their
masses. Image Credit:
Hubble Heritage Team
(STScI/AURA/NASA).
The other method used to
figure out the mass of a
galaxy is to observe a binary
galaxy system. This is just
like using binary stars to
determine the masses of stars.
There is only one small
problem. Galaxies are like
giant lumbering elephants -
they don't appear to move
very fast as we see them. This is in part due to their great distances. If a jet plane were
30 feet from you and it went by it would obviously look like it was moving fast, but
when it is 30,000 feet up it appears to be just inching along. It isn't actually going at a
slower speed, but the apparent distance it covers looks to be less since it is further
away from you. Now extend this to galaxies. These beasts can be moving at hundreds
of kilometers per second around each other, but because they are so far away you
never see any visible motion in your lifetime. Even if you live to be 100 years old,
you won't see any motion. Even though we can't see the motion, we can still get a clue
as to how fast they are moving based upon how their spectra are screwed up by the
Doppler effect. There is a problem with this. The Doppler effect only measures
256 | P a g e

velocities toward or away from you, not velocities that are side-ways. No matter how
long you look or how carefully you analyze the spectra, you'll never accurately
measure the entire velocity of a pair of binary galaxies. Often astronomers have to
guess about the amount of velocity they are missing. I suppose I should not say guess,
since that sounds like we're throwing darts at a board and using those numbers. We
really don't do that, but instead use various statistics to figure out how much of the
velocity we are missing.
Any ways, after we estimate the velocities and the separation of the binary galaxies,
we can get their masses using Kepler's Third law, just like with binary stars -
M
1
+M
2
= a
3
/P
2

where M
1
and M
2
are the masses of the galaxies and a and P have their usual
meaning. As with binary stars, the masses are in solar masses, the distance is in A.U.s
and the period is in years. Isn't that nifty? Kepler's simple little law works for more
than just planets; it works for whole galaxies!!! You probably didn't think I'd get this
much mileage out of that one silly little law did you? Actually, you could even use
this method to get not only the mass of two galaxies going about one another, but of a
whole bunch of galaxies going about each other. All of their masses are influences on
one another, so they would all feel one another's pulls and would all obey Kepler's
Third law, but in this case one of the M values in the formula is the mass of the whole
group of galaxies.
Figuring out the luminosity of a galaxy is pretty tricky, since sometimes it is difficult
to figure out where the end of a galaxy is. Longer and longer exposure pictures of
galaxies (spiral galaxies especially) show larger and larger sizes and more and more
stuff. Just where does a galaxy end? At some point astronomers have to say enough is
enough; we'll only measure out to a certain distance and add up all the light to this
distance. Then the guessing game is played again - how much light are they missing
by not seeing the stuff beyond this point? How much light are they missing that is
covered up by dust and gas within our galaxy or by the dust and gas within the galaxy
they are looking at? What about the light sources that produce only IR or UV light?
There are quite a few fudge-factors that come into play in luminosity estimates for
galaxies because of the limitations we run into. Typically, we measure galaxy
luminosities in terms of the Sun's luminosity.
In spite of all of these problems, once the mass of a galaxy is known, it can be
compared to the luminosity of the galaxy. Why? You know that astronomers like to
do weird things with numbers. Actually, we just don't compare these numbers; we
divide them. This gives us the M/L ratio - Mass to Luminosity ratio. This is sort of
an important thing. What if galaxies are made up of average stars that are like the
Sun? If that were the case, you'd expect to measure the same amount of mass as you
would luminosity from the galaxy, so the M/L value should be close to one. What if a
257 | P a g e

galaxy is ultra luminous or abnormally bright for the amount of mass contained within
it? If that were the case, there will be a large L value compared to the M value, so
M/L would be low (much less than one). The third possibility is if there is a ton of
material but it is really dark and not giving off much light, so there would be a huge
amount of M but not so much L, and then we would have a large M/L value (much
larger than one).
That's all fine and dandy, but what do we see for the M/L values of galaxies? For
spirals, the M/L value is about 35, and for ellipticals, the M/L is about 70! What does
this mean? It tells us that there is a great deal of dim stuff in galaxies - it makes up a
lot of mass but doesn't show up in our telescopes (doesn't produce much light). This
means we are missing out on seeing a lot of the stuff that is out there. Remember, the
only way we can see stuff is if it produces light, and with such high M/L values, we
know that we are missing a lot of stuff.
It is sort of hard to determine the distribution of the different types of galaxies, such as
determining how many ellipticals, spirals, and irregulars there are. The problem is that
at great distances, the really small galaxies would escape detection. Even if there were
many small galaxies (like dwarf ellipticals, dwarf spheroidals or small irregulars), we
would have a hard time seeing all of them. Some astronomers say there are more
irregulars, while others believe that there are more ellipticals (especially the dwarf
ellipticals and spheroidals). We'll need more information before we can settle this
debate. Where have I heard that phrase before?
Galaxy Evolution
You've seen that there is quite a wide variety of galaxy types and that the content,
sizes, shapes and masses of galaxies are very diverse. Is there any link between the
various types? For example, if a galaxy observed today is an SBc, will it always be an
SBc or can it change into an Sc, SBb, or whatever? How does a galaxy end up being a
spiral, an elliptical or an irregular? What determines these things? Why is it that
someone always calls you just when you're ready to sit down to eat dinner? Oh, I
guess this last question really has nothing to do with galaxies.
Let's tackle the easiest parts first. We know that ellipticals have sort of spherical
blobby shapes and no new star formation in them. This means that they made their
stars very early in their lives and then that was it - no more star formation - sort of a
like a kid who eats all of their Halloween candy at one time. They end up rather
blobby and spherical. Spiral galaxies, on the other hand, are still making stars, so they
didn't use up all their gas when they formed. They are like the kid that slowly rationed
out the candy, eating only a little at a time so that it lasts longer. Why would they do
258 | P a g e

either of these things - using up all their gas or rationing it out? What could be the
reason for different galaxies to act in different ways?
One of the simplest explanations is to look at how a galaxy moves. An elliptical
galaxy has all of the material moving in a swarm about the center, while a spiral has a
well organized disk that is rotating. Why are they different? One of the reasons that
things may move the ways they do is angular momentum. You'll remember this was
the thing that makes the ice skater spin faster when the arms are brought in. Angular
momentum also has some say in how things will move and how fast they will move.
If an object, like pizza dough, is spinning really fast, it will tend to flatten out in a disk
as gravity pulls the material in. It is thought that this is what was happening with
spirals, that they were spinning pretty fast when they formed, so that the gas didn't all
blob together in the middle but got spread out into a disk. If the gas is spread out it
isn't going to go into making stars immediately (remember, you have to compress it to
a certain degree). Elliptical galaxies, on the other hand, may have originally been slow
spinners, so they didn't flatten out, and the gas just globbed together quickly (due to
gravity) and formed all the stars right away. If this idea about how they rotate is
correct, then we would expect to see more elliptical galaxies forming before spirals
forming, since the ellipticals' star formation isn't delayed. We do actually see this. At
great distances we see many well formed elliptical galaxies but very few fully formed
spirals - they are taking their sweet time to form. As we'll see later, having ellipticals
form quickly can have some interesting consequences, especially when it comes to
their cores (or more precisely, what forms in their cores - but I don't want to give that
away).
Let's get to some specifics of the life cycle of a spiral. It is now thought that spiral
galaxies can change quite a bit over their lives by having bars form in them and then
having these bars dissipate over time. The bar can redistribute the material in the disk
and the bulge over time and alter the rate of star formation, the chemical composition
and the galaxy type. Other things can also alter the shape of a spiral, such as having
another galaxy get a little too close to it and re-arrange the material. It is possible for a
galaxy to form as a type Sc, become an SBc, then an SBb, then an Sb and then an Sa
through various processes. Of course, these changes will take millions or billions of
years to occur, but they can happen nonetheless. There is a bunch of information on
this idea at the Hubble Web site.
Figure 2. Click on this icon to see several examples of how a galaxy
can change. The rate at which it collapses, interactions with other
objects or the influence of a bar can all play a role in altering the
shape or content of a galaxy.
259 | P a g e

Irregular galaxies tend to be produced by collisions or near-collisions of galaxies. The
strong gravitational pull of big galaxies on little galaxies can cause the material in the
little galaxies to be completely stripped off or re-arranged in a bizarre shape. There
are also cases where the little galaxy can screw up the big galaxy - no one is immune
from the forces of gravity; it just depends upon how things come together and mess up
one another. Our own galaxy is going to collide with the Andromeda galaxy in about
two billion years. A simulation showing how that may look to someone located far
away is provided here (animation visualization by Frank Summers (Space Telescope
Science Institute). Simulation by Chris Mihos (Case Western Reserve University) and
Lars Hernquist (Harvard University). Since this collision will start before the Sun
dies, it could result in our solar system being flung out of the galaxy. When the
collision finally finishes (in about 5 billion years) and everything settles down, the
Sun (and the planets) could be located far from the center of the newly merged galaxy,
possibly in the outer halo of the newly formed galaxy. Or it could be found amongst
the other stars in the newly merged elliptical galaxy. Either way, it's something we
can look forward to.
Figure 3. An image of a group of galaxies
known as Stephan's Quintet. These
galaxies are interacting with one another.
If you click here, you can see a computer
simulation showing how these galaxies are
possibly moving and interacting. Image
credit: N.A.Sharp/NOAO/AURA/NSF.
Animation credit: Credit: Joshua Barnes
(University of Hawaii).
The giant elliptical galaxies (cD types) are
thought to be produced by cannibalism,
and this idea tends to gain favor because there are often remains of the "victim"
galaxies still visible in the cores of the cD galaxies. When you look at the locations of
cD galaxies in groups of other galaxies, they tend to be right in the middle, just like a
big, fat spider in the middle of a web. It doesn't necessarily mean that the cD galaxy
will eat up all of its neighbor galaxies, but if something does get too close, it better
watch out.
Galaxy Clusters
Once people realized galaxies are objects outside of our own galaxy, they noticed that
galaxies are not randomly scattered across the Universe. They tend to be found in
groups or clusters - what an original name!
260 | P a g e

Figure 4. A typical galaxy cluster, in this case
the Virgo cluster. This one is rather large
having a few thousand galaxies in it. Only the
brightest are visible here. Image courtesy of
NOAO/AURA/NSF.
Our own galaxy is in such a cluster known as
the Local Group. This contains around 70
galaxies with the big ones being our own Milky
Way, the Andromeda Galaxy (M31), and the
Triangulum Galaxy (M33). There are many
smaller galaxies including the Large and Small
Magellanic Clouds, the little ellipticals that
hang around Andromeda and a bunch of others.
Actually, the little galaxies (irregulars and
dwarf ellipticals) outnumber the big galaxies by
a wide margin. The extent (diameter) of our
cluster is only around 1 million pc. It is such a pain to write out all of those zeros that
we'll just define a new unit of measure called the Megaparsec (Mpc), which is a
million parsecs. There is also the unit of a Megalightyear, which is obviously a
million lightyears. These two units are often used to describe distances to galaxies or
the sizes of galaxy clusters.
Figure 5. The Local Group seen in three dimensions.
If you click on the image you can see how the
members of the Local Group are spread out. The
Milky Way is identified. The other large black dot
indicates the locations of Andromeda and the
Triangulum galaxy - in this graph they are so close
together that their dots merge. The other small blue
and red dots are all of the other galaxies, mainly
irregular and dwarf galaxies.
While you might think that the Local Group is pretty good sized, there are actually
other clusters that are larger, and I mean much larger. Some clusters have a great
number of galaxies and cover a correspondingly larger span of the Universe than our
little Local Group. Clusters can contain thousands of galaxies spread over dozens of
Mpcs. Generally, clusters are distinguished mainly by their size. Those with a lot of
galaxies are termed as Rich, while those with few galaxies are known as Poor.
261 | P a g e

Figure 6. The Coma Cluster of galaxies.
This picture shows hundreds of galaxies.
Nearly every object in this picture is a
galaxy - only a few of the objects are stars
located in our own galaxy. This is clearly a
rich cluster. Image credit: Omar Lopez-
Cruz & Ian Shelton/NOAO/AURA/NSF.
Clusters are the way they are because
of gravity. One of the effects of gravity is
to bring some of the galaxies into close
proximity, which can lead to galaxy
collisions or mergers. Astronomers find that
there are many collisions going on, or have
gone on in the past. In some clusters
collisions can be on-going or happen every few billion years. In the end the galaxies
in the cluster could all merge into one large central galaxy - usually an elliptical. But
that would take a very long time. You can see a simulation of a galaxy cluster's
motion here - this simulation spans nearly the entire history of the Universe, around
14 billion years.
Based upon the ages of stars in them, we know that galaxies have been around for
billions of years and they are moving around in clusters at pretty good speeds, yet the
clusters stay together - the galaxies don't generally drift or fly away. They are
gravitationally bound systems - the galaxies are orbiting about one another and stay
together due to their mutual gravitational attraction. If we can determine how fast the
galaxies are going then we can figure out how much mass is needed to keep them in a
group - how much gravity is needed to counteract their outward velocity. This is
another way of measuring the masses of galaxies, or even the masses of a whole group
of galaxies at once.
If you remember back to the discussion of galaxy masses and luminosities, it seemed
that we always detected more mass than light from galaxies. Is this the result we get
when we look at a whole bunch of galaxies? Yes! When we look at whole clusters of
galaxies, we tend to see much less material than there needs to be to hold the group
together, or putting it another way, there is a very high M/L ratio. I'm really getting
sick of this high M/L thing - why do we always detect more mass than we can see?
Perhaps we're just using the wrong types of telescopes? Perhaps a lot of the matter in
galaxies and clusters of galaxies is only visible at IR, UV or radio wavelengths? I'm
sorry to say that's not the answer. We've tried to look for the material at these
wavelengths and we still don't see enough stuff to account for all of the mass that is in
the galaxy clusters or the individual galaxies. We just aren't seeing the stuff at all!
262 | P a g e

Now things are going to get a little silly. It appears that a large fraction of the stuff in
galaxies or even galaxy clusters is not visible to any telescope. We're basically saying
that we have a hard time seeing most of the material that makes up even our own
galaxy. What could this stuff be? This is that same sort of thing we saw with the
Milky Way Galaxy, the presence of a large amount of unseen material. If you
remember, we refer to this stuff as dark matter - sort of a cool name, don't you think?
Dark matter comprises 90% or more of the total mass of a cluster of galaxies. That
means when astronomers look at groups of galaxies, or even our own galaxy, they can
only see 10% of the stuff that is in the galaxy or the galaxy cluster. We are not seeing
most of the stuff! That is like someone who studies plants or flowers only being able
to see one petal or leaf from the plant. We're really missing most of the stuff that is
out there.
If most of the stuff that is out there is dark matter, what is it? What stuff can be so
abundant, yet remain undetectable? There are two main ideas as to what type of stuff
comprises the dark matter. One idea was put forth by people who do physics - there
are a lot of funny little particles that exist in nature, like the neutrino, which are really
hard to detect. Such particles are hard to detect because they don't interact with things
very well, sort of like Patrick Swayze in "Ghost" - you don't know he is there most of
the time unless you are Whoopi Goldberg, but even then you aren't detecting all of
him. Neutrinos are like this - they don't interact with much stuff, so we don't know
how many of them are out there. Now what would happen if a neutrino has mass, even
just a little mass? There are so many neutrinos, you would have a bunch of mass that
would not be visible! This is exactly what we are looking for - a particle that escapes
detection, is very abundant, and has some mass. This is one of the possible forms that
dark matter can take. We just can't call it a neutrino since it could be some other
particle, so we have to give it a really cool name (mainly because we don't know what
it really is) - WIMPs, which stands for Weakly Interacting Massive Particles. This
could be just what we were looking for. The only problem with this idea is we don't
know if these types of particles exist - remember, we can't detect them very well. It is
possible that there are other things out there besides neutrinos which we can't detect at
all, so we may never be able to say for sure if the WIMPs are the stuff that make up
the dark matter.
If the WIMPs aren't the stuff that makes up the dark matter, then what is?
Astronomers probably didn't like having the physics people come up with the answers
to an astronomical problem, so they came up withMACHOs (not because
astronomers are tough and physicists are wimps, or perhaps that is the reason...),
which stands for Massive Compact Halo Objects. What does that mean? It means
these are massive objects, but they are small in radius and they are found in the halos
of galaxies. What sort of massive objects are these? We've already run across quite a
263 | P a g e

few exotic, compact objects like black holes, neutron stars, and white dwarfs in our
discussion of stars. There are also all those really ultra-faint L and T type stars that
would be difficult to detect. What if there were tons of these little stars or lots of black
holes out there? They would contain a large amount of mass in them, yet they would
not be producing very much light. This is just we are looking for. Some astronomers
are actually trying to find MACHOs out there. They do this by watching a whole
bunch of stars - millions of stars each night. If a MACHO were to pass in front of one
of these stars, the light from the star would be effected in a predictable manner. Now
this sounds like a really boring thing to do - watch several million stars to see if any of
them flicker, but that is exactly what is done. Various projects that are doing this have
detected some MACHOs in the halo of our galaxy. We're still not sure if we're seeing
as many MACHOs as we should so that they could account for all of the Dark Matter
in our galaxy, but some of them have been found.
MACHOs are sort of the "normal" objects, while WIMPs are the "abnormal" objects
that can make up Dark Matter. Which is it? Is the Dark Matter in the form of WIMPs
or MACHOs? We have detected small amounts of both (remember, we can detect
neutrinos from the Sun and supernovae), but we're still not sure of the numbers. It is
possible that the Dark Matter is made up of various amounts of both WIMPs and
MACHOs, though we're not sure of the amounts. There are many projects involved in
the search for Dark Matter and here is a list of some of those projects. Even though we
don't know if it is made of WIMPs or MACHOs we're still working on it, so I can't
give you a definitive answer. Just give us a few more years and perhaps I'll be able to
tell you something more substantial - or perhaps not - who knows?
Superclusters
After astronomers started grouping galaxies into clusters they noticed that on larger
scales the clusters tended to be grouped together in clusters or what we
call Superclusters (what a clever name - sounds almost like a superhero). Our cluster,
the Local Cluster, is part of the Local Supercluster, which contains about 10 large
(rich) clusters - our cluster is one of the many small clusters in it. Actually, in the old
days the Local Supercluster was referred to as the Virgo Supercluster, mainly because
the Virgo Cluster was one of the bigger parts of it.
Figure 7. A graph showing the locations of superclusters
around our own (we're in the middle). This map shows
the material within about 1 billion lightyears distance
from us. Image is from Richard Powell's website.
If there are clusters of galaxies and clusters of
clusters, are there clusters of Superclusters? This is
264 | P a g e

getting silly, but it isn't a silly question, especially when you look out into space and
see how the superclusters are arranged out there. It appears that there is a kind of
structure to the Universe; things are not smoothly distributed. Clusters and
superclusters tend to be grouped together. When we map out large areas of the sky
and look at the distances and locations of galaxies, clusters of galaxies and
superclusters of galaxies, we see some very large scale structure in our Universe. The
image below is from a survey of the sky in two directions from the Earth. Notice how
many bubbles and holes there are in the image - stuff is clumpy!
There are regions with very little matter (called voids) and other regions with large
concentrations of galaxies (superclusters or even clusters of superclusters). Matter
isn't evenly spread out but appears to be rather clumpy. This is an important
characteristic of the Universe that we'll run into it later. There are many surveys that
are mapping out this structure to our Universe (often refered to as Large Scale
Structure) and the results from the various surveys are consistent - the Universe isn't
smooth!
Figure 8. A graph of around 220,000
galaxies surveyed in two directions of the
sky. The Earth is in the middle of the
wedges. Each dot represents a galaxy. To be
precise, the sizes of the dots are much too
large (not to scale), so there is really not as
much crowding in the distribution of
galaxies. With this in mind it is obvious that
the galaxies aren't uniformly spread out but
clumped together in superclusters or even larger scale features. There are also many
regions where there are no galaxies, which are very empty stretches of space. Click
on this image to get the larger version of it. Picture is from the 2dF Galaxy Redshift
Survey.



Now that you've read this section, you should be able to answer these questions....
What methods do astronomers use to determine the mass of galaxies, and what
are the limitations of those methods?
What is the significance of the Mass-to-Luminosity ratio?
What influence does angular momentum have on the evolution of galaxies and
the formation of various galaxy types?
What role does cannibalism play in galaxy evolution?
265 | P a g e

How large is our local galaxy cluster? What are the major members in the
cluster?
What are the possible types of dark matter?
What evidence exists in galaxy clusters that indicate the presence of dark
matter?
What is a supercluster?
What is the general structure of the distribution of galaxies in the Universe?
Distant and Weird Galaxies

What's covered here:
What is Hubble's Law all about?
What defines an active galaxy?
What different types of active galaxies exist?
What is a quasar?
What is the most distant object yet observed?

One thing that people started to do when they figured out that those fuzzy things out
there were actually separate, distant galaxies was to get as much information about
them as possible. Now this isn't an easy thing to do, especially when you consider just
how far away these things are. There are only certain things that can be observed with
galaxies, and one of those things is the spectrum of the galaxy. Now different parts of
a galaxy are moving in different directions and at different speeds, so astronomers
thought that there would just be a bunch of random velocities observed in the
spectrum, but that wasn't the case. They started noticing that along with the random
velocities there was an overall general motion of entire galaxies through space. This
would be sort of like watching how people move on a jet plane. If you looked at any
given time there would be people moving up and down the aisles (and waiting with
crossed legs to use the bathrooms), but if someone were to ask you how fast those
people were actually moving, you would say something like a couple of hundreds of
miles an hour. Their small individual motions are pretty measly compared to the
plane's velocity - the same is true with galaxies. They can have some fairly high
velocities involving the stuff within them but those velocities can be pretty small
compared to the overall motion of the galaxy through space.
Now there is another type of velocity involved with galaxies, and these are the
velocities that galaxies have when they are moving around one another, like in a
cluster or binary pair. This motion can be seen as sort of random, since some galaxies
266 | P a g e

in clusters would be moving towards us as they orbit around other galaxies in their
cluster, while others are moving sideways to our line of sight and others are moving
away from us. One galaxy that is moving towards us is the Andromeda Galaxy. As
mentioned in the previous set of notes, that galaxy will actually run into ours
sometime in the future - I don't think our insurance has coverage for such an event.
Galaxies move around one another in clusters at velocities that can be 100s of km/s -
pretty fast.
We know that these motions are occurring, since we can see the evidence of such
motions in the spectra of galaxies. If you were to make a survey of thousands of
galaxies, what sort of velocities would you expect to see? Would you expect to see all
sorts of different velocities due to the random motions of galaxies in their clusters?
After all, galaxies are just moving around in various random directions, aren't they?
Aren't some going to be moving toward us while others are moving away? It should
be pretty random, right? Am I ever going to stop asking questions and get to the
point?
Here's the weird thing - when you look at thousands of galaxies, you see pretty much
one type of motion - and that is motion away from us. It looks like pretty much every
distant galaxy out there is moving away from us! All of their spectra show large
redshifts (remember, velocity away from an observer shifts the spectral features to
longer, or redder, wavelengths). Now not every single galaxy has a redshifted
spectrum, but the number of blueshifted spectra seen in galaxies is pretty sparse, and
that is mainly for galaxies that are near us. Apart from these few galaxies, pretty much
all the others are moving away from us.
Now this is a really bizarre thing - but what does it mean, and what does it tell us
about galaxies and the Universe? Just measuring the spectra and getting the velocities
is only one part of the answer. It took some careful detective work to find the other
piece of information to figure out what was going on, and that piece of information
had to do with the distances of the galaxies. One fellow noticed that there was actually
a trend in the velocity values with the distance values. That fellow was Edwin
Hubble; you remember him, the guy who figured out that galaxies were separate,
distant objects. He used his methods of obtaining distances to galaxies (by
using Standard Candles) and combined that information with velocity data (obtained
from the spectra of galaxies). Just how did he combine that data? - in a graph, of
course.
If you plot up this data, you find that the greater the distance a galaxy is from us, the
faster it is moving away from us. Put another way, distant galaxies are receding
from us at great speeds. The greater the distance, the greater the speed. This
267 | P a g e

diagram, which was first constructed by Edwin Hubble is known, surprisingly, as
a Hubble Diagram.
Figure 1. A Hubble
Diagram, simply a plot of
galaxy velocity versus
galaxy distance. This
diagram is based upon
distances found using
Supernovae. The slope of
the line gives the value for
the Hubble Constant, H
o
.
Now this diagram will
only work on distant
galaxies - nearby galaxies
don't show this sort of
trend since they are
moving randomly around in their clusters - sometimes toward and sometimes away
from us. For distant galaxies there is a direct correlation between the speed a galaxy is
moving away from us and how far away it is from us. There must be a formula that
relates these two things, distance and velocity. Of course there is! The data appear to
fall along a straight line that can be drawn through the data, so the line represents the
slope of the formula, and it can relate the two quantities. Guess what we call this
formula? You guessed it; it's known as the Hubble Law.
Mathematically, the formula is written out as
v = H
o
d
where, v=velocity (in km/s), d=distance (in Mpc), and H
o
is the slope of the formula
and appears to be a constant, which we'll call the Goober constant - no, that's silly, we
better stick with the trend here and call it the Hubble constant (its units are
km/s/Mpc). This is a pretty simple formula, eh?
Not only is this formula pretty simple, but is probably one of the most important
formulas ever written. THIS IS REALLY, REALLY GREAT!!!! Why? There are
three reasons.
1. If you remember the discussion of Standard Candles, you know that there are some
problems in finding accurate distances. The Hubble Law provides astronomers with
another method for finding distances - so long as you know what the value of H
o
is.
How do we get H
o
? You need data from galaxies, in particular their distances and
velocities. Just take this data and plot it up like Hubble did, draw a line through it and
268 | P a g e

you'll get a value for the Hubble Constant, since it is equal to the slope of the line.
That sounds easy; get velocities and distances of galaxies so we can use the Hubble
Law to find distances to galaxies. Wait a minute, if we want to use the Hubble Law to
determine distances, we need to first have distances to determine what H
o
is before we
can use it to find distances - that's silly. This doesn't make sense. How can we find
distances with the Hubble Law if we first need to have distances to get the constant?
That's sort of the problem with the whole thing.
Astronomers can try many times to determine the value of the Hubble Constant using
distances measured with Standard Candles and velocities obtained from galaxy
spectra, but they rarely get the same value for the constant. This has been one of the
biggest problems in astronomy since Hubble first did this. Each time someone tries to
determine the value of the Hubble constant, they get a value that doesn't agree with
anyone else's value, so there has usually not been much agreement on what value to
use for the Hubble Constant. One of the reasons that astronomers don't agree with one
another about the value for the Hubble Constant is that they use different methods or
assumptions in getting distances, so they will get different values for H
o
. Astronomers
who do this experiment over and over even change their own ideas as to what the
value of H
o
is.
That's just dandy; how are we supposed to know what the value of the Hubble
Constant is? Actually, things have gotten better in the past few years. Astronomers
who have been using the Hubble Telescope (which was not built by Edwin Hubble, it
was just named for him) have been able to measure distances using Cepheids and
Supernovae in distant galaxies. Actually, several different groups of astronomers were
doing this, so of course they ended up with different values of the Hubble Constant.
However when you look at all of the data it tends to converge to a value of around 70
km/s/Mpc. This is actually pretty good; it used to be that the values for the Hubble
Constant that people calculated were widely divergent by large amounts. Typically, if
you wanted to use the Hubble Law to get distances, you would pick one of these
values or something in the middle, like 65 km/s/Mpc, and just go from there. Most
astronomers today use a value of 65-70 km/s/Mpc.
It is sort of a good thing that the Hubble Telescope was able to do what it was built
for, namely to help figure out what the value for the Hubble Constant is fairly
accurately. Now it's onto another reason why the Hubble Law is so cool.
2. What exactly is H
o
? What does it represent? According to the Hubble Law,
H
o
=Velocity/Distance = 1/Time. It is related to a time. What does this time represent?
The time value which H
o
represents is the time it took for the Universe to expand to
its current size, or another way of saying it is that it is a rough estimate for the age of
the Universe - good gravy, that's one of those basic questions of life, the universe and
269 | P a g e

everything. That's sort of an important number - not as important as your student ID
number or your weight, but it is one of those numbers that have been on people's
minds for a long time.
3. The last reason why the Hubble Law is so valuable is why it even exists at all.
That's a rather interesting thing - why does it exist? Why are those more distant
galaxies moving away from us at greater velocities? The Hubble Law exists because
of what the Universe is doing - it is Expanding Uniformly! If the Universe weren't
expanding or moving in some way, there would be no trend in the velocities. Hubble
wouldn't have been able to make that nifty plot of his if the Universe were sitting still.
We'll look at this aspect more closely later.
The Hubble Constant is a pretty important thing. Why can't we figure out what its
value is simply? It is easy, isn't it? No, it is not easy at all, but is a rather complex
step-by-step process. There are many places to mess up in the process.
To determine the scale of the Universe, you must first start closer at home. You need
to determine the distance to the Sun, the AU. Why? The concept of parallax is based
upon our orbit about the Sun, so the size of a parsec depends upon the size of the AU.
Once we know how big an AU is, you can start to find the distances to nearby stars
(using the parallax method), then the distances to more distant stars (using the
properties of the nearby stars and methods such as spectroscopic parallax), then the
distances to nearby galaxies (using various Standard Candles, which are based upon
the characteristics of those stars that you looked at previously in your own galaxy),
then the distances to more distant galaxies (using the characteristics of the nearby
galaxies) and so on.
As you go further out, more assumptions, more uncertainties, more inconsistencies
between various scientists, more guess work, and more shaky ground pops up. No
wonder those silly astronomers can't agree; it's too much of a mess to begin with!!!!
Distant Galaxies
As we look further out into space, we see objects as they appeared further in the past.
You have to remember that light travels at one speed, so you can't see something until
the light from that object gets to your eyeballs. It is sort of like waiting for the post
office to deliver a message. You can't read the message until it arrives. If the distance
is large, the delay in the message is greater. If you were to look at the Sun in the sky,
you would not be looking at where the Sun currently is, but where it was about seven
minutes ago, since that is how long it takes for light to go from the Sun to the Earth.
You could also say that the Sun is seven light minutes away. Alpha Centauri is four
light years away, so we are seeing it as it looked four years ago when we look at it in
270 | P a g e

the night sky. The center of our galaxy, Sgr A*, is about 25,000 light years away, so if
there were a great big explosion in the middle of our galaxy today, we wouldn't see it
until about 25,000 years from now. When you start to look at galaxies, the time delay
is even worse. Even our nearest large neighbor, the Andromeda galaxy, is about 2.25
million light years away.
When we look at galaxies that are billions of light years away, we are seeing them as
they looked billions of years ago. They are so far away and the light from them takes
so long to get here that we don't see them as they look today, but as they looked
billions of years ago, when they were younger objects. As we look further and further
out, the objects we see are younger and younger. If you remember how young stars
were sometimes rather active and did strange things then you may not be surprised to
learn that young galaxies are also rather peculiar. Many of these young galaxies are
quite unusual - even more unusual then young stars. They go through phases where
they will emit abnormally large amounts of light, often at unusual wavelengths, such
as radio or x-ray. Sometimes these galaxies also have unusual colors, spectra, and
shapes.
We generally throw all of these unusual galaxies into one main group, those that are
labeled as Active Galaxies. Sometimes only the centers of these galaxies are visible,
in part because they are so abnormally bright. In those cases, we may just refer to
them as Active Galactic Nuclei. Either way, these are amongst the strangest things
out there. Actually, about 10% of all galaxies are active, so you can't ignore them. I
suppose you could,
but that wouldn't be
very polite. Here are
some examples -
Figure 2. The
spectrum of a BL Lac
galaxy is compared to
that for a normal
galaxy. In the normal
galaxy there is a mix
of absorption lines
(from mainly stars)
and emission lines
(from hot gas clouds).
In the BL Lac galaxy
there are neither of
these things, which
makes it difficult to
271 | P a g e

determine the chemical nature, velocity and distance of these galaxies. Based on the
image from Bill Keel's slide set.
BL Lac Galaxies, which are also called blazars, were misnamed. The first one
discovered was given a star name (BL Lac), in particular a name associated with stars
that change brightness. This was a mistake, since we know that these are actually
galaxies with very bright central regions, so bright that they look like a bright star
when seen at great distances. It took some careful studies of their spectra and detailed
telescopic studies to show that these are actually the bright cores of galaxies. Once
that was figured out, astronomers started to notice all of their bizarre features.
By observing them for long periods, it was noted that they tend to have variable
brightnesses at different wavelengths (got brighter and fainter over time).
Observations of their spectra show there are virtually no emission or absorption
lines - this is a fairly unusual characteristic for a galaxy, especially an Active
Galaxy.
The radiation coming from these objects is known to be non-thermal (not
related to heat). The spectra do look rather uniform and flat, close to being
continuous spectra. The optical light from these objects is also highly polarized,
indicating the influence of a strong magnetic field.
Blazars produce light at all wavelengths, ranging from being strong radio
sources all the way to
being sources for
gamma-rays.

Figure 3. The changing
brightness of BL Lac is
illustrated over the years. This
shows the change in the radio
intensity of the galaxy. Based
on the image from Bill
Keel's slide set.
One of the annoying features of
blazars is that because they
don't have any spectral features
that can be seen clearly (no
emission or absorption lines), it is difficult to determine their velocities or (by using
Hubble's Law) their distances. Actually, blazars are kind of rare when compared to
some of these other strange galaxies.
272 | P a g e

Seyfert galaxies are the next strange beasts in the astronomical zoo. There are
actually two types of these. Like BL Lac, Seyferts show very abnormally bright cores.
However, unlike BL Lac galaxies, these have emission lines (hurray!). One aspect of
the emission lines is how they can appear as either broad or narrow - this is one of the
things that caused astronomers to divide them into two categories. Now why would an
emission line look narrow or broad? Don't they all look the same? - not really.
Figure 4. Spectra for the
two types of Seyfert
galaxies. Notice how the
emission lines for the
Type I Seyferts are wider
than that for Type II
Seyferts. Based on the
image from Bill
Keel's slide set.
Emission lines are
produced by hot gas. If
the hot gas is moving at a
pretty good rate of speed,
then the Doppler effect
comes into play - the
emission lines could be shifted to different wavelengths. If the hot gas is moving at
many different speeds, some fast, some slow, some toward, and some away, then the
Doppler effect would cause the spectral features to appear at a wider range of
wavelengths than if it was just sitting still and being boring. Therefore (I like that
word; it sounds so final), the presence of broad emission lines tells us that there are a
lot of different velocities associated with the gas that is producing the emission lines
and that these motions are large and varied in direction. Perhaps this is a clue as to
what is going on in these objects...
Here are some gory details about the two types of Seyferts -
First of all, both types are very luminous, and they sometimes change their
brightnesses (or the brightness of their cores) on relatively short time scales
(days or weeks). For these things to change brightness so quickly, they must
have relatively small energy sources - so the place where all the weird stuff is
going on must be pretty small (perhaps this tells us something about what these
things are like?).
When Seyferts are viewed much more closely, it seems that they have shapes
like spiral galaxies, except of course, for their bright, active cores (the link here
273 | P a g e

shows Hubble telescope images of the cores - note how bright and star-like
they are).

Now to get down to the nitty-gritty differences between the two types
Type I Seyferts
o There are broad spectral emission lines in them. The widths of these
lines indicates velocities around 1000 km/s.
o These are also very bright sources of UV and x-ray light, as well as the
normal visible light that comes from them and their very bright cores.
Type II Seyferts
o If the other type has broad emission lines, then these guys have narrow
emission lines, which indicates that they have very low velocities
associated with them.
o Along with having the bright cores, they are also bright at IR
wavelengths
First blazers, now Seyferts - quite different, but are they really that different? You'll
see...
That's enough of the weird galaxies; now for the galaxies that just look really weird
when we see them in different ways. This sort of has to do with the different types of
technologies that have come along over time. After World War II, radio telescope
technology really exploded on to the scene (not literally; you know what I mean). As
the technology got better, our views of some things changed. What looked like normal
galaxies with visible light telescopes looked really strange with radio telescopes. In
some cases, galaxies were producing more light at radio wavelengths than at the
visible light wavelengths. Galaxies that do this or do other weird things at radio
wavelengths are called Radio Galaxies. What we've got here are galaxies that at
visible
wavelengths look
like regular
galaxies, but pull
out your radio
telescope and you
got a completely
different, bizarre
galaxy.
Figure 5. A typical
radio galaxy. The
274 | P a g e

galaxy itself is not much larger than the dot that represents its location in the image.
The sizes of the lobes are much greater than the size of the galaxy they come from.
Also visible are the jets that come from the galaxy's core toward the lobes. Images
from NRAO.
Generally, when a galaxy has some sort of strange radio light coming from it, it tends
to come in two forms - either there is a strong source of light from the core, which
isn't so exciting, or it is spewing out radio emitting gas in various directions. This is
sort of what people think of when they think of radio galaxies. In the case of the big-
time spewage, there is usually some sort of structure that shows the motion of material
from the core, generally in a jet type of structure (sort of like that bipolar outflow you
have with accretion disks around black holes). The stuff that is getting spewed will
pile up over time, generally into big lobes. A classic view of a radio galaxy shows a
double lobe structure. These lobes are huge compared to the galaxies that are spitting
them out - they can be millions of light-years in size (galaxies tend to only be only
thousands of light years in size). These great big lobes are nothing more exciting than
radio emitting hydrogen gas. It's just that there is a lot of it out there!
Cen A is a classic example of a Radio Galaxy. In the visible light view, it looks like a
galaxy with a large dusty disk, a little bit unusual but not too bizarre. With the radio
telescope, you get a totally different picture. Notice how the lobes come out of the
disk - they are actually perpendicular to it.
275 | P a g e


Figure 6. Cen A, as seen in multiple wavelengths. On the left is a visible light image.
In the center is a radio (orange) and visible composite. On the right is a composite of
visible, microwave (orange) and x-ray (blue) light. These images help astronomers to
understand the core of Cen A and the jet features. You can read about the images,
view movies and learn about what astronomers think is happening at theNASA
website. Left image from Capella Observatory. Middle image from Capella
Observatory (optical), with radio data from Ilana Feain, Tim Cornwell, and Ron
Ekers (CSIRO/ATNF), R. Morganti (ASTRON), and N. Junkes (MPIfR). Right image
from ESO/WFI (visible); MPIfR/ESO/APEX/A.Weiss et al. (microwave);
NASA/CXC/CfA/R.Kraft et al. (X-ray).
Another good example of a radio galaxy is M87, an elliptical with a very strong jet
extending from it. M87 is relatively close, so the jet is seen even in some visible light
images. The jet is about 1800 pc long (about 6000 light-years). Hubble Space
Telescope observations of the core indicate motions of material around 550 km/s
about the center. By using Kepler's laws we can estimate the mass of the core to be
276 | P a g e

about 2.5 billion solar masses. For reference, the mass at the center of our galaxy is a
few million solar masses. That is a big difference!
Figure
7. Imag
es of
M87,
an
elliptic
al
galaxy
with a
jet. The
view at
the left
shows
the
regular
telesco
pic
view
(image
from
NOAO/AURA/NSF). The top right view is a radio telescope image, which mainly
shows the large jet coming out from the center of the galaxy (image from
NRAO/NSF). The lower right image is from the Hubble space telescope and it shows
the jet in more detail. Data from the Hubble telescope were used to measure the
velocity of material near the center of the galaxy and estimate the mass of the black
hole in the center. Hubble image from NASA and John Biretta (STScI/JHU).
Radio galaxies can come in a variety of shapes - they can be spirals or ellipticals,
though it seems that the most extreme ones tend to be ellipticals.
Now I suppose you think that these bizarre objects are out there at very great
distances. Actually, some of these strange radio galaxies are relatively close. The
really weird galaxies, the ones that astronomers have still not figured out, are amongst
the most distant objects out there. Let's take a look at these really extreme galaxies...
Quasars
Quasars were originally thought to be stars with very strange spectra. If you look at a
quasar you usually see a point of light - just like what you see when you look at a
277 | P a g e

normal star in our galaxy. Remember, galaxies traditionally look like fuzzy blobs, not
points of light. Quasars (short for quasi-stellar radio source) stood out from normal
stars because they had weird spectra. First of all, they have emission lines, which is
unusual for stars. Not only that, the spectral features did not correspond to any normal
spectral features. When you look at a star, the spectrum usually has normal elements
in it, like hydrogen, carbon, iron, etc. The wavelengths where the emission lines were
seen didn't correspond to any of these things. They were just weird.
Figure 8. On the left, an image of a typical quasar (the object in the middle with the
two lines pointing to it) as seen from the ground. Notice how much this looks like the
stars that are also in this image and are actually within our galaxy. To the right is a
collection of images of quasars taken by the Hubble Space Telescope. In this picture it
is possible to see hints of the galactic structure around the quasars. Image on the left
from Bill Keel'scollection, image on the right credits: John Bahcall (Institute for
Advanced Study, Princeton) Mike Disney (University of Wales) and NASA.
In 1963, Maarten Schmidt was looking at the spectra of some of these quasars and
noticed that the spacing of the emission features was about the same as that for
hydrogen emission features, but they were well off their normal wavelengths. Maarten
did the calculations and found that the spectra for the quasars would be what you
would get if the features were redshifted to longer wavelengths due to the motion of
the object being about 45,000 km/s = 15% the speed of light! This is a pretty high
velocity for a star! After some more study, Maarten figured out that the object he was
looking at (which has the really romantic name of 3C 273) was actually fuzzy,
meaning that it wasn't a star but a galaxy with a really bright core. To put this in
perspective, you should know that 3C 273 is one of the slower quasars out there!
278 | P a g e

If we use the Hubble law, we find that these incredibly large velocities (many of them
good fractions of the speed of light) appear to correspond to incredibly large distances
- amongst the greatest distances measured!
We need to have word of CAUTION here about the redshifts. The redshifts of
quasars tend to be pretty large, so values of redshifts (the symbol for redshift
is z = / ) of two or three are not unusual. This means that the light has been
shifted to two or three times its current value (if normal wavelength for a spectral
feature is 1200, and the redshift, z, is three, then the observed spectral feature is seen
at a wavelength value of 1200 + 3x1200 = 4800 ). The formula for Doppler shift is
z= / = v/c
Does this mean that a redshift (z) of two, implies a velocity of two times the speed of
light?
No. When such high velocities are observed you need to use a different redshift
formula, one that is based upon Special Relativity. We have to use this formula since
nothing can go greater than the speed of light, c. TheRelativistic Redshift formula
allows values for redshifts (z) greater than one, but velocities are always less than c.
The Relativistic Redshift formula is

Quasars are among the most distant objects seen and they have some rather extreme
characteristics -
very fast moving - this is based upon the sizes of their spectra's redshifts. The
current maximum redshift for a quasar is one with z=6.4 (which corresponds to
a velocity of more than 96% the speed of light!). See Figure 9 for the spectra of
the current record holders
very far away - we know this since Hubble's Law says if you are moving very
fast, you are very far away
there are usually emission lines in their spectra, and often the spectral features
are broad, which indicate high speed motions within the quasar (sort of like
Type I Seyferts)
very luminous - we know this since we know that quasars have huge distances
but don't look really faint. You'd expect that very distant objects would also be
very faint, but this is not true for quasars; they tend to have respectable
brightnesses. Actually, quasars are often detected accidentally when people are
studying what they think are a bunch of stars in our galaxy. They are not hard
279 | P a g e

to detect. Based upon their distances and their apparent magnitudes we know
that they are thousands of times more luminous than normal galaxies like the
Milky Way.
sometimes the light source changes brightness over time, indicating that they
are variable energy sources
based upon the rate at which the quasars change their brightness, we know that
the energy source in them covers a pretty small region of space
lower redshifted absorption lines are often seen in the spectra of quasars. These
are not part of the quasar but are due to stuff between us and the quasar -
remember the light from a quasar has to travel a great distance, so there is a
pretty good chance that it will run into something along the way. Most of this
stuff is hydrogen, and hydrogen is often detected at a wavelength for a feature
known as Lyman alpha, so the proliferation of the absorption features is often
called a Lyman alpha forest (Figure 10).

Figure 9. Spectra of the currently highest
redshifted quasar and some that are

Figure 10. The Lyman Alpha Forest is shown in the
spectrum of the quasar on the bottom. The nearer
quasar's light is not absorbed by many intervening
gas clouds, so its spectrum doesn't have as many
absorption features as the more distant quasar.
Based on the image from Bill Keel's slide set.
280 | P a g e

nearly as fast. The Lyman Alpha feature
(large bump) is supposed to be at a
wavelength of 1216 ! Click on the image
to see the full size version. Image
from Sloan Digital Sky Survey. Image
credit: Donald Schneider and Xiaohui
Fan, SDSS Collaboration.
Now we have strange galaxies which go from being the most extreme (quasars) to
being just a little quirky (radio galaxies). Are they all drastically different, or are they
all just variations of one common set up? It might be easier to deal with them if there
is just one model or concept that can explain all of these weird things - that way we
don't have to come up with all sorts of different exotic ideas and figure out why there
would even be different exotic models. Sometimes in science, it's better to search for a
single solution than for many diverse solutions. Often the simplest one is correct. The
basic upshot is that astronomers think that there is one model that can explain all of
these strange galaxies - from the strange spectral features and the unusual energy
sources to the crazy structures that are seen. The type of active galaxy that is observed
may only depend upon the observer's point of view. Here is what the model is all
about.
Figure 11. The basic model for
Active Galaxies. It consists of a
massive black hole (located in
the center), an accretion disk
around the black hole, a bunch
of high velocity clouds
(yellowish ones around the
disk), low velocity clouds
(bluish ones located above and
below the accretion disk), a
dusty torus (cut open here so
you can see the stuff in the
middle), and jets going
perpendicular from the center
(cut off in this view). The
relative sizes of objects are not
correct here, but are just
shown so that you can get an
idea of where everything is. Of
course, the colors are not
281 | P a g e

realistic.
The model is made up of the following components
A black hole in the center (this can be as massive as millions or billions of solar
masses ). Remember, black holes can have a large amount of mass, but they
won't cover huge amounts of area (usually smaller than a solar system in
radius).
An accretion disk forms around the black hole. This would be much larger in
size than the black hole and would be very hot and dense near the center -
producing UV light and x-rays in most cases or even some gamma-rays. Even
though the accretion disk would cover more space than the black hole, it is still
relatively small compared to the size of the galaxy that it is located within.
Clouds of material would be orbiting around the disk. These clouds will be
heated up by the extreme temperatures of the accretion disk. These clouds have
to move at pretty high speeds so they don't fall into the accretion disk or,
ultimately, into the black hole. They are basically hot, fast, gas clouds near the
black hole.
The high energy accretion disk can cause the formation of jets perpendicular to
the disk. This is due to the strong magnetic field that arises from the hot
material in the disk. Stuff (mainly hydrogen) travels from these jets at huge
speeds, up to 99% the speed of light! These jets are visible in a variety of
wavelengths depending upon the amount of material they eject and their
temperature.
A dusty, wide torus (donut shaped feature) is around the whole thing. This
torus would be cooler and hard to look through (since it is dusty), so it can
block the view of the accretion disk, the black hole, and the hot clouds.
Located away from the high velocity disk and the black hole is where you
would find other clouds of material, but this stuff is moving at relatively low
velocities. Also, this material would tend to be found away from the dusty
torus.
How does this explain everything? First of all, not all strange galaxies will have each
component in equal amounts - some may not have very significant black holes, or jets,
or things like that. That is something that should be remembered.
How do the different active galaxies show up? Let's say you are looking along the
edge of the disk, which will be pretty boring, since the dusty torus blocks all the
action in the center. You aren't seeing any of the high velocities or high-energy action
in the center. All you are seeing is the dusty torus, which would be a good IR source,
and the low velocity clouds away from the center. This is a good way to explain the
appearance of Seyfert Type II galaxies.
282 | P a g e

Now let's look at it from an angle. Now we can see the action in the middle, including
all those high velocity clouds. This will produce broad emission lines. You are also
seeing stuff that is a source of high energy light, like UV, x-rays or maybe even
gamma-rays, so now you're seeing the stuff that we associate with Type I Seyferts.
Another direction you can look is straight down the throat of one of those jets. Now
you'd think that this would show you all sorts of stuff, but actually it doesn't, since the
bright core is not really easy to see since it is behind the material being shot out of the
core in the jet. You are really just seeing most of the stuff that is associated with the
jet that is coming toward you. That material is from a jet, which is produced by the
strong magnetic field of the accretion disk, so you would see non-thermal light from
this source as well as highly polarized light. This is what you'd get if you were to look
at a BL Lac (Blazar).
You can also see from most of these directions the ingredients for a radio galaxy,
though that would depend upon how much radio-emitting material is being ejected by
the jets. A quasar could be observed along the same viewing angles as the Type I
Seyfert views or some of the other angles, but since quasars are so distant you might
not be able to detect evidence of all of these other features. A summary of these views
in the active galaxy model is available for viewing here.
Does this model really make sense? In a way it does. First of all, in recent studies by
the Hubble Space Telescope and the Chandra x-ray telescope indicate that massive
black holes appear to be present in the cores of most galaxies. In fact it appears that
black holes are quite common and very active in most galaxies, and remember, you
can determine their masses by seeing how objects move around them (good old
Kepler's laws). Hereis a news item about such a black hole, one that is only a few
million solar masses in a distant galaxy. In fact, it seems that it would be unusual for a
galaxy not to have a massive black hole in its core! Recent observations with all the
main space telescopes have shown evidence of a star being torn about by a massive
black hole in the core of a galaxy. Just read about it in this link.
If you remember the material about radio telescopes, you will recall that they expell
material at great distances from their cores. What if something is in the way?
Apparently that has happened. The so-called "Death Star Galaxy" is blasting material
from its core in the direction of a nearby, small galaxy, which is causing all sorts of
disruption to the small galaxy. You can read about it at the Chandra website. There
are three videos that go with this - one of a model, one of the real data, and one
showing how they relate to one another.
One interesting aspect of these likely black holes is that it appears that black holes
tend to be either small (few dozen solar masses) or very large (millions or billions of
283 | P a g e

solar masses). There are very few instances of "in-between" black holes - those that
would have perhaps a few thousand solar masses worth of material. Recently
observations of the cores of globular clusters indicate that perhaps there could be
some "in-between" black holes in them.
Getting back to the very massive black holes, you might remember that the radio
galaxy M87 (Figure 7) is likely to have a huge black hole in its core, and there are
many galaxies with visible disks in their cores, jets, and the things described in the
above model for an active galaxy. So why aren't there more active galaxies out there?
It is likely that when the galaxies were first forming, that they were initially quasars.
At that time, they were sucking in huge amounts of material into the accretion disk
and black hole, so that when we look at them now, we see them as very active, high
energy objects. However, this won't last forever; eventually the party has to shut down
and the mechanism that produces all of this energy will fade away as less and less
material is sucked in. If you don't feed the black hole, you don't get the fireworks
show. Over time a quasar will fade away, possibly going through a Seyfert phase and
then finally ending up as just a boring galaxy. This helps to explain why we only see
quasars at great distances (which corresponds to looking further into the past, when
galaxies were younger). Actually, you don't have to put a huge amount of material
into these things to get all of the observed energy that comes out of them. Maybe only
about 1 solar mass of stuff per year needs to be fed into the black hole to provide the
observed energy output.
Figure 12. Views of
NGC 4261. On the left
is a composite of the
visible light image
(whitish blob) and
radio telescope image
(yellow lobes). On the
right is an up close
view of the core from
the Hubble Space
Telescope. Notice that
this is pretty much a
disk structure located
at the center of the
galaxy, which is in line
with the model for
Active Galaxies
described above.
284 | P a g e

Image credit: Hubble Space Telescope.
Gravitational Lensing
Another thing that is seen occasionally with quasars and other distant galaxies is one
of the effects of Generally Relativity, the distortion of space due to massive objects.
Let's say you have a bunch of galaxies in a cluster. You then have a lot of mass, and
since mass distorts space, you have a bunch of distorted space in the area of the
cluster. Anything that has to travel through a cluster of galaxies, like light from very
distant objects, will have to deal with the spatial distortions that pop up due to the
galaxy cluster. Under some special circumstances, you don't even need to have a
bunch of galaxies, but just a well placed large galaxy in front of another distant object,
like a quasar. This distortion is just like the distortion seen in the starlight that travels
close to the sun. In this case, since the light isn't coming from a star but from a distant
galaxy, you don't have a single point of light but rather a larger blob to screw up. The
effects of this distortion can be really nifty.
Figure 13. The set
up for
gravitational
lensing. The light
from the distant
quasar passes
near the large
galaxy and the
path is bent. An
observer on the
Earth may think
that there are
actually two
quasars since they see light from the same quasar coming from two slightly different
directions.
At times the distortion will cause us to see not one but two images of the same quasar.
Sometimes we see more images of the same quasar; in several cases up to four images
of the same quasar have appeared in various spatial distortion events. There is one
image that shows five images of the same quasar being distorted (you can see images
of that here). When you have something doing these things to the light from a distant
object, then you have what is known as a Gravitational Lens. A gravitational lens
can do more than just make multiple images appear. There are several instances where
a large concentration of galaxies distorts space like a fun-house mirror distorts your
285 | P a g e

view of the world around you. The images of the more distant galaxies aren't actually
reproduced in this case but are smeared out into unusual shapes.
Figure 14. The Einstein Cross. The four bright
sources around the larger light source are
actually four images of the same quasar that has
been split up into four different images. The
large light source is the core of the galaxy that
is distorting the light from the distant quasar.
Image from Bill Keel's slide set.
Very distant objects are more likely to have
their light screwed up, so we can look at some
of these distant objects more closely to see just
how far away they are. This sort of cuts down
on the "searching for a needle in a haystack" for distant objects, since we know they
won't be as likely to be distorted if they are nearby. This also is one of the ways we
know that quasars are very far away - their light would not likely be distorted if it
didn't have to travel so far (there was a time when quite a few astronomers thought
that quasars can't really be so distant, but gravitational lensing kind of shot their
arguments down). Actually, some of the most distant objects ever discovered were
found through the study of a gravitational lens system. In one case, the redshift (z) for
the distorted galaxy was measured at 4.92, which corresponds to a velocity of about
94% the speed of light.
286 | P a g e


Figure 15. A cluster of galaxies distorting the light from even more distant galaxies.
There are all sorts of streaky images of galaxies shown here around the large cluster.
This is one of the best examples of gravitational lensing, especially since it is
happening on such a large scale. Credits: NASA, A. Fruchter and the ERO Team
(STScI, ST-ECF).
Gravitational lensing is easily visible in large concentrations of mass like a giant
cluster of galaxies. However it is also possible to see a similar effect on a much
smaller scale. Astronomers look for indications of dark matter in our own galaxy by
trying to see the influence that unseen but massive objects have on the light from
more distant stars behind them. The unseen objects will cause the light from distant
stars to become brighter for a short period of time in a well defined manner. However
to see just one of these "lensing events" astronomers have to monitor literally millions
of stars to see if any of them do get brighter. This is further complicated by the fact
that there are stars that normally do change their brightnesses (remember Cepheids
and RR Lyrae?). So those normal variable stars have to be taken into account when
people are searching for the small scale gravitational lensing effects of dark matter.
Several of the current searches for dark matter, particularly the MACHO searches, use
this method to find the unseen objects.
287 | P a g e

In 2007, a report from the Hubble Space Telescope indicated that gravitational lensing
was used to infer the existence of a large distribution of dark matter in a distant galaxy
cluster. You can read about the discovery by following this link. In this case, the
distortions appear to happen at a pretty good distance away from the actual location of
the cluster, indicating that dark matter can extend quite a ways out from the galaxy or
galaxy cluster that it is located in. Remember, we can't actually see dark matter, but
we can certainly detect it by its influence on matter and light. Figure 16 shows the
galaxy cluster and the evidence for the dark matter.
288 | P a g e


Figure 16. The effects of dark matter on an image of a galaxy cluster. The blue
smears are actually images of more distant galaxies, whose light is smeared by the
present of dark matter. If you click on the image, you'll see an animation that shows
the distribution of dark matter that the astronomers were able to derive based upon
the distortion of galaxies. Credits: NASA, ESA, M. J. Lee, and H. Ford (Johns
Hopkins University).
289 | P a g e

Most Distant Objects
What is the most distant object? If there is one thing that can be said about astronomy,
records don't last long. On March 1, 2004, a group of European astronomers using the
VLT discovered an object (named Abell 1835 IR1916, or "Abby" to its friends) with a
redshift estimated to be around 10. This object's light comes from a time when the age
of the Universe was only about 480 million years old, only about 3% its current age.
Unfortunately, there is a problem with this data, as other astronomers were not able to
verify the discovery. So they don't get the record. Oddly enough, only 8 days later, the
folks at Hubble Space Telescope released an image known as the Hubble Ultra Deep
Field (HUDF), which was obtained by pointing the telescope at one part of the sky for
an equivalent of 11 days. The image shown below shows a range of galaxies, some of
which may be from the earliest moments in the Universe. The Hubble folks think that
some of the little specs that you are seeing below are from objects with redshifts
possibly as high as 12! If this is the case, these would be amongst the first objects
formed in the Universe, since they would be originating from a time when the
Universe was only 370 million years old. Again, it should be pointed out that neither
of these actually has been confirmed, so they don't get to go into the record books just
yet. For your information, the most distant galaxy that has been confirmed to be
distant (based on its redshift) so far is UDFy-38135539 with a redshift of 8.55. The
discovery of this object was initially announced in 2009 and that was followed by
measurements of its spectra, which were announced in 2010. The image is courtesy of
NASA, ESA, G. Illingworth (UCO/Lick Observatory and University of California,
Santa Cruz) and the HUDF09 Team. The observations of this object are pretty clear
cut and it is the current record holder. This means that the light we see today left it
when the Universe was only about 4% of its current age - only 600 million years after
the Big Bang (or 13.1 billion years ago). Of course it is likely that someone will break
the record sometime in the future. It is rather interesting that the most distant known
object is a galaxy - and not a quasar. The most distant quasar currently known is
ULAS J1120+0641, which has a redshift of only 7.085. Clearly not as distant as
UDFy-blah-blah-blah, but still it is way out there!
I should mention that in April 2009 a gamma-ray burst (named GRB 090429B)
occurred which was estimated to originate from an object with a redshift greater than
9, perhaps even 9.4, which would make it the most distant object. Unfortunately
gamma-ray bursts don't last very long, so it is very difficult to measure it later with
other telescopes to find the source and accurately measure the redshift of the source.
You can learn more about the discovery here.
So is GRB 090429B the most distant object? Maybe. There is still a chance that its
redshift could be lower than the value announced in the press, since several
assumptions were used to determine the redshift. Remember, just because one
290 | P a g e

scientist observed something and says it is one thing doesn't make it true - all work
needs to be checked and verified before it is accepted. In fact there is another galaxy
which may have a redshift as large as 10.3, but since that has not be confirmed
through other means, it doesn't get to be included in my notes (yet).
So the record for the most distant object can change in the future, and with bigger and
bigger telescopes, it most likely will change. If you want to look in the direction of the
most distant galaxy known (UDFy-etc), you would have to look during the winter
towards the southern horizon, down and to the right of Orion, in the direction of the
faint constellation Fornax. The most distant gamma-ray burster came from a source in
Canes Venatici (below the handle of the Big Dipper).
291 | P a g e



Figure 17. The Hubble Ultra-Deep Field. This image, which is less than 1/10 the
diameter of a Full Moon, contains about 10,000 galaxies. Some of them are thought
to have redshifts greater than 10. If you click on the image, you'll get a view at a
larger version. In this larger view you should some small red objects which are very
distant faint galaxies with redshifts near 6. Image from NASA, ESA, S. Beckwith
(STScI) and the HUDF Team.
Images such as the HUDF and all of the redshift surveys help astronomers understand
the evolution of not only galaxies but also the Universe. They are able to see how
292 | P a g e

galaxies at great distances have some rather interesting characteristics; amongst these
is the existence of massive black holes early on (which supports the quasar model
discussed above). Not only do we see this with visible light telescopes, but also x-ray
telescopes such asChandra reveal galaxy evolution and black hole growth. So I guess
we should move onto the next step - discussing the Universe.

Now that you've read this section, you should be able to answer these questions....
What makes Hubble Law such an important discovery?
What does the value of the Hubble Constant tell us?
What are the characteristics of the various types of active galaxies - BL Lac,
Seyferts, Radio Galaxies?
What are the extreme characteristics of quasars and what basic assumption does
this information rely upon?
What other observations tell astronomers that quasars are very distant objects?
Why is a relativistic redshift formula needed when dealing with quasars?
What is the proposed model that explains the observed characteristics of the
different types of active galaxies and how is it able to explain all of those
different possible active galaxy types?
What is gravitational lensing?
At what sort of distances are the most distant objects?
The Universe

What's covered here:
What are the basic characteristics of the Universe?
How do we define or measure space?
How old is the Universe?

Some of the most basic questions that people have been pondering since people
started pondering things are: What is the Universe like? How old is the Universe?
How was the Universe created? How will the Universe end - if it does end? Where are
we in the Universe? Why is there never anything good to watch on tv? Maybe not that
last question, but the rest of those questions are sort of basic philosophical questions
that have been thought about over time. Now I think we need to start tackling them.
First we'll try to figure out what the Universe is like. The study of the nature of the
Universe is known as Cosmology.
293 | P a g e

Isaac Newton
How big is the Universe? That's a pretty straightforward
question. What's the answer? The first person to tackle this
question with some real scientific gusto was Isaac Newton.
Of course, he had an advantage over the people who
thought about it before he did, since he had the Law of
Gravity to help him in his quest for an answer. First of all,
what if the Universe had a given size, so that it extended
only out to a certain distance - that is to say, what if the
Universe was finite? Newton thought that this can't be the
case. Why? When he looked up in the sky beyond the
Moon, the Sun and the Planets, he did not detect any motions of objects. He thought
that the Universe was static. This doesn't mean that it had lousy reception, but that the
Universe did not appear to move on large scales. What does that tell you? Newton
knew that the Universe had stuff in it - stars, fuzzy things, etc., and he knew that these
things had mass. All of the stuff in the Universe was attracted to all of the other stuff
in the Universe (just basic gravity at work here). If that is the case, then the stuff on
the edge of a finite Universe would be attracted by all of the other mass in the
Universe and would be pulled toward the center of the Universe. If that happens, the
stuff on the edge of the Universe would be moving toward the middle. That sort of
motion wasn't seen by Newton or anyone else during his time - the saw no large scale
motions. The stuff at the edge (if there is an edge) isn't moving inwards. What could
prevent that? - probably more stuff further out. What is preventing the other material
from being pulled inwards? There'll have to be even more stuff out there to prevent
that stuff from being pulled in. What about the stuff further out; won't it be pulled in?
- not if there is even more stuff. I think you can see where this is going. If there were
an edge to the Universe, the stuff at the edge would not sit still, and since Newton
thought that stuff was sitting still (since he saw no motion), that meant there was no
edge.
Figure 1. Newton's argument
against a finite Universe. If the
Universe started out having a
certain size, then the gravitational
pull of stuff near the center of the
Universe would cause the stuff near
the edge of the Universe to be pulled
inward. Eventually, stuff on the edge
of the Universe would be moving
inward to the center and there
294 | P a g e

would be obvious large scale motion toward the center.
What do I mean no edge to the Universe? - just what I said. Newton thought that the
Universe was infinite in size. This was the only way that he could explain the lack of
motion of objects on the largest scales. He of course didn't know about all of the
motions we see today, since his technology wasn't able to detect it. Newton wasn't
really happy with the idea of the Universe being infinite. If that were the case, then
that would mean there must be an infinite amount of mass as well. If there was an
infinite amount of mass, what does that say about the gravity? - you got it - infinite!
Actually, saying the Universe is infinite doesn't really make sense when you look at
the gravity of the whole Universe, or the Universe as Newton saw it, but that was the
best he could make of it.
Figure 2. Why having an infinite Universe may be
a bad thing. When astronomers look through their
telescopes at the night sky, they see galaxies
scattered around and a lot of dark space (view on
the left). If the Universe was infinite, then there
would be an infinite amount of starlight and the
telescopic view would be what is shown on the
right - a blazing white sky.
While an infinite Universe may seem reasonable, there are problems with this. One of
the problems that comes from having an infinite Universe is known as Olbers's
Paradox. This one is a bit confusing, so bear with me. If the Universe is infinitely
large, infinitely old, and filled with stuff in every direction (so that gravity is in a
perfect balance and nothing moves), then when you look in any direction in the sky
you should see something. You may have to look a great distance, but eventually you
will see something. You should see a galaxy, a star or something in all possible
directions in the sky. When you go outside at night, you don't see something in every
single direction of the sky; there are a lot of places where nothing is seen. Another
way of stating Olbers's Paradox is to ask the question, "If the Universe is infinite, why
is the night sky dark?" An infinite Universe would have an infinite number of objects
in it, which are giving off various types of light depending on their temperatures, so
there should be an amount of light in the Universe. If this is case, there would be
constant brightness everywhere and at all times - day or night.
That's a fine pickle. How do we get out of this predicament? Answer: the Universe
could be infinite in size, but it is not infinitely old. It is not infinitely old, so the stuff
that is in certain directions isn't visible to us yet, because it takes time for light to
travel from those distances to us. The light from the most distant objects hasn't
reached us yet. We will only be able to see these distant objects if we wait long
295 | P a g e

enough for their light to get to us. This is the problem with having light travel at a
certain speed - you can't see everything right away. As you'll see later, the Universe is
expanding, and this affects the light from distant objects - stretches it out into
wavelengths that we can't see with our eyes. Light can leave a distant galaxy at visible
wavelengths, but the light will get so stretched out into longer wavelengths by the
expanding Universe that our eyes won't see it - another reason why the sky isn't full of
bright lights at night.
So is the Universe infinite in size? It could very well be, however we'll never be able
to see it as such. Why? It has to do with that fun little concept from Special
Relativity, the speed of light is constant. We can only see parts of the Universe whose
light has gotten here. In other words, if the Universe is 5 billion years old, we could
see out to a distance of, at most, 5 billion lightyears. If it is 10 billion years old, we
could only see out to a distance of 10 billion lightyears. We can only see out to a
distance from which light has gotten to us. Actually the distances I just used are a bit
incorrect, since I didn't account for the expansion of the Universe, but even that would
not help us to see areas outside of the observable Universe. To see an entire, infinite
Universe, you'd have to wait until the Universe were infinitely old.
It seems that the idea of having an infinite Universe is okay, so long as the Universe
isn't infinitely old. What else is known about the Universe? One thing we can look at
is how the Universe looks to us - what are some general things about it that we can
describe? First of all, when you look out over great distances you see stuff everywhere
on large scales. This distribution of material is known as Homogeneity. Now we
know that galaxies aren't uniformly spread around and that clusters aren't uniformly
spread around, but when you look at the largest scales, distances in the billions of
parsecs or so, then things look pretty smoothed out. That's sort of like describing the
ocean. When you are on a boat in the ocean, you may notice how non-uniform it is,
that there are waves that are high or low. If you look at the ocean from a great
distance, like in the space shuttle, it looks pretty smooth. On large scales (or the
largest scales) the Universe is pretty homogeneous (uniform).
Another feature of the Universe is known as Isotropy. This concept says that the
Universe looks the same no matter where you are looking and no matter where you
are located in the Universe. This sort of prevents there from being a center to the
Universe, since everyone sees (again on the largest scales) the same stuff. There is no
preferred location; all the seats are good.
When the concepts of homogeneity and isotropy are combined together, they make up
the Cosmological Principle. This is sort of a rule that needs to be followed when
people think about making models of the Universe. Another way of looking at the
cosmological principle is to say that the Universe is boring! Think about it - both parts
296 | P a g e

of the C.P. refer to how uniform, same, non-unique the Universe is - dullsville, man;
bland as bland can be.
The Cosmological Principle helps us describe the Universe, but there is a much more
important rule that we use when we try to figure out how the Universe works. That is
the concept of Universality, which basically says the laws of physics are the same
everywhere, which means that if gravity behaves a certain way here, it should behave
the same way elsewhere in the Universe. Atoms should behave the same way, light
should be the same, and all the rules should be the same no matter where in the
Universe you are. We really need this rule to be true or else it would be impossible to
understand the Universe, if the laws of physics were random or screwed up all over
the place.
More Models of the Universe
We've seen what Newton thought of the Universe, but what did modern astronomers
think of it? By modern I mean astronomers working over the past 100 years. In the
early part of the 1900s the technology available to astronomers was just getting
sophisticated enough to cause some major disruptions in the theories of the Universe.
One set of observations was of galaxy spectra by a fellow named Vesto
Slipher (1875-1969). He obtained many spectra of spiral nebulae (remember, at this
time people still didn't know what the spiral nebulae were). He noticed that practically
all of the spiral nebulae showed redshifts - that they were moving away from us.
Slipher's observations led to Edwin Hubble's discovery that the Universe was
expanding (in 1929), as is illustrated in the Hubble Law. One aspect of this expansion
is that the more distant galaxies are moving away from us the fastest. What does this
tell us about the expansion? The type of expansion observed by Hubble is a Uniform
Expansion. So what? Is this really all that important?
First, remember that there is no preferred location in the Universe - no center, and
everyone will see the same uniform expansion regardless of their location (remember
the Cosmological Principle). Let's see what happens when we expand the Universe
uniformly. Each part expands in the same way. To illustrate this simply, let's pretend
that the Universe is a straight line with a bunch of galaxies along the line. If the
Universe were to expand uniformly, then we could have the space between each
galaxy double - all would follow the same expansion rate. What do we see?
297 | P a g e


Figure 3. The uniform expansion of the Universe. When the galaxies start out, they
are equally spaced apart. After a time, the Universe expands uniformly; in this case,
the distance between each galaxy doubles. We see the nearby galaxies moving only a
little way (slower velocity), while the more distant galaxies travel a greater distance
(higher velocity).
We see, after the expansion, that the galaxies close to us didn't go very far, so they
didn't move very fast. Galaxies that are much further away went a very great distance
in the same amount of time. To do this, they must be moving very fast. This is exactly
what the Hubble Law shows. It doesn't matter if you are located on galaxy A, B or X,
you would see the same thing. This is in part due to the tendency to not imagine
ourselves moving, but perceive everyone else as moving. This also means that there
is no center to the Universe; the Universe isn't expanding away from just us, but from
everyone. It is hard to visualize the expansion using a one or two dimensional model,
since you are seeing that model relative to a background, but without anything to
compare it to there is no fixed center of the expansion. You could also go the entirely
opposite direction and say that everywhere in the Universe is the center of the
Universe, since the Universe is expanding away from every point - ouch, my head
hurts...

Figure 4. The uniform expansion of the Universe seen from other locations. Here an
alien in what we call galaxy D is observing the expansion. In this case, they also see
all other galaxies (including our own) moving away from them. If they didn't see such
a thing, then there would be no isotropy in the Universe, and some parts of the
298 | P a g e

Universe would see different things - that's not possible. No matter where you are in
the Universe, it looks like everything else is moving away from you. Don't take it
personally.
When we say that the Universe is expanding do we really mean that or do we mean to
say that stuff is moving outwards? Actually, here's where things get a little hairy. The
expansion of the Universe is just that - the Universe actually expanding. What exactly
is expanding? Space is expanding. Huh? Space is expanding? Yes, that's what I said.
Space is expanding, and since things like galaxies are located within this space, they
get carried along by the expansion. The expansion of the Universe doesn't just mean
that galaxies are moving away from one another on their own, it means that the space
that those galaxies are located in is stretching out, causing the galaxies to move away
from one another. Galaxies have their own velocities around one another, in clusters
and such, so that motion has to be separated from the expansion velocity, but when
you get to very great distances the random motions of galaxies are so much smaller
than the expansion velocities. Good grief, this is silly, isn't it?
No, it isn't silly, because we have already messed up space with the concepts outlined
in General Relativity. If we can warp space, can't we also stretch it? Sure we can, but
how are we stretching it? In which directions is it getting stretched, how can we see
that and just how does space get screwed up by this stretching? These questions sort
of depend upon the curvature of the Universe. Uh-oh, it looks like we're going to be
looking at some more weird graphs of warped space! Remember, with the concepts of
General Relativity, space can be expanding into another direction, like into a fourth
spatial dimension. Can we see this? No, not directly, since we are three dimensional
creatures. To understand the possible shapes or the curvature of the Universe we live
in, we need to use some lower-dimensional analogies. Instead of describing space as a
three-dimensional object, we'll use a two-dimensional analogy which can be distorted
(at times) into three spatial dimensions. Now you can pretend that you are ants
crawling around in two-dimensional space, where you can perceive only two
directions (you can't see up-down). What do you see when you look around in your
two-dimensional Universe?
Figure 5. A two-dimensional flat
Universe. Creatures in this Universe only
know of two dimensions and things here
would be similar to what you find on a
flat sheet of paper. Triangles have angles
that add up to 180 degrees and parallel
lines would always stay parallel - never
getting any closer or further away. Even though this image has an edge, there would
be no edge in the flat Universe and it could easily expand uniformly.
299 | P a g e

The first type of curvature we'll look at is the simplest - no curvature, or
a flat Universe. Stuff in this type of Universe acts like it would when things are drawn
on a big flat piece of stretchy material. When it is stretched out uniformly, things keep
their relative shapes and dimensions. Things behave like you expect them to - straight
lines will always be straight and triangle corners always add up to 180 degrees. You
might wonder why I mention straight lines and triangles, but just stay tuned because
we're going to go to the strange Universes next.
Figure 6. A spherical (positve curvature) two-
dimensional Universe is shown. It should be
remembered that the sphere itself is not the
Universe, but rather the surface of the sphere is the
Universe. Creatures in this Universe would only
know of the two directions, like in the flat Universe,
and would not be able to see into the sphere. In this
Universe, you could make triangles with angles
greater than 180 degrees - even though they have straight lines that make up their
edges. Also, parallel lines will converge (come together) eventually.
The next type of surface has what is called positive curvature. This is best represented
in two dimensions by the surface of a sphere. Remember, the 2-D creatures in this
Universe don't see up or down, so they don't know that they are on a sphere. This is
sort of what you experience every day in your life when you look around (you live in
Iowa, after all). The Earth isn't really flat, but it looks pretty flat since it is so big.
Now in this Universe things can be really strange. One thing that can happen in this
Universe is that if you have two parallel straight lines, those lines will eventually
converge (come together). This is like how the lines of longitude on a globe all come
together at the poles - these lines are straight, but because they are located on a curved
surface, they come together. Also, you can draw a triangle on this surface made up of
three straight lines, but with angles that add up to greater than 180 degrees! Egad, this
is crazy! Another nifty aspect of this type of Universe is that if you get into a rocket
and blast off, you'll eventually end up at the same place you started from! That makes
it sort of like a trapped system; you can't get out of this Universe.


300 | P a g e

Figure 7. A saddle (negative curvature) shaped Universe. Actually, it looks rather like
an infinitely large Pringles potato chip. In this two-dimensional Universe, the
curvature is such that triangles have angles that add up to less than 180 degrees and
parallel lines diverge (get further apart).
I suppose if there is positive curvature, there must also be negative curvature. This
Universe is rather strange in that it is shaped like a saddle or a Pringles potato chip.
The way things act in this Universe are sort of the opposite as to the positive curvature
case. Straight lines that start out as parallel will diverge as you go further out into
space - they'll get further apart. Also, triangles can be drawn in this Universe that have
angles adding up to less than 180 degrees - strange.
What type of curvature does our Universe have? I'll tell you right now that it is really
hard to determine that, since it is sort of like how the Earth sort of seems flat. You
know that the Earth is a sphere, yet you are so small compared to it that you see it as
being very flat. The Universe is so darn huge that the curvature, if it exists, is so small
that we can't easily tell what type of curvature there is. In the cases of the positive and
negative curvature, you can blow those surfaces up to such huge sizes that the little
ants in them would think that the surfaces are flat or really close to flat.
I don't want to tell you what astronomers think the shape (curvature) of the Universe
is right now, so I'll save it until later, since it is sort of important in our current studies
of the Universe, and I love to build up the suspense.
Before we go into more about the Universe, in particular its origin, I want to mention
briefly what Einstein thought about the Universe. When he did his work on General
Relativity in 1917, he really didn't know what astronomers knew about the Universe,
so he asked a few of them what the Universe was like. They told him that it was
homogeneous and isotropic. They also told him that it was rather static (even though
Slipher had detected all of those spiral nebulae redshifts; this was before the Curtis-
Shapley Debate and when Hubble did all of his stuff, so no one really knew what
galaxies were any ways). Einstein did his calculations, but his formulas always came
out wrong - they showed that the Universe wasn't static, that it is probably moving.
No matter how hard he tried, the darn formulas never came out the way
he thought they should - or at least following what he was told, that the Universe
wasn't moving in any organized way. He, well, he basically, um, cheated. Yes, the
genius was so intent at getting what he thought was the right answer that he sort of
fudged the formula by adding an extra term to it, what is called the Cosmological
Constant. It is usually represented by the Greek letter . The Cosmological Constant
actually represents anti-gravity, just what Einstein needed to counteract the effects of
gravity and keep the Universe static. Later, when he learned that the Universe wasn't
301 | P a g e

static but expanding, he thought that he had made the biggest mistake of his life - but
had he? We'll have to keepEinstein's blunder in mind later on.
It's time to recap what we know about the Universe at this point. First of all, it is
expanding (Hubble's Law shows this). To be precise, we should actually say that
space is expanding and the stuff in it is just along for the ride. It also has some sort of
shape (curvature), though the Universe is too large for us to see what that shape is
easily - while it may look flat, it may actually be something else. We know that
galaxies evolve, since very distant galaxies look quite different from nearby galaxies,
and we also have the concepts of the Cosmological Principle and Universality to keep
in mind. We know that the Universe is expanding, so it must have been a lot smaller
in the past. That makes sense; just hit the rewind button so that the Universe goes
back to a smaller and smaller scale - how small? As small as we can get it; as small as
it was when it first started to expand - back to the time when it was created, back to
the dawn of time, to the event known as the Big Bang!
Age of the Universe
How long ago did the Universe form? How can we even try to determine such a
value? Well, that's the job of a scientist. First of all we have a couple of tools at our
disposal -
Hubble's Law
Speed of light=constant
Radioactive decay
Steller Evolution
Observations of the Universe
These help us to nail down an age for the Universe which we currently put at
around 13.7 billion years. That makes it pretty old, but how did that number come
about?
As mentioned previously, the value of the Hubble Constant (H
o
) is related to the age
of the Universe - and that's just simply a matter of algebra. So if we know the value of
the Hubble Constant, then that's the age of the Universe? Oh if only it were that easy!
Unfortunately, the Hubble Constant isn't. Isn't what? CONSTANT! The value has
changed over time as the rate of the Universe's expansion has changed. Think of what
happens when you throw a ball up into the air. When it leaves your hand it is moving
at its highest speed, but over time it slows down and slows down and slows down. So
the rate of its motion changes due to the effects of gravity. The Universe has to also
obey the effects of gravity, and therefore must also have changed the pace of its
expansion over time. Gravity would have slowed it down in the past, which tells us
302 | P a g e

the Hubble Constant would have changed over time. That's sort of why we have the
little "o" on it - that means the value it has now. So the value of the Hubble Constant
doesn't exactly give you the age of the Universe, but it does give an estimate to its
value.
What else can be used to determine the age of the Universe? We could use the fact
that the speed of light is a constant to look for the most distant galaxy. For example, if
astronomers find a galaxy that is 5 billion light years away, that means that the galaxy
has to be at least 5 billion year old (since the light had to be emitted and had to travel
for 5 billion years to reach us). And if the galaxy is 5 billion years old, then the
Universe has to be at least that old. So if we find the most distant galaxy, then we
should be able to determine a lower limit for the age of the Universe, right? Not quite.
You'd certainly determine how far away the galaxy is from us, but again, this is a
lower limit for the age. It would have taken time for the galaxy to form. Also, how do
you go about finding the most distant galaxy? Wouldn't it be very, very, very, faint
and therefore very difficult to even see? Certainly astronomers try to find objects that
are further away than any other objects, but to just try to randomly search for such
objects is very difficult. Astronomers search for not only the most distant galaxy but
the most distant supernova as well as gamma-ray bursts. These objects can be seen
over great distances, but again they only give us a glimpse of an object that is close to
the furthest distances we can see.
Recently astronomers announced the discovery of a very old star, one that is estimated
to be 13.2 billion years old! This star, HE 1523-0901, was found to have an unusual
composition, one that showed a decrease in the amounts of various radioactive
elements which helped astronomers to determine the age. You'll learn more about the
use of radioactive elements for age determination in the Planets part of this course.
But radioactive decay is one of the most reliable methods of determining how old
something is, even things like stars.
The old star mentioned above isn't alone. Many stars are very old, particularly the
stars found in globular clusters (remember those?). These stars would have been
formed at the same time, but they die at different rates depending upon their mass.
This allows astronomers to measure the age of a globular cluster by looking for the
largest Main Sequence stars in the clusters. This method requires the use of computer
models that predict how stars evolve, and these give fairly consistent results, again
with ages around 12-13 billion for the oldest stars in globular clusters.
Of course the best way to figure out how old something is would be to just look at it
and see. And in a way that's what many cosmologists do. They look at all aspects of
the Universe and try to piece together from the various observations a consistent
model to explain not only why it is the way it is today, but what it was like in the past,
303 | P a g e

and just how far back that past goes. This includes looking at the structure of the
Universe (clusters of clusters of clusters), the composition of the stars and galaxies
(metals form over time), and the current rate of expansion (Hubble's Law).
The age of the Universe is pretty old, but like all things, it had a beginning. That's
what we'll tackle in the next section.

Now that you've read this section, you should be able to answer these questions....
What did Newton take into account when he described the characteristics of the
Universe?
What is Olbers' Paradox and how is it solved?
What basic rules do cosmologists follow when describing the nature of the
Universe?
What is meant by the "curvature" of the Universe?
What are the possible option for the curvature of the Universe and how are they
different from one another?
What does Hubble's Law tell us about the current nature of the Universe?
What was Einstein's "big mistake"?
What methods are used to determine the age of the Universe?
The Origin and the Fate of the Universe

What's covered here:
What is the Big Bang theory about?
What are the various steps that were thought to happen during the earliest times
in the Universe?
What observational evidence exists that supports the Big Bang theory?
What is the likely fate of the Universe?

We now enter the realm of Cosmogony, the study of the origin of the Universe. The
general name for the theory of the Universe's origin is the Big Bang theory. While
most astronomers accept the basic concepts behind the Big Bang model, there are still
some problems with it and some details that need to be worked out. It covers most of
the bases, so most astronomers are pretty happy with it.
304 | P a g e

You'll remember that Einstein didn't propose a model that was anywhere near that of
the Big Bang, which includes aspects like the Universe evolving and changing over
time. Other astronomers of the time also proposed Universe models that were based
on Einstein's General Relativity, but these models did not get much approval since
they had bizarre features (like having no mass in the Universe or having a static
model). Eventually things did get a little bit more lively when two fellows
independently came up with pretty much the same model - the prototype for the Big
Bang. They were Aleksandr Friedmann and Abbe George LeMaitre - most of the
astronomical community didn't really notice their work, which they did in the early
1920s until about 1930. By that time, Friedmann was dead, so LeMaitre became the
one who was discussing the model with other astronomers.

Aleksandr Friedmann (1888-1925)
of Russia (left) and Abbe George
LeMaitre (1894-1966) of Belgium
(right). These two are credited
with developing the basics of the
Big Bang model.

The models that LeMaitre and Friedmann came up with had an expanding Universe.
The general scheme of the model is that the universe started out much smaller and
much hotter. Along the way various things happened that helped produce the Universe
that we see today. The details of the model have changed over the years, as
observations have caused some parts of the theory to be dropped or altered. One of the
foundations of the model is that if the Universe had a really hot temperature in the
past, then today it would still have a measurable temperature, though it would be
much lower. Of course, the big idea of the model is that it says the Universe is
expanding, which is where the whole Hubble Law comes in. I should mention that
none of these people called their theory the "Big Bang theory." That term was coined
by an astronomer named Fred Hoyle, who was actually sort of making fun of the
theory - he never believed that the Big Bang theory was anywhere near correct, in
spite of all the observations that support it.
Enough talking about the Big Bang theory - let's get into it. We'll go through a time
line, hitting the major events that need to be hit and explaining what is going on at
305 | P a g e

each step. I'll also include what the time is since the creation of the Universe and an
estimate of the temperature of the Universe at each point.
The Big Bang, time=0, Temperature = ? Infinite?
Things aren't too good at this point in the history of the Universe, mainly because
there is no history yet. At this first point we are pretty clueless as to what was going
on in the Universe. There are no current theories of physics that can give us specific
information about what the Universe was really like at this point. Why? In part,
because we don't understand the laws of physics that exist under these conditions.
Why is that? The rules of nature are sort of screwed up under these extreme
conditions. In fact, you can't tell the difference between the four normal forces of
nature - Gravity, Electro-magnetism, Strong Nuclear and Weak Nuclear. In the
earliest history of the Universe, all of these forces were pretty much the same - they
weren't distinct. The way that we describe this one merged force is to call
it Supergravity. If you were to take a physics class at this point in the history of the
Universe, it would be very simple - only one rule at work, so the textbook would be
pretty small. Unfortunately, we don't know what is in that textbook since we don't
precisely know what supergravity is or how it works. We're pretty clueless here.
That's why there is no value for the temperature of the Universe. Guessing that it is
infinite is probably not too bad.
You might wonder about the time before time=0 - what was it like before the
Universe formed? That question may have no meaning - if there were no Universe,
there was no time. Huh? There was no time as we would define it. To try to talk about
what the Universe was like before the earliest time, which we can't even precisely
describe, is pretty meaningless. There are ideas that the Universe goes through cycles,
so that perhaps there was a Universe that existed before this one, or perhaps there just
wasn't anything. Let's get back on the time line and tackle a time that we can handle.
Planck Era, time = 10
-43
second, Temperature=10
32
K.
Now we can describe something, since we know the rules of physics at this point.
Actually, the rules are still sort of unusual, since we still don't have the four normal
forces of nature at this point, but we do have two - Gravity and the GUT. Gravity is
the same old falling-apple-hits-your-head type of gravity, while the other, GUT, is the
Grand Unified Theory. This is where the other three forces are all merged together
into one law. If you were to take a physics class during this time in the history of the
Universe, you'd have only two rules in the book - a bit more complicated from the
previous step, but still not too bad - at least we understand these laws. We understand
the rules at this point, so we can write down some numbers to describe the Universe. I
can even tell you that we think the Universe was about 10
-35
meters in size at this
306 | P a g e

point, but that would be sort of meaningless. It is still just very hot and small. At this
point there really isn't much in the Universe except radiation, in part because it is so
hot.
End of GUT, time = 10
-35
seconds, Temperature = 10
27
K.
Now things are getting closer to being like the current Universe, though there are still
some strange things going on. At this point, there are now three forces in existence -
Gravity, Strong Nuclear Force and the Electro-Weak force. This last is a combination
of two forces, which, like the GUT, is sort of strange. However, with the Strong
Nuclear force out there as a distinct force, there can be some action going on in the
Universe that wasn't possible before. One thing that can happen is that the building
blocks for atoms can be produced - things like quarks and leptons will be produced
(quarks make up protons and neutrons; leptons make up electrons and neutrinos).
How was all this matter created? We just go back to Einstein's E=mc
2
. This law says
that a certain amount of matter will create a certain amount of energy, and it also says
that a certain amount of energy will create a certain amount of mass - this is the
situation that we need to look at. The early Universe was incredibly hot and full of
radiation - so full that the radiation (light) could spontaneously produce particles
(matter) and their anti-particles (anti-matter). In case you are wondering, anti-matter is
not just something that you read about in science fiction but is actually real stuff,
though it is sort of an opposite form of the regular matter. One freaky feature of anti-
matter is that if it contacts matter, then both are destroyed and their masses are
converted back to energy. While it might sound like this is just magic, getting matter
out of energy, you should know that this type of event has been observed in high-
energy physics experiments, so we do have experimental support for this part of the
theory.
Figure 1. The way that matter is produced in the early
Universe. Here we have the creation of an electron and
anti-electron (positron) out of energy, in this case a
photon with enough energy to create the mass of both
particles. Of course, matter and anti-matter don't get along, so when they come into
contact with one another they destroy one another, and you get a photon out of that.
Energy is converted to mass and then mass is converted to energy.
If both a particle and an anti-particle are created, and the Universe is today full of
matter, where is all the anti-matter? Conversely, if matter and anti-matter are each
created, then why should there be any matter at all; wouldn't all of it be destroyed by
the anti-matter that was also created? According to this idea, there should be nothing
but radiation in the Universe! I suppose everything that you see must be an illusion
(including you).
307 | P a g e

We get past this problem by, well, basically, um, cheating. There are some rules of
physics that allow what might be considered the impossible to actually happen. It is
possible that during the production of particles there was a slight excess of matter
produced - something like one extra quark for every billion reactions that produced a
quark and antiquark. The Universe had so many of these reactions going on at this
stage that it is possible for there to be enough extra quarks (and extra leptons) to be
produced to provide all the material needed to form all of the mass of the current
Universe. While it seems like we are sort of cheating here - well, yes, we are sort of
cheating. It is one of the basic rules of quantum physics that - stated simply -
sometimes the impossible happens. It actually has a more complicated name than that,
but it is completely within the realm of physics for there to be more matter created
during the early Universe than anti-matter. Of course, if that weren't the case, you
wouldn't exist.
Inflation, time = 10
-35
to 10
-33
seconds, Temperature = 10
27
K.
Now here is a case where the original theory of the Big Bang had to be supplemented
with some additional work. Why? One of the problems with the original theory was
that it could not explain why the Universe looks so flat. In fact, it is so flat looking
today that if you wind the clock back to the beginning, it would have to be really flat
then as well, since the scales of expansion are so huge that it could not have deviated
much from flatness from the beginning. Our models of the Universe say that it doesn't
have to be flat - yet it looks really flat. Another problem is how uniform everything
appears to be in the Universe, that every part of the Universe has a very similar nature
to every other part (remember homogeneity). Yet those parts are very distant and
would not have been in contact even in the earliest days of the Universe (since they
are so far away). This is sort of like having a pizza that is the same temperature at all
parts, even after it is cut up and divided amongst a bunch of people. All parts of the
pizza were in contact in the past, so they all had the same temperature and keep that
same temperature. What if each part of the pizza as we see it now is 20 miles away?
How could all parts have the same temperature now? They're too far apart and can't
get back together in enough time. Someone thought that at this point in the Universe
(when the quarks and leptons were being made) there was a huge amount of energy
around and this energy may have caused the Universe to expand drastically at this
time. By drastically I mean it got about 10
30
-10
40
times larger than it was before the
expansion.
Inflation helps solve the problem of how flat the Universe looks (just think of
suddenly blowing up a gum ball from its normal to a size that is 10
13
billion light-
years and then standing on it - it would look pretty flat, wouldn't it?). Also, this says
that even though the current Universe is really huge, before Inflation it was a lot
308 | P a g e

smaller and the different parts of it would have had a chance to get equal temperatures
and other uniform characteristics.
End of Unified Forces, time=10
-12
s, Temperature = 10
15
K.
We've finally reached the point where the forces of nature are normal - there are four
distinct forces and they stay that way from now on. At this point, some of those
quarks are starting to get together to form neutrons and protons.
Heavy Particle Era, time=10
-7
s, Temperature=10
14
K.
Here we have big time productions of protons and neutrons. Remember, to make them
you need quarks and a Universe that can allow them to come together and make these
things, which is what happens at this time. There were also some of the anti-particles
still around, but they are gradually decreasing in number.
Light Particle Era, time=10
-4
s, Temperature=10
12
K.
As the Universe cools down, the radiation's energy decreases and the only particles
that can be made are those with low mass - so that's what this era is about. The light
weight electrons are being produced here.
Nucleosynthesis Era, time=3 minutes, Temperature = 10
9
K.
Figure 2. Click to see a larger version. This graph shows
how the composition of the Universe changed with time since
the Big Bang - not much time, though. The abundance of
protons (p) slightly decreased as did the neutrons (n) as
some heavier nuclei were formed. This included deuterium
(D, a heavy form of hydrogen), helium (He), lithium (Li) and
beryllium (Be). The scales are logarithmic, so there are
large jumps between each point in the graph in most of these
values.
We're out of the tiny fractions of seconds eras - we've made it to three minutes since
the start of the Universe. At this point, the Universe is like the interior of a star. There
are protons all over the place, and they can combine together, just like in the core of
star, to make helium nuclei (remember the proton-proton chain?). This is a pretty
important stage, since it determines the chemical composition of the Universe. This
step helps explain the amount of helium that we see in the Universe today. It would
take far too much time for only stars to produce all of the helium we see in the
Universe today, so a step in the early Universe where helium is produced in vast
309 | P a g e

quantities is useful. Other things were also made at this point, like deuterium (heavy
hydrogen) and possibly some other elements. At this point the composition of the
Universe ends up being about 75% hydrogen and 25% helium, which is very close to
what we see today. The amount of helium produced by stars is so small that it hasn't
changed this ratio a great deal, even after all of this time.
Recombination (or Decoupling), time= about 400,000 years, Temperature = 3000
K.
Up to this time, radiation dominated the Universe. What does that mean? It means that
whenever matter wanted to do something, like form into atoms or molecules, the
radiation would rip it apart - it was still too hot for atoms to form completely. The
building blocks were there (protons, electrons, neutrons, and helium nuclei), but they
just couldn't get together. By this time the Universe was cool enough for the protons,
electrons and the helium nuclei to get together and form complete atoms - and the
radiation could not break them up. From this point forth, matter was able to overcome
the forces of radiation; it was able to take over domination of the Universe. You know
what that means - if matter (mass) is all important now, then that makes gravity also
all important.
Galaxy Formation - time = about 500 million years, Temperature = 10 K
This would be when the earliest galaxies would have formed. One thing that isn't
explained by the Big Bang theory is the order of the galaxy - cluster formation. What
the heck does that mean? You could make galaxies, clusters of galaxies and
superclusters of galaxies it two basic ways, either via the top-down method or
the bottom-up method. The top-down method says that big structures, like
superclusters, formed first. These later fragmented into smaller parts that became
clusters of galaxies and then broke down into smaller pieces like individual galaxies.
The bottom-up method says that the little pieces (galaxies) formed first and then these
came together to form the clusters, which then came together to form the superclusters
and other large scale structures. Which is it? That sort of depends upon the type of
stuff in the Universe - and since most of the matter in the Universe is in the form
of dark matter, the way that that stuff acts will help us figure it out. For the top-down
method, big things would first form. This would be the case if material was moving
too fast to form into small bits (galaxies) but not so fast that the big structures couldn't
form. Fast moving stuff is often called "hot," so the top-down method is linked with
what is called hot dark matter. Of course, the other method, with the material
moving very slowly, would allow the formation of small structures first, so galaxies
would form first and then the bigger stuff (clusters and superclusters) later. Slow
moving stuff is "cold," so the cold dark matter model is the bottom-up scenario.
310 | P a g e

Which is it? I'm not telling - at least not yet. I'll get back to it later.
500 million years to Today, Temperature = 3 K
Things progressed with the formation of galaxies, clusters and superclusters, but that
formation was likely impacted by several factors. What about the massive black holes
in the centers of galaxies? Where did they come from? Did they help or hinder the
process of galaxy formation? What about all the energy generated by quasars? Would
that influence galaxy formation? What types of galaxies formed? Did they remain in
that form so they look like they currently look?
Observations of the characteristics of the early Universe are very difficult, and even
observations in the not-to-distant past can be rather perplexing. There are a few things
we do find strong evidence for. It appears that massive black holes seemed to have
formed early, and quickly in the early Universe, based upon observations by
the Chandra telescope. These could have then led to the formation of active galaxies
like quasars, Seyferts and Blazars that we see today (of course we see them as they
looked in the past due to their great distance). The expulsion of large amounts of gas
and energy may also influence the evolution of objects, possibly even hinder the
formation of objects (since it is harder for material to come together when it is hot).
Also in the early stages of a galaxy's formation, it would have a great deal of star
formation and would also release a lot of energy (particularly ultraviolet light) into
space. Such a heating event may also influence the environment around galaxies.
We see evidence for many of these things, with galaxies changing color over time as
the stars within them evolve, as they transition through various active galaxy phases,
and as they undergo mergers and collisions. The end result is that we have a Universe
which was initially energy rich, to one that has matter in it (70% hydrogen, 28%
helium and 2% other stuff), and is still changing.
Support for the Big Bang Theory
This Big Bang thing is all fine and dandy, but is any of it correct? Did the Universe
really form this way? Where is the evidence? We have the stuff that Hubble originally
came up with - that we see the Universe expanding. We also have the concept that
comes from Olbers's Paradox, that the Universe can't be infinitely old; it had to have
some initial time of origin. There is also the way that we perceive the Universe - that
the concepts of the Cosmological Principle and Universality should be correct (or else
we're in big trouble). There is also the observed helium and deuterium abundance of
the Universe - how else but through a phase of large scale nucleosynthesis could this
much stuff be created? While this evidence does support the idea behind the Big
Bang, there is still no smoking gun - we always need more evidence, not just
311 | P a g e

circumstantial evidence but incriminating evidence (oh, dear, I've been watching too
much Law & Order again).
When the Big Bang theory was being worked on in the 1940s, astronomers
determined that there should be some radiation left over from the Big Bang. After all,
the Universe was thought to have originated from a very hot, radiation filled space.
That radiation should still be around everywhere in the Universe, but it would be
diminished by the expansion and cooling of the Universe. The temperature of the
Universe was estimated in 1948 to be around 5 K. At that time there wasn't a high
enough level of technology available to search for it, so the estimate was sort of
forgotten. According to Wien's Law, an object with a temperature of about 5 K would
be producing radiation at a wavelength that corresponds to microwave light, and in
1948 there were no microwave detecting telescopes out there.
A strange thing happened in 1965. Two radio astronomers (Arno Penzias and Bob
Wilson) working for Bell Labs (the once nation-wide phone company) wanted to do
some basic radio measurements of the sky using a telescope that was originally
designed for satellite communications. They started using this rather strange looking
radio telescope by first checking how well it worked. They noticed right away that
there was an annoying static noise coming from all directions of the sky. This is not
normal. Radio signals should just come from objects that give off radio light and these
would be only found in certain locations of the sky, not all over the place. They
thought that there was something wrong with the telescope and decided to check out
its internal workings. One thing they found were some pigeons roosting in the
telescope and that these pigeons had produced "a layer of white, sticky, dielectric
substance coating the inside of the antenna." This is just science-talk for pigeon poop.
After cleaning up the pigeon poop and double and triple checking the electronics they
were still amazed to find that the annoying buzz was still there. They knew that it was
coming from space, but they had no idea what was causing it. Further tests on the
radiation told them that it corresponded to an object which had a temperature of about
3.5 K (the exact temperature wasn't known; this was really just an estimate they
made).
Figure 3. Robert Wilson (left) and Arno Penzias
(right) shown in front of the telescope they used to
discover the Cosmic Background Radiation (in
1965). Copyright Lucent Technologies.
Just down the road (literally) at Princeton
University, physicists Robert Dicke and P. J. E.
Peebles were going to try to find the radiation left
over from the Big Bang, but they didn't have the
312 | P a g e

equipment on hand to detect it. They knew that the radiation of the Big Bang should
have been cooled down to a pretty low value, less than 10 K, but they didn't have a
telescope to view it and they were working hard to make one that would detect this
radiation. Eventually a mutual colleague of the Princeton guys (a fellow named Bernie
Burke), who had heard what the Bell Lab guys had done, suggested to the Princeton
guys that they talk to the Bell Lab guys. They did. The basic upshot were two papers
that were published in 1965 (one from the Princeton group and one from the Bell Labs
group) in the Astrophysical Journal, and in 1978 a Nobel Prize for Physics went to the
Bell Lab guys. Why? What the two folks at Bell Labs discovered was the energy left
over from the Big Bang - the radiation that was created from the early, hot Universe
that has been stretched out and cooled down to a temperature of around 3 K today.
The 3 K radiation is called various things, like the Cosmic Background Radiation,
the Cosmic Microwave Background Radiation, the 3 K background radiation or
all other sorts of variations of these words. Let's just call it the CBR (Cosmic
Background Radiation). This radiation is one of the strongest supports for the Big
Bang theory. The radiation is seen in all directions of the sky and has the same
temperature everywhere, so it is an indication about the smoothness of the Universe.
The smoothness of the CBR is, however, a very bad thing! Remember that radiation at
one time dominated the early Universe and under those circumstances matter would
have been "controlled" by radiation. What this really means is that matter would not
have been able to clump together unless the radiation was also clumping together.
When we look at the Universe, we see that matter is not uniformly spread out across
the Universe like the background radiation appears to be. There are all those large
scale structures, clustering of superclusters and so forth. If the radiation were perfectly
uniform in all directions, then the matter should also be perfectly uniform and stuff
would not have come together to form galaxies, stars, people, etc. It looks like we're
in trouble with the Big Bang theory.
Astronomers realized that the radiation appears very smooth when measured from the
surface of the Earth. As you know, the view from Earth is limited due to the fuzziness
and blurring caused by the atmosphere. The best way to measure the radiation would
be to launch a satellite to measure it from space. In 1989 the COBE - Cosmic
Background Explorer - satellite was launched. Unlike observations from the ground,
the CBR could be observed by COBE in all directions and at very high precision.
Within about one month of launch, the COBE satellite measured the temperature of
the CBR and found it to be 2.735 K (I'll still call it 3 K, since that's easier to
remember). Actually, the temperature that it measured was so precise that all graphs
showing the readings at different wavelengths have such small uncertainties (errors in
measurements) that you can't even plot those uncertainties - they are just too small. In
January of 1990 (less than two months after the launch), the results showing the
313 | P a g e

radiation temperature were presented at the American Astronomical Society meeting
in Washington D.C. When the plot showing the radiation was displayed, it received a
standing ovation - those crazy
astronomers.
Figure 4. The data from the
COBE satellite showing that the
Cosmic Background radiation
has a temperature of about 2.73
K (based upon the shape of the
curve and the location of the peak
- Wien's Law) and is pretty much
a black body. The boxes show
where the satellite data are
plotted and the line is the
theoretical line for a black body.
Data from the COBE mission.
One thing that I should point out about that graph - there is a line drawn through the
data points. This is the line a true, perfect black body would have if it has a
temperature of 2.735 K. WOW! Do you get what I just said? The radiation from the
Big Bang is the closest thing out there to a real black body - a perfect radiation
source! It seems sort of appropriate that the only thing in the Universe that is close to
being a perfect black body is the Universe itself. After some additional work it was
shown that the temperature is really about 2.728 (with a 0.004 uncertainty) K. I think
I'll still stick to calling it 3 K.
The radiation is that of a black body, but what about the clumpiness of the Universe
problem? It took some time, but a thorough examination of the COBE results showed
that the CBR doesn't have exactly the same value everywhere. Also, by examining the
data very carefully, they found that the radiation had a temperature variation of around
0.0001 K. This is sort of a small temperature variation, but it is enough to do the job.
The radiation is clumpy enough to lead to a
clumpy Universe.
Figure 5. A map of the entire sky based upon the
COBE data. The different colors correspond to
slightly different temperatures, indicating that the
Cosmic Background Radiation is not of a uniform
temperature but does have some small variations.
Image from the COBE mission.
314 | P a g e

There have been many other CBR experiments that have improved upon the COBE
data. Many of these projects are based in Antarctica, since the atmosphere there is
rather dry and observing conditions are good. Some of the current projects include
BOOMERanG, BEAST, ACBAR, and many more. Even though these experiments
are all unique in how they measure the CBR, they tend to give the same results - that
the radiation is clumpy on very small scales, which indicates that the early Universe
had very small scale structures very early in its history. This is evidence that supports
the cold dark matter formation theory or the bottom-up scheme. Actually, similar
results (supporting cold dark matter) were also obtained by the Hubble Telescope
observation of distant galaxies (2004) as well as other visible light surveys of
galaxies. Most of the CBR experiments mentioned above were rather limited in the
areas of they sky they surveyed. Another recent survey from space was completed by
the WMAP spacecraft (operating 2001-2010), which provided the best data yet.


The BOOMERanG project is shown at left - a mapping of the sky using a balloon from
Antarctica. This is similar to the COBE project, except that the BOOMERanG instrument is
about 35 times more sensitive. A part of the map it produced is shown at right. Click on the
image at right to see a larger version of the map and how it compares to the COBE data.
Clicking on the image at left takes you to the BOOMERanG website. Images courtesy the
BOOMERanG project.
Currently most astronomers who are trying to figure out what the Universe is like
have been using the cold dark matter theory to model the growth of structures in the
Universe. This has been done by various groups, including some that have made some
rather nifty images. You can check out the views of the early Universe according to a
group from Princeton (movie by Paul Bode and collaborators), which shows the early
clumping of structures. This view shows a region of space 100 million light years in
size - pretty big! Another animation from a group at Los Alamos, Cal Tech, and
Mount Stromlo shows the quick clumping of material early in the history of the
Universe. In both cases, material clumps together rather quickly, and the clumps grow
in size as they pull in more and more stuff. These views of the Universe are being
constantly adjusted due to discoveries and clearer results from all of the CBR
mapping projects. For the most part there is good agreement between our theories and
315 | P a g e

our observations on the origin of the Universe. Now let's get to the other end of the
Universe - its end!
Fate of the Universe
Now that we think we know how the Universe started, how do we think it will end?
Will it end? Is there an end? Traditionally, we believed that the fate of the Universe
depended only on one force. That force is... (all together now) Gravity. Why? Gravity
is the only thing that we knew of that could stop the expansion of the Universe and
possibly bring it all back together. It is the only force in nature with enough power to
do such a feat. Remember, there are no limits over how far gravity extends. You are
pulling on and being pulled on by all the objects in the Universe that have mass. Even
though you don't feel the pull (and neither do they), the effect is still there. When the
influence of all of the matter of the Universe is accounted for, we need to know if it
just might be enough to overcome the expansion of the Universe.
If it all depends upon gravity, what does gravity depend on? If you just look at the
formula for gravity, there are two values that can be changed, and they are mass and
distance. As you know, astronomers have a really difficult time measuring distances
and our measurements of masses are also sort of dubious. To determine individual
values for mass and distance is quite tricky. Actually, we don't do that - we sort of
combine mass and distance together into one value - density. All we have to do is
figure out the average density of the Universe. If it is dense enough, then there is
enough material in a small enough area to force the Universe to stop its expansion and
collapse down again. If it isn't dense enough, then the expansion won't stop.
What is "dense enough"? We define the critical density as the density needed to stop
the expansion of the Universe. To complicate matters, the value of the critical density
depends on the value of the Hubble constant, H
o
, and the shape (curvature) of the
Universe. Even though we don't know the value of the Hubble Constant precisely, we
do have a rather good idea of what the critical density is. For current estimates of the
Hubble Constant, the critical density is close to 8 x 10
-27
kg/m
3
, or about 5 protons per
cubic meter of space. This isn't very dense compared to stuff that you are familiar
with, and actually it is an incredibly low density. What is the actual density of the
Universe? How does it compare to the critical density? Unfortunately, we don't know
(yes, that is a pretty lame answer, but there is a reason). When we do measure the
density of the Universe, we get a value for the density that is nowhere close to the
critical density. We always get very low values, sometimes up to 100 times less than
the critical density. I guess that means there isn't enough matter in the Universe to
stop the expansion. No, it just means we get a density that is pretty low. You must
remember, the density that we get is based upon our accounting for all of the stuff we
can see. What about all of that stuff that we can't see? What about all of that dark
316 | P a g e

matter? What affect would that have on our value of the measured density of the
Universe? If there is a huge amount of dark matter then the value for the density that
we get would be much greater, perhaps even greater than the critical density. We don't
know precisely how much dark matter there is out there, so we don't know what we
are missing. I guess we can't use density of the Universe to determine the fate of the
Universe.
Even though we don't know the density of the Universe, we can still see what the
results of its value would be for the fate of the Universe. If the density of the Universe
is less than the critical density, then the Universe does not have enough mass (gravity)
to stop the expansion and it will continue to expand forever. In this case the Universe
is OPEN. If the Universe were to expand forever, then clusters of galaxies would be
getting further apart from one another, and eventually all the matter in galaxies that
could go into making stars would be used up. Finally all the stars would die and the
Universe would be a very dark, cold place - large, but dark and cold. If the density of
the Universe is greater than the critical density, eventually the force of gravity will
slow down the expansion and cause it to stop. In this case the Universe is CLOSED.
In the case of a closed Universe, not only would the expansion be stopped but also
eventually reversed. The Universe will start heading toward a collapse after the
expansion is stopped. This is sort of how a ball that you throw in the air goes first
away from the Earth (expansion of the Universe) and then returns to the Earth
(collapse of the Universe). Gravity will bring everything back together again.
Galaxies would start getting closer, and they would all head for a Big Crunch. It is
also possible that a closed Universe could experience another Big Bang after the Big
Crunch. This is due to the way that material would have been destroyed as it was
crunched - it would have been turned to radiation, and there would be a lot of that. A
lot of radiation is what you would have when you start out with a Big Bang. Of
course, it isn't guaranteed that there would be another Big Bang following the Big
Crunch, but you never know.
There is also the very unlikely possibility that the density of the Universe equals the
critical density, in which case the Universe is FLAT. It has to be exactly, precisely,
completely EQUAL to the critical density, no more or less. In that case, the Universe
will keep expanding but slow down as it does so. In a way this is very similar to the
Open Universe case, except here the Universe will keep slowing down, slowing down,
and slowing down, until, far into the future, the expansion is so slow that it isn't
observable. The odds of the density of the Universe being exactly, precisely equal to
the critical density made many astronomers think that this could not be the case.
However, our ideas about this have been changing.
The different, traditional scenarios for the fate of the Universe, the shapes and the
densities are presented in a table here. You may wonder why I say "traditional" - or
317 | P a g e

maybe you don't. Either way, our "traditional" view of the future of the Universe has
recently been thrown out the window - just keep reading.
A Surprising Turn in the Universe's Story
In 1997, the stuff that I have been describing would have been about as far as I needed
to go when talking about the Universe. This changed in 1998 when some very strange
and unexpected observations were made. Several groups had been observing very
distant galaxies containing supernovae to see what the expansion of the Universe was
like long ago. Remember, when you look further out in space, you are looking at
objects that used to be at those locations and looking at how the Universe was
behaving long ago. If the expansion of the Universe was slowing down due to the
force of gravity, then the distances of these supernovae would be different than what
you would get if the expansion never changed speed - by just using the Hubble Law
and a current Hubble Constant. If the rate of the expansion changes over time, then
the Hubble Constant will change over time. Actually, calling it a "constant" isn't really
a good idea, since it was different in the past and will probably be different in the
future. The astronomers making the observations of the distant supernovae expected
to find that the Universe is slowing down its expansion due to gravity, since we
thought that the Universe was either flat or possibly even closed. And as you know by
now, gravity is the all important force, and we (astronomers) thought it was the only
thing that could control the fate of the Universe.
What did the supernovae observers see? They determined that the expansion of the
Universe was not slowing down but instead was speeding up. The Universe is
ACCELERATING!?!?!?! This is such a drastic result that a lot of people have a
hard time believing it. But observations by even more astronomers have done nothing
more than support this conclusion, and they can't all be wrong, can they? It is
unlikely, since they are using different methods and measuring different things (not
just supernovae). Why is an accelerating Universe such a bad thing? Go back to
Newton's Laws of Motion. You have F=ma. To accelerate something (a), you need to
exert a force on it (F). Something is pushing on the Universe? Not really, but
something has to be not only overcoming the slowing down effects of gravity but
exceeding them. Dang, this is weird stuff. Also, by having a force fighting the effects
of gravity and overcoming them, worrying about the amount of dark matter in the
Universe is kind of not so important - the density of the Universe doesn't seem to
make much of a difference to its fate. The density and dark matter are still important
in helping us determine the structure of the Universe and how that came about, but it
kind of ruins all that stuff I was saying about Closed and Open Universes. It's time to
turn to the mysterious force that is overcoming gravity.
318 | P a g e

What is going on with the Universe? What can overcome gravity? Believe it or not,
something that can be best described as anti-gravity has been proposed. The way that
this anti-gravity is included in Einstein's General Relativity formulas is by adding in
that term that Einstein stuck in his formulas when he thought that there was no motion
to the Universe - the Cosmological Constant, . Einstein thought he screwed up
when he stuck this term in his formulas. Now it seems he might have been correct to
include it - but not for the reasons he thought it was needed.
That's how we deal with the anti-gravity in formulas, but what is this anti-gravity
thing really? A concept known as dark energy has been proposed as the form that the
anti-gravity force takes, energy that exists in a "vacuum". If you remember back to the
Big Bang discussion, matter and anti-matter can be made from energy (radiation). If
this is the case, then this "vacuum" would have some energy, which can be translated
into a pressure, which can help to push the expansion of the Universe. I know that this
sounds like magic and other silliness, but various projects have shown that this may
actually be the case. This dark energy is not too unusual since it is sort of like the
energy that went into driving the expansion of the Universe during the Inflation stage.
Recent work by the Hubble Space Telescope shows indications of the influence of this
repulsive force as far back as 9 billion years ago! Observations by the ultraviolet
telescope GALEX also confirm the presence of dark energy. Observations by
the Chandra telescope indicate that dark energy may have even influenced the
formation of early galaxies (in a negative way). So it now appears that we have a
Universe that is made up mainly of dark matter (which works to pull it together),
being overwhelmed by dark energy (which works to expand it). A cosmic tug-of-
war is going on.
People often confuse dark energy and dark matter, which isn't unusual given their
similar names. Here's a link to a radio interview with Neil deGrasse Tyson who
explains them in terms that should help you keep them straight in your head.
Here's another interesting thing you should ponder, E=mc
2
. I'm sure you remember
that interesting little formula, but what does it have to do with the Universe? This
formula shows the relationship between energy andmatter. So if you want to take an
inventory of the stuff in the Universe, you have to include not only regular stuff
(matter), but also dark matter, energy and dark energy! So how much of that stuff is
out there? I'll get to that in a second.
Another result that has come about which is equally unusual is that the most likely
curvature (shape) for the Universe based upon these and other observations is that it is
perfectly Flat. This result is seen when comparing the BOOMERanG, WMAP and
distant supernovae results (and many other studies) to see what is the most likely
319 | P a g e

shape for the Universe, along with determining what the expansion of the Universe is
doing. Ultimately this means that the Universe isn't just Open, but WIDE OPEN! If
this dark energy keeps driving the expansion of the Universe, then there will be no
closing the Universe, no collapse. The Universe will just get bigger and bigger at a
faster and faster rate. How's that for mind blowing?
So does this mean that the Universe is really never ending? Some have proposed that
if the acceleration rate of the expansion increases, it will be harder to keep things
together, since gravity will have to fight the ever accelerating expansion "force". First
superclusters of galaxies would break apart. Then clusters of galaxies would break up.
Eventually it would be difficult for galaxies to hold themselves together. Star clusters
and then solar systems would be torn apart. Eventually it would difficult for individual
objects like stars and planets to stay together. Ultimately all matter would be broken
apart, since even the forces that hold atoms together would fail. Models that follow
this scenario predict an end to the Universe (as we know it) in about 20 billion years.
This theory is known as the Big Rip. Will it happen? Not likely. While it may seem
interesting to have things ripped apart down to the smallest levels, it doesn't appear
that a Big Rip will happen. At least we don't think so at this time (but remember, our
ideas about what the Universe was doing have changed not so long ago, so stay tuned
for new findings that may change all of this - again).
Recently accurate measurements of the Universe have come from the WMAP mission
that was launched in June 2001 and finished operations in 2010. It was designed, like
COBE, to view the entire sky, but with a much greater precision than COBE. In
February 2003, many results from WMAP were released and they provide us with the
best information to date about the Universe. As with any mission, the WMAP results
have been possibly overwritten by the results from the Planck mission, which has a
higher resolution map of the cosmic microwave background radiation. This allows
astronomers to refine the models of the Universe and check to see the best fitting
values. Of course over time the results from WMAP and Planck will no doubt be
revised and improved upon with even better satellites. Just so you can see the results,
the table below shows the values from both WMAP and Planck. While these are the
best measurements we have available today, they are of course not exactly 100%
precise, though they seem to be pretty well accepted by the astronomical community.
-
Value WMAP Planck
Age of Universe 13.75 billion years 13.819 billion years
Hubble Constant 71.0 67.11
% Dark Energy 73.4% 68.25%
% Dark Matter 22.2% 26.71%
320 | P a g e

% Normal Matter 4.49% 4.90%
For either of these missions there are some rather consistent results -
The density of the Universe is pretty much equal to the critical density,
meaning that the Universe is Flat. The error on this value is less than 1%,
which isn't much.
Most of the Universe is made up of the Dark Energy (remember, energy=mass)
Only about a quarter of the Universe is made up of matter, though most of that
is dark matter. Only about 4-5% of the entire Universe is made up of
"normal" stuff, and most of that is hydrogen.
The Hubble constant is about 70 km/s/Mpc. This is pretty close to the value
that people using the Hubble Space Telescope got.
The Age of the Universe is 13.7-13.8 billion years. So basically if you said the
Universe is about 14 billion years old, you won't be too far off the mark.
Era of Decoupling happened about 375,000 years after the Big Bang.
The Universe's expansion is accelerating.
The fate of the Universe could be just continued expansion or the Big Rip -
here's a table with these new possible fates.

Figure 6. Map of the sky from the WMAP satellite. The different colors indicate the
variations in the CMR. This telescope has a better resolution than COBE (Figure 5)
and was able to see even small scale clumpiness in the radiation. Image courtesy
NASA/WMAP Science Team.
321 | P a g e

That pretty much answers it, right? Of course it doesn't. There is still a lot of stuff to
be worked out, mainly details and finding out more about what happened in various
parts of the Universe's past. There is quite a bit more to do yet and I'm sure there will
be more newspaper headlines about stuff like this in the future. I'm sure some of you
don't believe any of the stuff I've said about the creation of the Universe, or any of the
results that have been observed about the origin and fate of the Universe, but that's
fine; you're entitled to your opinion. Of course, this is the stuff that will be on the test.

Figure 7. Map of the sky from the Planck satellite. Like the previous image, the colors
indicate the slight temperature variations in the CMR. This telescope has a better
resolution than WMAP (Figure 6) and has probably the most precise measurements to
date. see even small scale clumpiness in the radiation. Image courtesy ESA and the
Planck Collaboration.


Now that you've read this section, you should be able to answer these questions....
How do the characteristics of the Universe change over time?
Why is our knowledge about the very beginning of the Universe limited?
How did the buildling blocks of matter form in the early Universe?
How did forces in the early Universe change over time?
322 | P a g e

Why does the Big Bang theory require the early Universe to be hot, and dense
in the first few minutes?
What observational characteristics support various parts of the Big Bang
theory?
What current observations have changed our view of the nature of the
Universe?
What is the Cosmological Constant, and why does it appear to be important
today?
What are the current possible fates for our Universe?
What is most of the Universe made up of?
The Solar System

What's covered here:
What exactly is a planet?
What types of planets are there?
How did the solar system form?
How do astronomers determine the age of solar system objects?
Do other solar systems/planetary systems exist?

click to see larger version - images courtesy Calvin J. Hamilton.

What is in the Solar System? - pretty much everything found
around the Sun, which includes the Sun, planets, satellites
(moons), asteroids, comets and anything else that is the area
(dust and debris is pretty much all that remains). It should be
noted that the proper way to refer to a moon around a planet is
the term satellite. Satellites can be natural (like a moon) or artificial (like weather
satellites). This also helps to avoid confusing our Moon with others.
Pretty much everyone learned about the planets when they were in elementary school
and you learned that the order of the planets going from the Sun outwards is Mercury,
Venus, Earth, Mars, Jupiter, Saturn, Uranus, and Neptune. I suppose you were
wondering where is Pluto on this list? Isn't it a planet? I think the best way to answer
this is that Pluto should never have been considered a planet in the first place. It
doesn't have the physical characteristics of any other planets, it isn't even the largest
thing in its neighborhood, and if we are going to consider calling Pluto a planet, then
323 | P a g e

there would be a bunch of other objects out that that should also be called planets -
which means you would have more to memorize.
In the Summer of 2006, the International Astronomical Union (the largest
organization for astronomers around the world) had a committee come up with the
definition for a planet in an effort to decide if Pluto should be a planet. Since planets
were objects that have been known of since ancient times, no one had ever thought
about defining what a planet actually was until astronomers started discovering a
bunch of other planet like objects in our solar system. As it turns out the committee
came up with pretty lame guidelines, which were voted against, but another group
came up with the criteria for being part of the planet club. These are the following:
1. Must orbit the Sun
2. Have enough mass so that it is a fairly round object
3. Is the dominant object in its orbital path around the Sun (biggest thing around)
Pluto does not meet the third criterion, so it is not a planet. The same holds true for
other objects such as asteroids and comets and a whole bunch of other small objects
out there. Objects like Pluto that are big, but not the only big thing in their area are
now refered to as dwarf planets. Sort of a downgrade, but it does make sense.
Anything that is not a planet or dwarf planet is just a small solar system body.
I don't know why there was this big fuss - I decided well before the International
Astronomical Union that Pluto wasn't a planet. They should have just asked me! So no
matter how you view it, as far as this course is concerned there are only 8 planets in
our solar system!
We'll start the study of the solar system by looking at how it came about and how it
compares to other solar systems out there (yes, there are other solar systems out
there). One of the most obvious aspects of our solar system is how the planets are
divided up into two types.
Terrestrial Planets
Figure 1. The terrestrial
planets shown to scale.
Images from NASA.
These are planets that are
similar to the Earth (that's
what the terra means) and
include Mercury, Venus,
324 | P a g e

Earth (duh!) and Mars. These planets have the following characteristics.
They all have a relatively small mass and radius. Earth is the biggest one
amongst this group. While you might think that the Earth is pretty large, as
you'll see, it is actually pretty puny.
These worlds have warm, solid surfaces. The temperatures range between
values of 100 K to about 900 K. While 100 K is pretty cool for you and me (it's
actually about -280 F), it is pretty warm compared to other places.
These planets have dense, rocky compositions. A good fraction of their
interiors have compositions of nickel and iron, making them have densities that
range from 3000 - 5000 kg/m
3
.
They are all found close to the Sun, since of course these are the first four
planets.
Those planets that have an atmosphere have one that is made up of heavy gases
- things like N
2
, CO
2
, O
2
, and H
2
O. Mercury is the only one without an
atmosphere.
Satellites (natural ones) are very rare - only three satellites are found amongst
the four terrestrial planets.
Jovian Planets
Figure 2. Jovian Planets shown
to scale. Image courtesy Calvin J.
Hamilton.
These are the planets that are
similar to Jupiter (Jove is another
name for Jupiter, by Jove!).
Included in this group are Jupiter,
Saturn, Uranus, and Neptune.
These planets have the following
characteristics (which are in some
ways exact opposites from the characteristics of the terrestrial planets).
These are the big planets; they have very large masses and radii. Masses are
from about 15 to 300 times larger than the Earth.
These planets don't really have surfaces, they just have cloud layer after cloud
layer. They do thicken up, but you can't really stand on these worlds.
The gaseous compositions mean that these guys will have very low densities. In
fact, these are the lowest density planets, with average densities being in the
neighborhood of the density of water (1000 kg/m
3
).
They are all found in the outer solar system.
325 | P a g e

The atmospheres (or upper cloud layers) are composed mainly of light gases -
hydrogen and helium are the most abundant, and there are also small amounts
of hydrogen rich molecules, such as methane, water and ammonia in them,
though they tend to be in very minute amounts.
These worlds also
have large numbers of
satellites going around
them.
Figure 3. The planets in the
solar system shown to scale.
Image courtesy Calvin J.
Hamilton.
The descriptions of the two
types of planets provides another reason to exclude Pluto from the planet club. Its
characteristics don't fit into either category. This only adds more support to the idea
that it shouldn't be a planet. So if you aren't happy with the rather arbitrary criteria put
together by the International Astronomical Union for the characteristics of a planet,
the basic characteristics of the planets in our solar system leave Pluto out in the cold
(literally and figuratively).
Formation of the Solar System
The current configuration and characteristics of the solar system should provide us
with some clues as to how it formed. If you are going to try to understand the
processes that went into the formation of the solar system and are attempting to
theoretically duplicate it, then you better make sure your model duplicates these
characteristics and others that are important aspects of the solar system. It may also be
the case that there were rare events in the formation of the planets in our solar system
that we don't expect to occur in other planetary systems. Of course, there is also the
question of how likely is it that a planetary system forms around a star - so many
questions, so little time!
Here are some of the features of the solar system, which can be thought of as clues to
how the solar system formed.
Most of the material in the solar system is in the Sun (actually, about 99.9 % of
it is). The planets and all the other junk are pretty minute in comparison to the
Sun.
326 | P a g e

All the planets orbit the Sun in about the same orbital plane, the ecliptic
(Mercury deviates from this the most with a 7 degree tilt).
The planets orbit the Sun in the same direction, counter-clockwise as viewed
from above the north pole.
The planets' orbits are very close to being circular (Mercury is the least circular
with an eccentricity of 0.21).
Most planets rotate on their axes in the same direction. The exceptions for this
are Venus and Uranus. Venus is completely tipped over, while Uranus is sort of
on its side.
The planets' satellites tend to follow the motions of the planets; they are
orbiting and rotating in a counter-clockwise direction. There are actually many
exceptions to this rule, but most do follow it fairly well.
The density of the planets decrease with increasing distance from the Sun. This
also indicates that the compositions of the planets change with increasing
distance from the Sun. When you include things like asteroids, their
compositions also follow this trend.
The outer planets are more massive than the inner planets and are made mainly
of hydrogen and helium.
There is extensive cratering of both planets and satellites, and it is thought that
most of this cratering occurred at about the same time.
How is a solar system made? What are the ingredients that we need to make
one? Actually, you could answer these questions by asking how the Sun was
made since it is the main component of the solar system. Remember, it has most of the
mass of the solar system. If someone were to do a survey of our solar system, they'd
most likely say that there is a G2 Main Sequence star in it and not much else. Maybe
they'd also mention Jupiter, but beyond that, everything else is pretty minuscule. If
you are going to make a solar system, you are going to primarily be involved in
making a star (the Sun in this case). We have already gone over this process of how
stars form, but we did not look at how the other stuff - planets, satellites, etc. form
since they are such small components of the whole process.
To make a solar system you need to start out with a gas and dust cloud composed of
about 70% hydrogen, 27% helium, and 3% all else. We're talking about a gas cloud
that will form into the solar system, so we usually refer to it as the solar nebula. Now
you know where all of this stuff has come from, don't you? At least, if you have been
paying attention the past 11 weeks or so, you should know. First of all, hydrogen and
helium are the most abundant elements in the Universe and were formed in the Big
Bang (even though some He is formed in other stars, that doesn't contribute a lot to
the overall abundance). All of the other stuff (the stuff that makes up 3% of the mass)
came from secondary sources, through the fusion of elements in the cores of stars.
327 | P a g e

These stars then spew this material out through various mechanisms such as stellar
winds, planetary nebula phases, and the ultimate littering of the galaxy, supernova
explosions. The most common elements found in this 3% are things like oxygen,
carbon, nitrogen, silicon, iron, and so forth.
You have a solar nebula sitting out there that will start collapsing down to make the
Sun and some other stuff. We don't want it to form in any general way, do we? We
have to have it ending up as a solar system like we see today, with all the planets
spread out around the Sun in the ecliptic. How can we do that? One thing that will
help the process is to have the nebula rotating a little bit. What good is that? This has
to do with that concept ofangular momentum that we have run into before (see the
stuff about pulsars). As something collapses, it spins faster. The gas cloud starts
collapsing, with most of the stuff collapsing into where the Sun will be. Does that
mean there will only be a Sun formed in the middle? No, that's not what's going to
happen. If you have something spinning really fast, there is a tendency for it to stretch
out. Have you ever seen a person make a "real" pizza? I don't mean someone taking a
pizza out of a freezer and putting it into a oven, I mean someone making a pizza the
old fashioned way. One thing they do is toss the dough up into the air. Is that all they
do? No, they also spin it. Why? The rotation (angular momentum) helps to spread out
the material (dough). The collapse speeds up the rotation, which in turn spreads stuff
out into a flat disk (like a pizza or a solar system). This explains why all of the planets
and most of the material in the solar system are found in this disk (the ecliptic) and are
moving around the center in the same direction.
Now that you have all of the stuff in a nice, neat, flat disk, you can start making
planets - well, not right away. Material will first come together in little bits and some
bits will come together more easily than others. Why? The temperature of the disk
will play a role into how quickly things form into bigger chunks of material and if it is
even possible for some things to form. Near the Sun it is very difficult for the
lightweight gases and ices to come together since the strong radiation from the newly
forming Sun will tend to break these things up easily. Only stuff with strong
gravitational attraction (heavy stuff) will be able to come together and stay together.
This would include the heavy elements or what you might think of as the stuff that
makes up rocks and metals. You have to remember, this high density stuff is sort of
rare in the solar nebula compared to the lightweight stuff, so you aren't going to be
making many of these rocky/metallic chunks or very large chunks of rocky/metallic
material. Further from the Sun, the temperature of the cloud is cooler and gases and
ices can coalesce easily - as well as the less common metals and silicates. There are
more of the light weight (low density) elements, so they will far outnumber the
heavier metals and rocks. You can see right away that the densities of the planets vary
according to the heat (distances) from the newly forming Sun. The sizes of the planets
328 | P a g e

are influenced in a similar way; since the inner planets are made up of the less
abundant heavy elements, they will end up having lower masses, while the outer
planets are made up of the most common materials, hydrogen and helium, and
therefore they will be much bigger. Another aspect of this temperature dependency on
the formation of the planets is the size of Jupiter. Jupiter is located closer to the Sun
than the other Jovians and is about as close as a Jovian type planet can get - it is also
the most massive. When Jupiter formed it sort of gathered up most of the available
mass, since it was close to the location where all the abundant lightweight stuff was
able the coalesce. It is closest Jovian to the Sun, so it is closer to where more of the
mass is. The end result is that Jupiter is by far the most massive planet in the solar
system.
Let's get back to making planets. What is driving the whole solar system formation
process? Gravity is the force behind it all (as you should know by now, "Gravity
Rules!"). The material is coming together, forming larger and larger
chunks. Condensation and accretion will cause concentrations to develop in the solar
nebula until the chunks reach sizes where they dominate their areas - they are
now planetismals. These are the basic building blocks for a planet. They vary in sizes
from tiny fractions of a centimeter to several kilometers. They are just basically big
enough (and therefore have enough gravity) to pull in more material so that they can
get bigger. Of course, if you combine enough of these planetismals together you will
make your basic planet.
Figure 4. A planet that comes together
gradually will have a mish-mash
composition. Once it gets heated up, the
layers will sort themselves out, with the
high density stuff sinking to the middle
and the low density stuff rising to the
surface.
Now the planets that you make will depend upon what material is in the area. The
inner solar system is full of rocky and metallic planetismals, while the outer solar
system is dominated by gaseous and icy planetismals. The material comes together
and makes bigger and bigger chunks. Now you have a great big glob of different
planetismals - so why don't planets look like big mixed up globs of material? As these
things come together and make up the planets, there is a lot of heat released
byradioactive elements (I'll talk more about this stuff later), and this helps to heat up
the interiors of the planets. When the rocks, gases and metals are heated and become
sort of squishy they rearrange themselves, so that the high density stuff (the metals)
sink to the center and the lower density stuff (the gases) rise to the surface. This is just
like how some salad dressings separate their material - high density stuff sinks, low
329 | P a g e

density stuff floats. To be really technical, we can say that the planet
becomes differentiated once the stuff gets sorted out. That just means that the
different elements get arranged according to density - high density in the center, low
density on the surface.
While the planets were forming, some of the big ones, like the Jovians, may have
done the same thing that the solar nebula did. By spinning fast, the clouds of material
that would become something like Jupiter would form into a disk and then parts of the
disk could form into small objects like moons. The big Jovian planets can be thought
of as mini-solar systems as they make their satellites. They have so much mass that it
is easy for them to form satellites. It is also possible that they can catch stray satellites
- objects that formed elsewhere but get too close to the Jovians so they are trapped by
their strong gravity into an orbit.
As planets (and satellites) start to take shape and their surfaces start to solidify, they
will experience a rather nasty episode of the solar system cleaning up. During the first
500 million years of the solar system's history, there would have been a lot of big
planetismals floating around that weren't incorporated into planets. When the planets
and their moons had pretty much formed, these miscellaneous planetismals would
have slammed into them. The earliest part of the solar system's history would have
involved a lot of damage to the planets and satellites. This era is generally referred to
as the Heavy Bombardment era. This would have been the time that the really big
craters currently seen on the Moon and Mercury were made. It should be remembered
that all planets, satellites, asteroids and other objects were hit by these planetismals at
this time as well, but in many cases these impact craters were covered up or erased.
The most intense part of the heavy bombardment era is thought to have occurred
around 3.8-4.1 billion years ago. By this time the modern structures and physical
characteristics of the planets were pretty well established and they ended up as targets
of big planetismals. I should also mention that we think the formation of the solar
system probably started around 4.6 billion years ago - this is a number that most
astronomers and geologists are pretty happy with (I'll explain why in a little while).
Another cleaning up process involved the Sun. While planets and planetismals were
doing their thing, there would still have been some stuff floating around that wasn't
really part of a planet, or perhaps it was material that was ejected by a planet. This
would have included a lot of hydrogen and helium that couldn't be parts of the
terrestrial planets since they were too hot to hold onto them. Most of these light gases
in the inner solar system were cleared out due to the strong solar wind as the Sun
became a true star. The outer planets, however, are far enough away from the Sun that
the solar winds didn't do much damage. Also, because the Jovian planets are so huge,
they are able to hold onto their outer layers better (more mass, more gravity).
330 | P a g e

Figure 5. The various steps in the formation of the solar system. Top left - Gas and
dust cloud - the solar nebula starts to contract. Top center - The protosun starts to
form, pulling in most of the mass. The rotation of the disk increases due to angular
momentum. Top right - The disk forms around the protosun, which is getting hotter
and hotter due to contraction. Bottom left - The temperature of the Sun reaches a
point where it will influence the area around it and start to blow the lightweight
material from the inner solar system, leaving behind only the heavier dust and metals
(mainly). Bottom center - The largest planetismals in the various locations will start
to pull in more material, clearing out the areas around them. Bottom right -
Eventually everything comes together to make the solar system we see today.
It appears from our analysis of this process that it seems to have occurred rather
quickly, on the order of a few hundred million years. We also have some hints about
this process by looking at other young star systems. If you go back to the star
formation part of the course you might remember that we often see disks of material
around young stars. These disks could be indicators of the start of the planetary
system formation process. The disks are pretty big, so it isn't too difficult to see them,
especially with techniques that let us block out the light from the star or viewing them
with infrared telescopes. Quite a few stars have been found with some very tantalizing
looking disks. Many have been observed with infrared telescopes
like IRAS and Spitzer, while in some cases they can be seen with the Hubble Space
Telescope. In some cases it appears that similar things like asteroid belts are also seen
around other stars, as can be seen here. It also appears that the formation of planets can
also lead to their destruction, which is the case seen here. A recent observation by the
infrared Herschel telescope has revealed a water vapor rich disk of material around a
star that is in the process of forming. Of course some of this water will form into icy
objects, some will stay in a gas form and some can become liquid water - that will
depend upon where in the disk the material is located.
331 | P a g e

Some recent observations of young stars find evidence of planet formation, in these
cases around stars that are only a few million years old. The planets that appear to
have formed would be the jovian types, which makes sense due to their large mass
and gravitational pull. Also, observations of very small stars (brown dwarfs) appear to
also show indications of the planet formation process, with dusty disks of material
surrounding them. As more data comes in, we find that the process of solar system
formation is pretty widespread and in some cases, fairly easy
to do.
Figure 6. (click on to see larger image) Some examples of
disks or rings of material around some stars, as seen by the
Hubble Space Telescope. In both pictures, the light from the
star is blocked out so that the dusty disk is easier to see - the
blocked area is indicated by the large circle, while the star itself is pretty small. A
scale for comparison is provided so that you can see how these systems compare to
our solar system. Notice how the rings look sort of like the drawings in Figure 5! AU
Microscopii Image Credit: NASA, ESA, J.E. Krist (STScI/JPL), D.R. Ardila (JHU),
D.A. Golimowski (JHU), M. Clampin (NASA/GSFC), H.C. Ford (JHU), G.D.
Illingworth (UCO-Lick), G.F. Hartig (STScI) and the ACS Science Team. HD 107146
Image Credit: NASA, ESA, D.R. Ardila (JHU), D.A. Golimowski (JHU), J.E. Krist
(STScI/JPL), M. Clampin (NASA/GSFC), J.P. Williams (UH/IfA), J.P. Blakeslee
(JHU), H.C. Ford (JHU), G.F. Hartig (STScI), G.D. Illingworth (UCO-Lick) and the
ACS Science Team.
Now you are probably wondering how we know when all of this stuff happened. It's
due to the presence of radioactive material - I told you I was going to explain this
stuff. There are some things out there that haven't changed since they formed in the
early days of the formation of the solar system. These are rocks that sometimes fall to
the Earth and are picked up as meteorites. Meteorites can tell us a lot about the early
solar system (unlike Earth rocks, which have been reprocessed many times and aren't
"pristine"). When you look inside of a meteorite, you'll find, on occasion, some
radioactive material. What's so important about this? First of all, radioactive material
that is trapped inside of a rock must come from an energetic source, and the best
candidate for that is a supernova. Also, the fact that the radioactive material was
trapped inside of the rock fairly early in the history of the solar system indicates that
the formation process was pretty quick. We know this because we find within
meteorites short lived isotopes of aluminum 26 and iron 60 - if the supernova
happened a long time ago, these isotopes would have decayed before being trapped in
the meteorites. Most importantly of all, radioactive material can be used to determine
the age of the solar system.
332 | P a g e

How is this done? We use the radioactive material's half-life to get the age of the
material. The half-life is the time it takes for 1/2 of the radioactive material to decay,
often into a non-radioactive form. For example, Iodine 129 decays into Xenon 129
with a half life of 17 million years. If there is a significant amount of Xenon 129 in a
meteorite, then there must have been a lot of Iodine 129 trapped in the sample to
begin with and there should still be some left. One thing about half-lifes is that you
are always cutting the radioactive material in 1/2, so you never actually get down to
there being no radioactive stuff remaining. If the rock were originally all Iodine 129
and only 1/8 of it is still Iodine 129, you know that it went through three half-lives (it
was cut in half three times, which gives you only 1/8 of the radioactive material left).
The rock would be about 3x17 = 51 million years old.
Not only is the radioactive material useful in determining the age of the rock, it also
provides some valuable information by its very presence -
1. The formation of rocks took place rather quickly since a large amount of
radioactive stuff was locked into these early rocks. If there was a significant
time lag, then more of the stuff would have decayed and less radioactive
material would be included in the rocks. This tells us that the solar system
formation process was rather quick, maybe taking only a few million years.
2. There must have been a high energy source for the radioactive elements to be
produced in the first place, before the solar nebula started the planetary
formation process. The most likely source would be a type II supernova (based
upon the composition of meteorites).
It is also possible that a supernova explosion occurred near the current location of the
solar system. It is also possible that the explosion could have triggered the formation
of our solar system! Remember, you need something to start a gas cloud collapsing -
some sort of shock wave - and a supernova is a good source of not only the
radioactive material but also the push needed to start the solar system forming. It is
possible that the creation of our solar system was due to the death of another star!
Other Solar Systems
Now that we've looked at how our solar system formed, we need to ask whether we
got it right. Are we correct in our theories? The only problem with that is that we don't
have much information about other solar systems (actually, we should call them
planetary systems, since solar refers to our Sun). Are there any other planetary
systems? Yes, there are. Actually, astronomers have found many planets outside of
our solar system.
333 | P a g e

How do you do that? Actually, it isn't very easy. It is very difficult to actually see a
planet next to another star, mainly because the light from the star would be so bright
that it would be nearly impossible to see any little object next to it. The main method
of finding planets is to see if the stars they orbit have small velocities visible in their
spectra. This is sort of like the way that spectroscopic binary stars are detected - by
looking at their changing spectra. The only difference here is that the star will not
move very fast if it is being pulled by a puny (when compared to the star) planet.
Astronomers look for velocities that are only a few m/s. In a spectrum, this
corresponds to a blueshift or redshift of a fraction of an angstrom - and such a small
shift is very difficult, though not impossible, to see.
Figure 7. Click to see full size image. A chart showing
the orbital sizes and masses of some of the planets
discovered outside of our solar system. The yellow dots
represent the stars they orbit, and the blue dots represent
the planets, though they are of course not to scale. The
extra-solar planets' masses are given in terms of Jupiter's
mass. The distances of the planets from the stars they
orbit are given in A. U.s. Image from the Exoplanets
Website.
Astronomers are so clever that they've come up with
equipment and ways to accurately measure such small
redshifts and blueshifts. There have been over 750 planets
discovered so far (as of August 2013 - odds are it will be up to 800 by the time you
read this). However, we should be careful in calling them planets. It is possible that
some of the objects could actually be very small stars (brown dwarfs), but they
usually have masses that are very similar to planets in our solar system. Most often the
planet's masses are similar to Jupiter or Saturn (hundreds of times the Earth's mass)
and are found within 5 AU of the star they orbit. If you click here you'll see a graph of
many exoplanet masses plotted versus their orbital distances. It also shows the masses
and distances of planets in our solar system for comparison. Only a few planets have
been found with masses similar to the Earth or smaller. The smallest one discovered
so far has a mass that is about 100 times less that of the Earth's. The masses of these
objects aren't known precisely, since the tilt of their orbits makes it difficult to
measure the actual size of their orbits. At this time, this method will only tell us about
the most massive planets, since they will have greatest influence on the stars they
orbit.
There was a breakthrough in the planet hunting area when the first picture of a extra-
solar (outside of our solar system) planet was obtained. A planet was found around the
star 2MASSWJ1207334-393254 (or 2M1207 for short). In this case the planet is large
334 | P a g e

enough to be visible to earth-based telescopes. Images of the planet were obtained
using the VLT as well as the Hubble Space Telescope. In a couple of other cases planets
have been "seen" when they have passed in front of or behind the star they orbit,
causing a very small eclipse of the star's light. In 2008 the Hubble Space Telescope spied
a planet around the relatively bright star Fomalhaut, which also has a dusty disk
around it. And at the exact time that discovery was announced, the folks at
the Gemini observatory proved an image of 3 planets around another star!
While most exoplanets have been found by looking at the changing spectra of a star, it
is possible to find some planets by looking for them to eclipse the star they orbit. In
2009 a spacecraft was launched which had the job of looking for such objects.
The Kepler mission has already shown that it can find small objects around stars by
carefully measuring the light variation caused by a planet's passage in front of the star
it orbits. In August 2009, one of the first tests of the spacecraft showed that it could
observe passages of even small earth-like planets in front of stars. The press release
about that event is here. In March of 2011, the Kepler project announced the discovery
of 1,235 candidate planets around various stars. The graphic available here shows the
stars to scale, with the black dots representing the relative sizes of the planets that
were possibly discovered. The lone star near the upper right is the Sun with the black
dot representing Jupiter. One of the most surprising aspects of the Kepler mission is
that it is only looking at a small part of the sky - about the area of your hand held at
arm's length. At the present time the Kepler mission may be over due to technical
problems, but it has possibly found over 3000 new planets - though many of these
discoveries have yet to be verified (remember - science is about evidence, so more
results from various observers help or hinder any potential discovery).
The results from these planet searches are a bit confusing, though. Many of these large
planets are found in orbits that are much closer than the Jovian planets in our solar
system. How is that possible? Shouldn't only small planets be found near a star? In
most cases all we know about the objects are the masses - not what they are made out
of. In one case astronomers were able to estimate the density of the planet and found a
value that is between that of water and rock. It is thought that this object (Gliese 436)
may be made of rock, gases, and liquids in a manner that would survive around the
star that it orbits. Also, many of the planets have very elliptical orbits. Why aren't
their orbits circular like the planets in our solar system? There must be some way that
these large planets can form close to a star. We're still working on that one.
Regardless of these little dilemmas, it is still sort of neat knowing that there are quite a
few planetary systems out there, so that it is possible that we are not alone!
Probably one of the most intriguing result to come from the planet searchers was the
discovery of a planet around Gliese 581. This planet is located at a distance around its
star that is within the star's habitable zone. This zone marks the region where water
335 | P a g e

on the surface of a planet can exist as a liquid. In our solar system the zone goes from
around the area of Venus out to around Mars, with the Earth right in the middle.
Scientists generally view the existance of water in a liquid form as a requirement for
life to exist. Personally I would have thought that coffee was more important, but
that's just me. Remember, just because the planet around Gliese 581 is located within
the habitable zone though doesn't mean there is life on it - it just means that
water could exist as a liquid. Be careful when you read such announcements in the
news about major discoveries - just because they say water could exist, doesn't mean
it actually does exist on the surface of this planet. The Kepler mission has fournd
nearly earth-sized planets that are located in Habitable Zones of stars similar to our
Sun, but of course just because a planet is found in the correct location doesn't mean it
has water on the surface. Just be on the look out for more news in the future about
such systems.
While liquid water has yet to be confirmed on a planet outside of our solar system, it
does appear that water in other forms exists on planets. The first indication of this was
in the spectra of a planet around the star HD209458b, which appears to show water in
the planet's atmosphere. While this may seem exciting, it isn't really unexpected, since
water is a fairly common molecule in the Universe and we would expect to see in
many locations in its various forms.
No matter how you look at it, this is certainly an exciting time for planetary
exploration!

Now that you've read this section, you should be able to answer these questions....
What are the currently defined criteria for an object to be considered a planet?
What type of object is Pluto?
What are the different characteristics of terrestrial and jovian planets?
What observational characteristics of our solar system provide us with clues to
how it formed?
What caused the density gradient in the solar system?
What role does radioactive material play in the formation of our solar system?
How do astronomers search for other planetary systems?
The Earth and Moon
336 | P a g e


Earth Symbol


What's covered here:
What are the physical characteristics of the Earth's interior, surface and
atmosphere?
What are the physical characteristics of the Moon?
Where did the Moon come from?

Even though you might not think of it as such, the Earth is a planet. Here we're just
going to treat it like one. However, the Earth is a very important planet, not because
you live there, but because it provides us with a basis for comparison to the other
terrestrial planets. Of course, if you would like to learn more about the Earth, you can
take a course in Geology or Meteorology (weather) in the Earth Science department
(in no way should this be considered advertising, just plain old good advice) - but
enough of the sales pitch; on to the science! The Earth also has a large satellite (the
Moon). We'll get to that one later.
What exactly do we know about the Earth? What is it made of? Even though
geologists have never dug a hole very far into the Earth, we still have a pretty good
idea about what is going on deep inside under the surface layers that we can see. Our
study of the Earth's structure will start with the center. How can we do that? How do
we know what the center of the Earth is like?
Geologists use the information obtained in earthquakes. Even though earthquakes are
very destructive, they are also very informative since they produce various types of
waves which travel through the interior of the Earth. There are actually many types of
waves that are produced, but we'll only look at two of them.
337 | P a g e

First, we'll tackle the P-waves
They are a form of pressure waves
and are most similar to sound waves
in how they travel (by pushing
material)
Pressure wave and Primary waves are other names for them.
They travel faster than the other type of wave we'll discuss here.
They move via a pushing or compression motion, so they can travel through
both solids and liquids - this is an
important feature!
S-waves are the other type of wave
These move in a rather wiggly,
transverse motion, sort of what you
get if you take a rope and jerk it up and down quickly - a hump will travel
down the length of the rope.
They are also called Shear waves and Secondary waves.
These travel slower than P-waves.
They are produced by moving material up and down, so they can travel only
through solids, not through liquids.
These and other earthquake waves originate from the focus, which is the actual
location of the earthquake's origin. This point can be located far below the surface of
the Earth or it could be very close to the surface; it depends upon the structure of the
Earth's crust at that location. More often you hear about the location of the epicenter,
which is the point above the focus located on the surface of the Earth. This is where
some silly news reporter will often stand, showing you where an earthquake
originated from, though of course they are wrong, since it really originated from the
focus.
There are seismographs all over the world, so geologists can measure the travel times
between the arrivals of the P- and S-waves (remember, they travel at different speeds)
and this gives them an idea of the distance to the location of the earthquake's focus.
First the P-waves will show up, and some time later, the S-waves show up. The
amount of time between the arrivals gives an estimate to the distance of the focus.
This only tells you how far away the earthquake is, not its direction from your
location. To pin down the point from which the waves originated you need to use at
least three separate readings. When you combine the data together, you'll get a pretty
good idea of the earthquake's origin.
338 | P a g e

There are earthquake monitoring stations around the world that can determine where
an earthquake occurred based upon the arrival times of the waves. Also, those stations
will be able to learn about the interior of the Earth based upon how the waves travel
through the Earth (or don't travel through the Earth). Figure 1 shows the location of an
earthquake and how the waves from it travel through the Earth. For regions near the
earthquake, both P- and S-waves are detected. On the other side of the Earth from the
focus, only P-waves are detected. There are also regions where P- and S- waves are
not detected - the shadow zones.
Figure 1. The motions of
P- and S-waves are shown.
The P-waves get through to
most parts of the Earth,
while the S-waves are
much more limited in their
travels. The regions where
no waves are seen are the
shadow zones, which for
the two types of waves
extend over different
distances.
There must be some sort of liquid in the center of the Earth to prevent the motions of
S-waves through it. Also, since the P-waves are slightly deflected when they get in the
middle due to the changing characteristics of the center, we know that there must be a
solid core as well. From this information we can determine the structure of the Earth's
interior, which is diagrammed in Figure 2. The cores are fairly large; about 55% of the
radius of the Earth is contained within them. This gives the Earth a fairly high average
density. Above the nickel-iron (Ni-Fe) core is the rocky mantle, which has a lower
density. At the outermost part of the mantle you will find the lithosphere, which
includes not only the upper mantle, but also the crust (more on this in a minute).
Figure 2. A cut-away view of the
interior of the Earth.
Approximately 1/2 the radius is
comprised of the two iron cores.
Lithosphere refers to the region
containing the crust. Graphic
from the USGS.
This type of interior model is
good since it also helps to
339 | P a g e

explain the strong magnetic field of the Earth. What is needed to make a strong
magnetic field?Planets and satellites can have strong magnetic fields if they have two
characteristics - they need a large core of electrically conducting material and also a
high rotation rate. If both of these criteria are fulfilled, then the object has a pretty
strong magnetic field. I should point out here that there are many things out there that
conduct electricity, not just iron like in the case of the Earth's core. Now, if one of
these features is not present or is really insubstantial, then the magnetic field tends to
be rather weak. In the case of the Earth, the nickel-iron core is both mobile (since it is
part liquid) and large. The relatively fast rotation rate for the Earth also helps.
Amongst the terrestrial planets, the Earth has the strongest magnetic field by a wide
margin. The area of the Earth's magnetic field (the magnetosphere) isn't nicely
shaped but is sort of stretched back by strong solar winds. Particularly strong solar
outbursts, like coronal mass ejections, can significantly alter the shape of the
magnetosphere, causing it to narrow. Here is a little movie showing such an
interaction.
Figure 3. The protective
nature of the Earth's
magnetic field is shown.
The Earth's strong
magnetic field is able to
deflect from the Earth
many of the dangerous
particles from the Sun.
The strong solar wind
causes the magnetic field
to be squished a little bit,
so it doesn't keep its nice
shape. Particles that get
through the magnetic
field get caught up in the Van Allen belts, which are sort of like of collecting areas for
charged particles. Image courtesy of SOHO team. SOHO
The magnetic field is important for the Earth since it helps to protect us from the high-
energy particles from the Sun. Such particles are usually deflected by the field, though
some particles are trapped for a while in the Van Allen Belts. These were discovered
back in 1958 by University of Iowa astrophysicist James Van Allen (I guess that's
why they're called the Van Allen Belts - unless they also hold up his pants). If the Van
Allen belts get overloaded with particles, then they can leak through to the Earth's
upper atmosphere, interact with the gases there and produce what are called
the aurorae (also called the Northern lights). We have more aurorae when the Sun is
340 | P a g e

very active, when there are many sunspots and the Sun is spewing out many more of
those charged particles than usual. Of course, the further north you are, the better your
chances of seeing the aurora.
The core of the earth is amazing, but don't let it freak you out!
Figure 4. Aurora images from the far
north. The two images are from the collect
of Jan Curtis, who has many fine images of
the various forms of the aurora. Check it
out!
The Crust - plate system
The inside of the Earth isn't so exciting; the
real fun doesn't start until you get to the
surface - the crust! The crust is part of the
lithosphere and is located above the mantle. You might be interested to know that
there are actually two types of crust, and they are sort of like pizza crusts. No, they
aren't filled with cheese or anything silly like that - they just have different
thicknesses. As with pizza, you have a choice of thin and thick crust.
1. The Oceanic crust is the thinner of the two, only about 5 km thick. It is made
up of mainly basaltic rocks. We'll be running into a lot of basaltic rocks, and
you probably know these best as the stuff that comes out of volcanoes - lava,
basically. These rocks are typically dark in color and have high densities.
2. The Continental crust is thicker, with a typical thickness of 30 km. The rocks
found here are mainly granites (granitic). This is a very common rock in Iowa,
since there are quite a few buildings made with granite and such in the
midwest. These tend to be lighter in color compared to the basaltic rocks. The
continental crust also has a lower density.
If you were to go into the deepest gold mine on the Earth, you would notice that it
gets hotter as you get deeper. Why? Some people presume that it is the heat from the
core of the Earth and stuff like lava comes directly from the core, but that is not the
case. Lava doesn't come from the center of the Earth. Actually, all that heat is from
the radioactive decay of material in the mantle. This material is just doing what
radioactive matter always does, breaking down and releasing heat. This material is
sort of buried inside of the Earth under the crust, so the heat can't get away easily and
is trapped inside. The heat does slowly move outward toward the surface of the Earth,
like heat in a boiling pot of water on a stove, though of course it moves much more
slowly than the bubbles in a pot of hot water. The basic upshot is that the Earth is
341 | P a g e

slowly giving off heat and cooling down. Of course, it will take a long time for the
Earth to cool down, since it is sort of large (at least for a terrestrial planet) and the
heat can't get out of the Earth easily.
There is all of this heat inside of the Earth, so it causes the rocks in the upper part of
the mantle to act a little strangely. They get a bit gooey - that isn't really a good term
for them, but that is sort of how they act. This upper mantle region is known as
the asthenosphere. The material here only moves slowly, but the point is that it does
move - it isn't rigid and sitting still. You could sort of think of the asthenosphere as a
big pot of hot, gooey chocolate. It moves slowly, but it does move. Its motion carries
heat toward the surface via the energy transport process known as convection. You
might remember this (or perhaps you don't) as the cause of the granules in the
photosphere of the Sun.
As the asthenosphere moves slowly, the crust on top of it also moves. In certain
locations the crust can be broken apart due to the motion of the asthenosphere in
differing directions. This is particularly true for the oceanic crust, which is thinner
than the continental crust. As the crust is ripped apart (slowly), material will fill in the
gap, and this would be mainly in the form of basaltic lava. The oceanic crust is found
at the bottom of the ocean, so we refer to this process as sea floor spreading. New
oceanic crust is being produced by the influx of lava into the rift (crack in the crust)
and this motion of the crust can also lead to increases in the size of the oceanic crust.
The best example of this is seen in the ever increasing gap between Africa/Europe and
South/North America. At some time in the past, these continents were connected, but
the sea floor spreading has increased the distance between them - and is continuing to
increase the separation. If you look at the location of this crack, you'll find the island
of Iceland, which is a great big volcanic land mass - it's just full of all sorts of
volcanic activity. Not only do you have volcanoes in Iceland but also things like hot
springs, since heat is able to easily escape there. Even though Iceland isn't on the
bottom of the ocean, all of the activity seen on Iceland is due to sea floor spreading.
342 | P a g e



Figure 5. (Above left) The convective motion of the
asthenosphere causes the crust to move in some
rather interesting ways. In places where the crust is
thin, as in the ocean, a rift or crack can form. This
will slowly fill in with basaltic material (lava) as the
two pieces of crust are pushed further apart. The
motions of the continents can be seen in the image
immediately above, which shows the early land mass
that ultimately broke up into the current continents.
The ridge seen in the Atlantic ocean (picture at
right) shows this process as the continents to the
west, North and South America, and the continents
to the east, Europe and Africa, get further and
further apart. Image from NOAA/NGDC.

Does this mean that the Earth's crust is increasing in size? Not really - new oceanic
crust is being created, but the crust is moving along and can collide with other crusts,
which can lead to all sorts of interesting events. The map below shows the various
boundaries between different sections of crust (which are divided into plates). At the
boundaries of the plates you would find some of the more exciting events occurring
on the Earth.
343 | P a g e

Figure
6. The plates
that the crust
of the Earth
is divided up
into are
shown.
Actually, only
the major
plates are
shown; there
are many
smaller
plates that
can't be seen on this scale. Along these plate boundaries are the locations where
earthquakes and volcanoes are prone to occur. Image courtesy of NOAA.
Once one crust starts heading toward another, you don't just have a collision. Usually,
as in the case of an oceanic crust and a continental crust coming together, you
get subduction. This is where one crust is forced below another. Why? It has to do
with their densities. The oceanic crust is more dense than the continental crust, so it
will "sink" while the continental crust will "float" above it. Basically, higher density
stuff goes downward while lower density stuff stays on top. The oceanic crust is
pushed downwards - but that's not all. Due to this downward motion, you tend to get
large trenches in these areas, such as those seen near South America, the Philippines
and Japan. Also, a great deal of heat is generated by the friction of the plate motions,
and the basaltic material is mixed with the continental granitic material, making a
much less runny (fluid) lava. This will work its way to the surface and form volcanoes
near the subduction region, which are different from the volcanoes you would find in
the areas of sea floor spreading (like the volcanoes of Iceland).
Figure 7. Animation showing the subduction process. An
oceanic crust moves toward the continental crust. Due to
its higher density it is forced downward (subducted). The
friction between the two crusts generates a great deal of
heat, which will rise to the surface of the continental
crust and form volcanoes.
Where you have mainly oceanic material, you tend to get Basaltic Volcanoes. These
are usually very low elevation structures and are very spread out, since the lava from
these is rather fluid and flows pretty easily. Volcanoes such as those in Hawaii and
Iceland are of this type.
344 | P a g e

The other type of volcano, found near areas of subduction, are Composite Volcanoes.
Some composite volcanoes are also known as Andesitic Volcanoes, since the rocks
that come from them is andesite (like basalt, but has more silicon in it and other
differences). The lava from these is different than that found in the basaltic volcanoes
(remember, it's a mix of basaltic and granitic material), so these volcanoes tend to be
taller and not as spread out. This is due to the lava being less fluid, so it doesn't spread
out as easily. Volcanoes that fall into this category include Mount St. Helens, the
volcanoes in the Andes, and the Cascade mountains of the Pacific Northwest. People
have to watch out for some of these volcanoes since they can cause widespread
damage. The composite volcanoes tend to be much more catastrophic when they
erupt. They tend to bottle up their energy until they blow their tops rather violently.
Basaltic volcanoes can be pretty bad as well, but they don't get bottled up so much and
instead erupt fairly frequently, so they don't have as much energy to expel at any one
moment - kind of like the difference between constipation and diarrhea.



Figure 8. Upper left, a basaltic volcano seen in Hawaii. Upper right, a composite volcano in
Washington State. The graph directly above shows a comparison of basaltic and composite
volcanoes, in this case the island of Hawaii and Mount Rainier. Notice how the basaltic volcano
has very fluid lava - it's the stuff that is usually shown in movies and this helps to spread out the
volcano. The composite volcano shown at the top, in this case Mt St. Helens, doesn't have as big
an outflow of lava as do Hawaii's volcanoes. Images and graphs from the USGS.
There are also volcanoes that are not produced by continental subduction or sea floor
spreading but by hot spots, places where the crust is broken through by the strong
upward swell of hot material. Hawaii is an example of this. If you were to look at the
345 | P a g e

islands that make up the state of Hawaii as well as the chain of underwater mountains
that stretch off to the northwest, you'd actually be looking at a bunch of extinct
volcanoes. These all came from the same hot spot, but since the volcanoes were built
up on a plate that is moving, the volcanic mountain eventually got too far from the hot
spot to remain active. Currently only the big island of Hawaii is over the hot spot, and
all other parts of the chain are extinct.
While it is easier to break through the thinner oceanic crust, it is also possible to crack
the continental crust. An example of this can be seen in the East African rift, where
the area around Ethiopia, Kenya and Tanzania will eventually break off from the main
continent. There are also hot spots under continental crust, and one place where this
can be seen is in Yellowstone park.
Due to all of this motion of the crust and volcanism, if you want to study old rocks
and old surfaces on the Earth then you need to look at continents (continental crust)
rather than at the ocean floor. There are many parts of the continental crust that do get
re-surfaced by weathering effects or water erosion, so not all parts of the continental
crust are old. When we talk about "ages" of planetary surfaces, we tend to refer to
when the surface was formed. If something like volcanism, glaciers or other erosion
effects has altered the surface significantly, then the surface is "young," especially if
these things happen on relatively short times scales. As you'll see, the surface of the
Earth is pretty young compared to other objects in the solar system. This is something
you should keep in mind when we look at other planets.
You might be wondering about all of these plate motions that have been going on for
some time. Why do they occur? If you remember, the plates move because the
asthenosphere is hot and mobile and it is using convection to move heat from the
interior of the Earth to the surface. You'll remember that the Earth is hot because of
radioactive decay. A planet that is hot is gradually losing heat and as a result you get
things like plate tectonics, volcanoes and earthquakes. A planet that has cooled down
completely would be inactive - no heat to drive any motions. This is something we'll
have to watch out for when we look at other planets.
Atmosphere
Fortunately for us, the Earth has an atmosphere and, of course, the other things that go
with an atmosphere, such as rain, snow, sleet, hurricanes and all those other fun
things. We won't look at that aspect of it (and, of course, if you are interested in
learning more about weather, take an Elements of Weather class). We'll stick more to
the physical characteristics of the Earth's atmosphere.
346 | P a g e

Let's start out with the basic composition of the Earth's atmosphere, which is as
follows:
78% Nitrogen (N
2
)
21 % Oxygen (O
2
)
between 0.1 % and 3% Water (H
2
O)
0.03% Carbon Dioxide (CO
2
)
Even though there is only a small amount of CO
2
in the above list, it is known that the
amount has been gradually increasing over the past few centuries (when we started
measuring the amount of atmospheric gases and how they change over time). This has
been linked to the increased amount of industrialization in the world, but that's a topic
for a Crapstone, uh, Capstone class.
If you look at the composition of the atmosphere listed above, you'll see that it is
pretty heavy in the amounts of oxygen, nitrogen and carbon with just a bit of
hydrogen. Is this the atmosphere that the Earth started out with? No, remember that
the cloud out of which the solar system formed was predominantly hydrogen and
helium, not the heavy elements (like carbon, nitrogen and oxygen) that are found in
the Earth's current atmospheric composition. The first early atmosphere (the primary
atmosphere) had more of the lightweight gases and hydrogen rich molecules, stuff
like H
2
, NH
3
(ammonia), and CH
4
(methane). Due to the close proximity of the Earth
to the Sun, the atmosphere would have gotten a lot of heat; this would have caused
many of these lightweight molecules to break down and most of the hydrogen would
have been lost to space. That's because hydrogen is such a lightweight gas and you
don't have to heat it up very much to get the atoms moving fast enough to fly off into
space. The early hydrogen-rich atmosphere would have slowly drained away. The
current atmosphere is a secondary atmosphere, sort of a replacement
atmosphere. Where did it come from? To answer that we need to take a look at
volcanoes again.
Volcanic gases are made up of the following -
79% Water vapor (H
2
O)
12% Carbon Dioxide (CO
2
)
1.3 % Nitrogen (N
2
)
other small amounts of stuff (argon, sulfur)
This is all quite interesting, but our atmosphere isn't predominantly Water and CO
2
!
That's okay, because we don't have to have all of this stuff stay in the atmosphere; it
could go elsewhere. The water went into composing the oceans around the Earth. Our
planet is at a distance from the Sun where water can be in a range of possible states
347 | P a g e

(gas, liquid and solid). Fortunately for us, most of it went into a liquid state. That
takes care of the first thing on the list, but what about the CO
2
?
With liquid oceans, gases such as CO
2
can be trapped in the sedimentation (rock
formation) processes that are operating. Virtually all the volcanic CO
2
of the early
Earth was removed by the oceans and is currently in the Earth's rocks. If the CO
2
that
is currently trapped in the Earth's rocks were to be released from the rocks, we would
have an atmosphere that would be very nasty and rather toxic to us.
We also have many biological processes occurring in the past as well as today - LIFE!
Things such as pond scum, which would have been quite abundant in the early Earth,
would have been able to help alter the composition of the atmosphere. The amount of
oxygen in the atmosphere of the Earth increased steadily in the early history of the
Earth until it reached a level near that seen today. It is also likely that a lot of the
water vapor could have been broken up into hydrogen and oxygen, with the hydrogen
escaping out of the atmosphere. Some people suspect that a good fraction of the water
on the Earth got here from objects from space like comets (which are composed
mainly of water). Others question the importance of comets in the whole process. Our
current atmosphere is pretty nice, since it keeps us alive and protects us. It and the
Earth's magnetic field help to deflect or diminish the effect of nasty particles from
space. Remember, most types of light can't get to the surface of the Earth because of
our atmosphere. It also keeps us relatively stable, temperature wise. If you ever go
outside on a clear winter night, you may notice that it is very cold. On a cloudy winter
night, the temperature doesn't go down as low. Clouds and water vapor help to blanket
the Earth and trap some heat in. This is what is known as a Greenhouse effect,
though it isn't as drastic as some Greenhouse effects that you'll see. On the Earth, the
main Greenhouse gas is water vapor, but carbon dioxide (CO
2
) can also act as a
Greenhouse gas. As I mentioned before, the amount of carbon dioxide in our
atmosphere has been increasing over time, and some people point to this as a possible
disaster, since it will increase the overall temperature of the Earth.
As mentioned before, when you look at the surface of the Earth, you can see extensive
evidence of erosion processes due to water and wind, as well as other effects -
landslides, volcanism, earthquakes, etc. The surface of the Earth is continually
changing due to these processes and also the influence of so-called intelligent life.
When you look around the surface of the Earth, you see very few impact craters, not
because the Earth avoided getting hit by objects during the early history of the solar
system, but because it has covered up the effects. We got hit quite a bit, but because of
our continually changing surface, the evidence has been removed. A lack of craters on
a planet's surface indicates that there are processes at work that are able to cover up or
erase the craters fairly quickly. Due to this, the Earth is considered to have a
348 | P a g e

very young planetary surface. You may not know this, but there is actually a rather
large crater in the state of Iowa. Just check out this web site for information about it!
Objects such as volcanoes, lava flows, impact craters, sand dunes, ice caps, faults and
large, smooth plains are seen not only on the Earth but other planets as well, as we
shall see. Try to remember what these things are and what they indicate about a planet
when we check them out on other worlds.

Moon

Moon Symbol

The Moon is the only other world that humans have walked on. Due to its rather large
size, it isn't unrealistic to consider the Moon a planet, though technically we can't call
it one since it is really just a very large satellite of the Earth. Some astronomers like to
think of the Earth and the Moon as a binary planet system, but that is just a little too
radical for me.
What are its characteristics? First of all, there is no atmosphere, so that makes that
discussion short. However, because of the lack of an atmosphere, there is nothing to
regulate the surface temperature. Remember how the Earth's temperature can be
influenced by cloud cover? - with more clouds, more heat is trapped in. Without such
a luxury on the Moon, there is nothing to trap or store any of the heat, so there is a
wide temperature variation. This causes the daytime temperature of the Moon to be
349 | P a g e

pretty high, around 400 K (which is 266 degrees F), while the nighttime temperature
is only about 100 K (-280 degrees F).
When we look at the surface of the Moon, we can see some things that are seen
throughout the solar system. These include -
Many craters due to impacts (there are about 30,000 craters visible from the
Earth),
Mare. The name means "sea" and these are of course not seas, since there is no
liquid water on the Moon (since there is no atmosphere). Their smooth, dark
appearances caused early astronomers to give them this name. These are, in
fact, large impact craters or basins that have filled in with dark basaltic lava.
That helps to explain their circular shape. It is interesting to note that there is
quite a gap between when the impact occurred that made the basin and when
the basin filled up with lava. Even though they look pretty big, they only cover
about 17% of the Moon's surface.
Highlands. These are the amongst the oldest regions of the Moon, and are
rougher and more heavily cratered than the Mare. Compared to the Mare, they
are much lighter in color. These are the most common feature, covering about
83% of the Moon's surface.
Mountains. These are found mainly in highlands and associated with the rims
of large impact craters. Various chains of mountains are actually the edges of
craters that haven't been eroded away or covered up by the formation of the
Mare.
Volcanic features. There are only small numbers of volcanic mountains as well
as a few lava flow tubes - objects known as rilles.
Large crack-like features, which are technically known as graben (a type of
fault). Recently careful analysis of these features have classified some of the
faults as scraps, which is a long cliff-like feature, which may be the result of
the cooling of the Moon, and the subsequent shrinkage as it lost heat.
Before continuing on I should say something about craters. Craters are, for the most
part, due to impacts. It can be confusing, though, since sometimes the opening of a
volcano is referred to as a crater as well. I'll try to stick to the usage of craters as an
impact feature. As you'll see, there are many forms of craters - big, small, really huge,
and so on. Some craters have a peak in the middle, sort of as a bouncing-back effect;
some don't. Some craters have rays (bright streaks radiating from them); some don't.
Some craters have well defined rims; some don't. There are many reasons why craters
have certain features, which can include the type of terrain that they were produced in,
the size of the object hitting the surface, the angle and speed with which it hit, and the
amount of erosion that can remove or cover features. What might produce a certain
350 | P a g e

type of crater on one world will produce a different type of crater on another world -
just something to keep in mind.
The surface of the Moon does not experience significant changes to its features like
the surface of the Earth does, so it has a very old surface. Actually, it appears that the
Moon is geologically dead. This is because there does not appear to be any current
volcanism and there has been very little change in the crust for some time. Why is the
Moon so "dead" and the Earth so "active" (geologically speaking)? First of all, you
have to remember where the activity comes from - heat. That heat is mainly produced
by the radioactive decay of material. The Moon has much less mass in it compared to
the Earth, so there was much less radioactivity and therefore much less heat. The
Moon is a much smaller object with a much thinner mantle, so it doesn't take very
long for the heat to escape into space. The Moon has pretty much cooled off. This is
sort of how big and small objects cool down - big things take much longer to cool
down completely because they are so darned big! A puny thing like the Moon cooled
down quite quickly. The
Moon is pretty dead
today.
Figure 9. A current
model of the lunar
interior based upon
seismic information
obtained by the Apollo
mission instruments. This
diagram and most
images in this section are
from NASA.
One interesting feature of
the Moon is that its
composition is very
similar to that of the
Earth's mantle. This
provides a clue as to its
origin. Its composition
also tells us that there
isn't much iron in the
Moon. Only recently
have we been able to model the Moon accurately, in part from information initially
obtained during the Apollo missions. It appears that the Moon's internal structure may
be quite similar to the Earth's though of course on a much smaller scale. Some of the
351 | P a g e

rocks on the Moon do have some magnetic remnants in them, indicating that some
time in the past the Moon's magnetic field was stronger, possibly when it was hotter
and more active. Today it really doesn't have a significant magnetic field.
We can only see one side of the Moon from the Earth, so it wasn't until the 1960s that
spacecraft traveled to and photographed the side that is turned away from the Earth.
Surprisingly, the two sides of the Moon have quite different characteristics.
The Near Side is the side that always faces the Earth. It has many large mare
but not a lot of old craters.
The Far Side faces away from the Earth and has many impact craters and
highlands. There are significantly more craters seen on the far side, and for the
most part, no mare.
Why the differences between the two sides? There is a very simple reason for the
difference. The crust on the far side is thicker. If you measured the average elevation
on the two sides of the Moon, you'd see that the near side has a lower than average
elevation, while the far side has a higher than average elevation - it's thicker on the far
side! Due to this, impacts, even large impacts, did not break through the crust and fill
in with lava to produce mare in the way this happened on the near side, where the
crust is thinner. Click here to see a movie of the Moon rotating around. (Movie from
Calvin J. Hamilton).
352 | P a g e


Figure 10. Views of the Near
and Far sides are shown
above, illustrating how the two
sides differ. A cross section of
the Moon interior is shown at
right, showing the differences
in the near and far side crusts.
A slightly thinner crust on the
near side allowed for the
formation of Mare there. Moon
images courtesy of NASA.

353 | P a g e

Why do we only see one side? It's because the Moon is tidally locked to the Earth.
This stuff was discussed way back in the section of the course dealing with the phases
of the Moon and that stuff, so you may want to refresh your memory and take a look
at that material again. Our tidal pull on the Moon is so strong (since we're bigger) that
we have locked one side of the Moon to always face the Earth. The Moon pulls on us
as well, causing things like tides, and these raise the ocean levels twice a day in some
parts of the world. The Moon is pulling on the Earth and we are pulling on the Moon.
We've been pulling on it (and it has been pulling on us) since the day the Moon
formed. You should also remember that the tidal pull of the Moon is slowing down
the rotation rate of the Earth, so our days are getting longer and longer. We know that
in the past, the Earth rotated much faster and the Moon was closer to the Earth in its
orbit. During these early years, the Moon would have been moving around the Earth
very fast - remember Kepler's Third Law - so its orbital period would have been much
shorter, perhaps only a few days (rather than nearly 30 days). At this time the tides
would have been many times higher than they are today - big time flooding in some
spots!
As I said, things are changing. The tidal forces are gradually causing a slowdown in
the Earth's rotation, and as a side effect, the Moon is moving further away from the
Earth. Eventually the Earth will slow down considerably and the Moon will be so far
away that only one side of the Earth will be aimed toward the Moon. The Earth and
the Moon will both be tidally locked! By the time this happens the Sun will be dying,
so it won't make much difference anyways. We know that this stuff is happening since
we can measure the rotation rate of the Earth as well as the increasing distance that the
Moon is getting from us. It's moving away from us at the astounding speed of 3.8 cm
every year.
Apollo Missions

Figure 11. Images from the
exploration of the Moon by
the Apollo astronauts.
We learned a great deal about the Moon from the six Apollo missions that took place
between July 1969 - December 1972 (were you alive then?). The 12 astronauts who
walked on the Moon spent 300 hours on the surface, gathered 850 pounds of rock and
dirt, and occasionally hit a golf ball around. Due to safety considerations, they landed
354 | P a g e

only on the Near side and in the smoothest places - the Mare. Due to this, most of the
rocks they gathered were from the Mare regions. The later Apollo missions made use
of a lunar rover that allowed them to travel closer to the highland areas and get a
wider range of lunar rocks.
One interesting aspect of the Moon's surface is the layer of material that covers it,
the regolith. This is not a soil, since technically soil needs to have biological
organisms in it. The Moon's regolith is a dry, powdery mix of small rock fragments
and dust. It is thought to have formed from all the years of impacts on the surface of
the Moon, not just big impacts, but the small erosions caused by little things hitting
the surface. The Apolloastronauts collected quite a few rocks, including
Basalts, which are just like the lava that comes out of volcanoes on the Earth.
These are relatively high density and dark in color. They contain iron,
manganese, and titanium. Obviously, most of this came from the mare regions.
These rocks are relatively young, only about 3.1 to 3.8 billion years old.
Anorthosites, which were collected mainly from the highland regions. These
are also seen on the Earth (in the old mountain ranges like the Adirondacks)
and are a mix of aluminum, calcium, and silicates. These tend to be lighter in
color and lower in density when compared to the basalts. These are also older
than the basalts, closer to 4 or 4.3 billion years old.
Breccia, which is just material cemented together by impacts. These would
contain mixtures of different rock types.


Figure 12. Various rocks returned from the Moon. Left, some mare basalt. Notice how porous it
is, very much like volcanic basalt on the Earth. Center, anorthosite, the rock from the highlands.
Right, breccia, formed by impacts that cement together rocks.
While the Apollo missions mark the only time humans went to the Moon, other
missions have continued to study it. The recent Clementine mission (1994) carefully
mapped out some of the more obscure regions of the Moon and made a rather
interesting discovery. Based upon the way that radiation was reflected off of the
surface, it appears that there may be a great deal of water ice located in deep craters at
the lunar poles. This finding was also supported by the later Lunar Prospector mission
355 | P a g e

(1998). How is this possible? It is easily possible if the ice is found in a crater that is
deep enough so that no direct sunlight falls upon it. Remember how cold the unlit side
of the Moon is? - pretty cold. If ice gets into a deep, dark crater, then the ice will stay
there - nothing will be able to reach it to melt it. The ice may have originally come
from ice laden objects like comets that struck the surface long ago. The scientists were
so sure that there was some water up there that the folks at NASA decided to crash
the Lunar Prospector spacecraft near one of the suspected ice rich spots. They were
hoping to vaporize some of the water and then telescopes on the Earth could see the
plume of material. Unfortunately, nothing was seen after the crash, though there could
be several reasons for that, not just that there is no ice. Stay tuned for further
developments in this area.
It took about 10 years before we decided to again check out the Moon in detail and at
this time there have been several missions looking for water on the Moon. Amongst
these are the Chandrayaan-1 mission launched by India, which in 2009 found
evidence for water molecules and molecules that were thought to form from water on
the Moon. This mission looked at the light reflected from the surface of the Moon for
these molecules, so basically they looked at the spectra. Other missions seemed to
confirm this remote detection of water molecules. A much more "intense" mission
was LCROSS, which used direct impacts into the surface of the Moon, like the Lunar
Prospector, but they hit it with 2 components - the booster rocket and the spacecraft
itself. Both hit a crater near the southpole of the Moon October 9, 2009. And the
results? Well, yes, they detected water, but not in huge amounts, but in amounts
greater than previously thought. Mixed in with the water was rocky material, methane,
carbon dioxide, ammonia and carbon monoxide. Observations by the Lunar
Reconnaissance Orbiter also indicated that there may be a great deal of ice located in
the south polar crater Shackleton. The problem with some of these studies is that there
is no direct observation since any sunlight would melt the ice. But by looking at how
light is reflected from the surface using lasers, astronomers have found very smooth
and "shiny" surfaces, which may be due to ice. So you can certainly find water on the
Moon (in ice form), but not a massive amount. If you plan to go to the Moon anytime
soon, you'll certainly want to take along your own filled up water bottle.
Where did the Moon come from? We have quite a few clues as to its origin. There is
the fact that the composition is very similar to the Earth's mantle, and the fact that in
the past the Moon was much closer to the Earth. The lack of water in the rocks and
regolith also tells us something about the way that it formed. Right now, the currently
favored theory about the origin of the Moon is the Impact theory. It is unlikely that
the Moon and the Earth would have been able to form in the early solar system as two
separate objects, since gravity would have tended to pull them together into one large
planet. The basics of the impact theory are that an object about 1/3 the size of the
356 | P a g e

Earth came along and hit the early Earth (soon after it formed and differentiated),
sending up a lot of debris into space. The effects of this early impact on the Earth
would have been covered over long ago, so there is no evidence of it still around
today. The debris that was tossed up into space would have been heated up and would
have lost most of its water (remember, the Moon is pretty dry). Over time this
material would have gradually come together and form into the thing we call the
Moon. At first the Moon would have been pretty close to the Earth; as time went on
the rotations of both the Earth and the Moon slowed down, and the Moon would have
moved further away from the Earth. The chances of such a big impact occurring in the
early solar system is pretty good, mainly because there was so much stuff out there
that could hit planets.
Figure 13. The formation of the Moon
step by step. A large object strikes the
Earth. Only the outer layers of the Earth
are greatly involved in the process. The
impact throws a huge amount of
material into space. Most of the material
was from the mantle, so this explains the
current composition of the Moon pretty
well. Also, the energy from the explosion
would have baked the material
thoroughly, causing the lunar rocks to
be pretty dry. A concentration of
material would develop from the debris
in orbit, and this would gradually come
together and form the Moon. The Earth
would eventually recover from the
impact and be spherical once
again. Click here or here to see this in
an animation. Image and animation
from the Southwest Research Institute.
Now we can pretty much summarize the history of the Moon -
4.5 billion years ago (b.y.a.) - Moon formed (probably via the impact theory). It
should be noted that this time estimate allows enough time for the Earth to form
beforehand. Current age estimates for the Moon always seem to have it slightly
younger than the rest of the solar system.
4.5-4 b.y.a. - large impacts produced basins on the Moon. The impacts on the
near side of the Moon would be eventually covered up by lava (see the next
357 | P a g e

step). Those produced on the far side would eventually get covered up by later
smaller impacts. This is the time of the Heavy Bombardment.
3.5-3 b.y.a. - lava filled the basins on the near side to produce the Mare. The
delay in filling the basins is due to the low level of heat in the Moon and the
slow motion of the lava. The crust on the far side is thicker, so the lava wasn't
able to get to those basins, fill them in and smooth them over. All of the impact
scars remained on the far side and are still seen today by spacecraft that travel
there (since nothing, but more impacts, have ever altered them).
3 b.y.a to July 1969 - not a lot of action; just occasional craters. The Moon was
basically just getting further away from the Earth (which is what it had been
doing since it formed).
July 1969 - December 1972 - People started visiting the Moon. Apollo
11 landed, first of six missions to the moon (Apollo 11, 12, 14, 15, 16, 17).
July 1999 - Lunar Prospector crashed on the surface of the Moon, but no water
was detected.
October 2009 - LCROSS impacts the surface of Moon and detects some water.
The Future - Who knows? Will we ever go back to the Moon? Will there ever
be colonies on the Moon? We'll just have to wait and see.


Now that you've read this section, you should be able to answer these questions....
How can earthquakes be used to probe the interior of the Earth?
What is the interior of the Earth like?
What is needed to produce a strong magnetic field in a planet?
What evidence do we have that the surface of the Earth changes over time?
What causes the continents to move around?
What are the different types of volcanoes on the Earth and what characteristics
do each have?
How did our current atmosphere develop?
What surface features are seen on the Moon?
What are the differences between the near and far side of the Moon? What
causes those differences?
What was learned from the Apollo missions to the Moon?
What evidence exists that indicate the Moon may have formed from an impact?
Mercury
358 | P a g e


Mercury Symbol


What's covered here:
What is unusual about Mercury?
How much have we explored Mercury and Venus?
Is Venus similar to the Earth or quite different?
What are the surfaces of Mercury and Venus like?

According to mythology, Mercury was the messenger of the Gods. This had to do in
part with its rapid motion around the Sun. It is the closest planet to the Sun, therefore
it has the shortest orbital period. Also due to its proximity to the Sun, it is very
difficult to see in the morning or evening twilight. It rarely gets more than 25 degrees
away from the Sun.
Mercury is so close to the Sun that astronomers thought that there would probably be
orbital coupling, like the Earth-Moon system (a tidally locked system). They expected
the rotation period to be the same as the orbital period, so that one side of Mercury
would always face the Sun, just like one side of the Moon always faces the Earth.
Does the orbital period equal the rotation period for Mercury? It is a small planet and
difficult to observe, so astronomers were not sure of its rotation period. They had to
wait for the invention of radar to check it out. By bouncing a radar signal off the
planet, they could measure the speed of its rotation and therefore determine the period
of rotation (sort of the same way a highway patrol officer can get you clocked on the
radar for being slightly over the speed limit on the highway). Are the orbital period
and rotation period the same? No. The period for one orbit is about 88 days, while the
period for one rotation is 59 days (pretty slow spinner, eh?). At first glance this
doesn't seem to be very significant, but if you were to take the ratio of the periods,
59/88, you get a number close to 2/3. What's so great about that?
359 | P a g e

The basic upshot is that there is a kind of coupling between the orbit and rotation, but
not the one that was expected. For every two orbits around the Sun, Mercury
rotates three times on its axis. Figure 1 shows how the orientation of an astronaut
would change over time as Mercury went about the Sun. By the time the planet had
made one orbit, Mercury would have made 1.5 (or 3/2) rotations; one more orbit, and
there would be a total of three rotations - sort of unusual.
Figure 1. An astronaut's view on Mercury
would change slowly. An astronaut is located
on the side of the planet away from the Sun to
begin with (position 1). As the planet moves
around the Sun and rotates on its axis (both
counterclockwise, as seen here), then by the
time the planet has moved 1/3 of the way in its
orbit, it would have rotated around 1/2 way
(position 3). One complete rotation is reached
by position 5, where the astronaut is again
upright, and the planet has gone 2/3 of its way
in its orbit. By the time one orbit is completed
(position 7), the planet has completed 3/2 or
1.5 rotations. Notice how the numbers "two"
and "three" keep popping up here.
The Sun's mass influences the space Mercury occupies and adds some extra distortion
to Mercury's orbit in a manner known as orbital precession. Observations of its orbit
over many years showed that the locations of perihelion shifts (or you could say its
orbit wobbles). Early astronomers didn't know what caused it, and even Newton's law
of gravity wasn't able to adequately explain it. The distortion of space around the Sun,
as described by Einstein's theory of General Relativity, helps to explain the cause of
the orbital precession of Mercury, so that is one mystery that was solved by a better
theory (not to say that Newton's law of gravity is lousy, but sometimes it needs help).
We don't know a lot about Mercury, in part because we haven't really spent a lot of
time exploring it. Also since it is so close to the Sun and relatively small, it is very
difficult to view from the Earth, even with the largest telescopes. The earliest close
exploration was by the Mariner 10 spacecraft which flew by in 1974-75 and was able
to photograph only 1/2 of the surface. For nearly 34 that was the best information
available. In 2008, a spacecraft named MESSENGER flew past the planet for the first
time. After that it flew by 2 more times before changing its orbit enough to enter into
a long-term orbit about the planet. Since March 2011, MESSENGER has been a
"satellite" of Mercury and has been providing us with a significant amount of new
information about the planet. Before 2008, we only had images of 1/2 the surface and
360 | P a g e

now we have about 98% of the surface imaged. Expect the information in this section
to change quite a bit over the next few years.
Figure 2. Two views of
Mercury from the
MESSENGER spacecraft.
To the left is a true color
image showing how
Mercury would look to the
eye. The image on the
right is a false color
image based upon filtered
images that will be used to
analyze the geological
features and material
composition. Image from
NASA/Johns Hopkins
University Applied
Physics
Laboratory/Carnegie
Institution of Washington.
What does Mercury look
like? There are craters
scatter about the place,
just like some aresa of the Moon, but not as densely concentrated as the Moon's far
side. There is some evidence of large volcanic plains - not exactly like the mare on the
Moon, but relatively smooth in places. However, these plains are older than the mare,
so they have more craters on them (remember, it took a while for the mare on the
Moon to form, and there is less cratering there). Basically, the plains on Mercury are
sort of in between the ages of the heavily cratered highlands of the Moon and the
smooth lunar mare. Also like the Moon, there is no atmosphere. This will, of course,
cause there to be wide variations in temperature since heat is not retained at night. The
daytime temperature is much greater than on the Moon due to Mercury's close
proximity to the Sun. During the daytime on Mercury the temperature reaches 700 K
(800 degrees F), while the nighttime temperature is similar to that found on the Moon,
about 100 K (-280 degrees F).
361 | P a g e

Figure 3. A side-by-side
comparison of Mercury and the
Moon (left and right
respectively). Note the lack of
mare on Mercury, though the
cratering isn't as densely
concentrated as on the Moon's
highland regions. Images from
NASA.
There are some features seen on
the surface of Mercury that are
also observed on the Moon.
Amongst these are the scarps.
These are long cliffs that stretch
along the surface for hundreds of
km and can be up to 3 km high, which is significantly larger than those seen on the
Moon. And while the Moon has the large impact basins that are now the mare,
Mercury has older features that are not as smooth. The most prominent is the Caloris
Basin, a multi-ringed feature that is about 1550 km in diameter. Just by chance the
Caloris Basin was located on the boundary of the lit and unlit sides of Mercury when
the Mariner 10 spacecraft photographed it in the 1970, so only about 1/2 of it was
visible. The MESSENGER mission has been able to photograph the rest of the basin -
and actually we found out that it was a larger feature than we initially thought. This
means that if you took my astronomy course before 2008, you learned the wrong size
for the Caloris Basin - I told you astronomy was annoying! Just follow this link to see
the full image of the Caloris basin, and the previous assumed size (in yellow).
There are some puzzling features on Mercury that are a bit unexpected. First are the
relatively large lava flow features visible near the planet's north pole. While this only
covers a fraction of the surface, it is quite thick. Also there are a variety of features
called "hollows" which appear to related to volcanic activity on the surface. Hollows
appear as pitted features on the surface and they tend to be bluer than the other terrain,
which makes them rather distinct. It is also possible that these hollows may be
currently changing, indicating that Mercury may still be geologically active.
If you were to look at the side of the planet opposite of the Caloris Basin, you would
see a region known as the Jumbled or Weird Terrain. This is a very hilly area
covering about 500,000 square km. Is it a coincidence that these objects are on
opposites of the planet? - probably not. The likely scenario is that the impact which
produced the Caloris Basin produced a big shock wave, which spread out from the
impact site. The shock wave would eventually run into itself on the other side of the
362 | P a g e

planet, and when this happens the material in the area would suffer the consequences.
This is sort of like the effect you get when two water waves come together and spray
upward.

Figure 4. In the image on the left, various surface features are shown. The pink arrows show
large, older craters that have experienced lava floods (smoothed out). A scarp is visible running
through a crater on the bottom (blue arrow). The yellow arrow shows some of the unusual bright
material that has been observed in craters on Mercury, but have yet to be explaiend. Green
arrows show chains of secondary craters, which were produced by other impacts which throw up
a large amount of material. To the right is a false color image of the Caloris Basin showing the
composition variation of the basin relative to the surrounding area. The basin's area is shown as
the yellow-ish terrain. Image from NASA/Johns Hopkins University Applied Physics
Laboratory/Arizona State University/Carnegie Institution of Washington.
Recent findings from the MESSENGER mission indicate that it is possible for there to
be water (in the form of ice) in some of the polar craters. These areas show relatively
bright reflections from the IR instruments on the spacecraft was well as other
363 | P a g e

indicators that water ice my exist on the surface in these dark craters. If that is the
case, that would be another feature that Mercury shares with the Moon.
Mercury also has a very weak magnetic field, which was first detected by the Mariner
10 spacecraft and has been accurately measured by MESSENGER. It's about 1% of the
Earth's in terms of its strength. That's not very good since Mercury is so close to the
Sun, and the strong solar winds are able to flatten the field down significantly on the
sun-facing side of the planet. Large eruptions of particles from the Sun would have a
much easier time getting to the the surface of Mercury than the Earth.
By looking at Mercury's density and mass, we can determine the likely composition of
its interior. Mercury appears to have a very large iron-nickel core, making up a larger
fraction of its interior than the Earth's core. Why is the magnetic field so weak? In
part, because the planet is not a fast spinner (remember, one rotation takes 59 of our
days), and it is likely that this core is pretty rigid (not liquid like part of the Earth's
core). Mercury is such a small planet that it would have cooled off very quickly in its
history, while the Earth is still pretty hot. All of these effects make Mercury's
magnetic field pretty weak.
Figure 5. Mercury's interior is compared to the
Earth's. While the Earth is a larger planet, the
fraction of the interior taken up be the iron rich
core is much less than the fraction of Mercury's
interior. In fact, Mercury's core extends about
75% of the radius from the center - the Earth's
iron core extends about 55% from the center.
There is a actually a problem with Mercury
having such a large iron core. There is just too
much iron - objects in the solar system don't have such a large amount of iron in them,
and it seems unlikely that the planet should have formed with that much iron in the
first place. Currently astronomrs theorize that Mercury in the past was a much larger
object with a larger layer of rocky material. A large scale impact or several impacts
would have resulted in the loss of a large fraction of the lower density surface material
leaving behind a planet dominated by a large iron core. This is a reasonable theory
since you have the Sun nearby and it tends to pull in a large number of objects,
unfortunately for Mercury which may have gotten in the way of one of those objects.
Other theories think that the Sun's high temperature may have also stripped away
some of the surface.
The large iron core can also explain the presence of the scarps. Early in the planet's
history, the rock layer above the iron core (the mantle and crust) would have cooled
364 | P a g e

and solidified before the iron core would have cooled. Long after the crust solidified,
the iron core started to cool down. What happens when iron cools? In case you don't
know, it contracts - gets smaller. There is so much iron in the interior of the planet
that the contractions were pretty major. With the shrinking of the core the rock layers
above it sort of lost their support, so they collapsed in various places forming long
cliffs or, as we call them, scarps. Recent observations from the Messenger spacecraft
seem to support this theory, since it appears that Mercury has shrunk by about 1.5 km
during its history. That might not seem to be a large amount, but it could certainly
effect the surface features.
With the Messenger spacecraft still in orbit about Mercury I will likely have to update
the material in this section of the course again in the future.

Venus

Venus Symbol

In mythology, Venus is the Goddess of Love. That's about the only thing that's nice
about this lady. In terms of overall planetary features, Venus comes closest to being
like the Earth and is often called the Earth's twin. Venus and Earth have nearly
identical compositions, sizes, masses and densities, but that's about all that's the same.
The Earth is slightly more massive and has a slightly larger radius, but not by much. If
you were to view Venus through a telescope you would see a uniformly colored object
because it is completely covered by clouds. The image shown above is actually taken
in ultraviolet light so that cloud structures are seen, though the surface is still not
visible. For the most part, a regular telescope view of Venus is pretty dull, since the
cloud layers are very uniform (smooth) and you can't see any part of the surface. It
wasn't until radar was invented that we could even determine the rotation period of the
planet.
365 | P a g e

Using radar, we discovered that Venus rotates on its axis once every 243 days. It is a
very slow spinner (spinner, not spinster). Actually, it takes longer to spin around once
on its axis than it takes to go around the Sun! A day on Venus is longer than a
Venusian year. That'll really confuse you. Not only is it slow, but it is also spinning
backwards (or retrograde)! If you were on Venus (which would not be a good
situation, as you'll see), you would see the Sun rise in the west, not the east.
The clouds completely cover the planet, so it would be a good idea to take a look at
the atmosphere and see how it compares to the Earth's. Here is the composition of
Venus's atmosphere -
96% CO
2

4% N
2

Very small amounts of SO
2
, H
2
O, O
2
, etc.
Now here's an interesting thing. If Venus and the Earth are so physically similar (in
mass, radius, and density), why are their atmospheres so different? Again, as in the
case of the Earth, Venus's current atmosphere is a secondary atmosphere. It is not the
atmosphere that the planet had when it was formed. Just like on the Earth, the current
atmosphere would have gradually built up over time due to volcanic out gassing. Now
remember, volcanoes release such things as H
2
O, CO
2
and N
2
(along with stuff like
sulfur). CO
2
is the main component of Venus's atmosphere. Why isn't water? Unlike
on the Earth, the surface of Venus was too hot for liquid oceans to remain for long and
help remove the CO
2
from the atmosphere. If you don't have liquid water for a
significant amount of time, you can't get rid of the CO
2
and it just stays in the
atmosphere. It is likely that the water from the volcanoes would have evaporated and
was then broken up by the high temperatures with the hydrogen escaping into space
and the oxygen either escaping into space or staying on the planet and probably
oxidizing the surface rocks (rusting it). Without that water in liquid form, you're stuck
with the nice and toxic CO
2
rich atmosphere. Based upon recent observations of the
atmosphere, it is possible that there was even more water in the atmosphere of Venus
in the past than was on the Earth. But it was just too hot.
No matter where it came from, it's still a rather nasty atmosphere, since it is poisonous
to us - but wait, it gets better. Not only are the clouds annoying since they prevent us
from seeing the surface, but some of the cloud layers also contain sulfuric acid
(H
2
SO
4
), making Venus not a very nice place to live. If the atmosphere doesn't poison
you, it will erode or melt you gradually. But, as you'll see, it gets even better!
The high CO
2
content has another rather interesting effect on the characteristics of the
atmosphere apart from its poisonous nature. CO
2
is particularly good at trapping in the
heat in the atmosphere. This is that fun thing known as the Greenhouse effect. While
366 | P a g e

the cloud cover prevents most light from getting in, once it does, it can't get out again.
The thick atmosphere effectively traps the heat in, and current information from
infrared telescopes and probes on the surface indicate that the surface temperature
there is around 750 K (900 degrees F). There isn't much temperature variation on the
day and night sides of the planet due to the way that the atmosphere traps heat so
effectively. Venus, not Mercury, has the hottest average planet surface temperature in
the solar system.The atmosphere can poison you, erode you, and burn you up - what
else can it do? The other amazing aspect of Venus's atmosphere is that it is also very
oppressive (dense). The air pressure on Venus is about 100 times greater than that of
the Earth's surface air pressure. If you wanted to feel this much pressure on the Earth,
you'd have to dive down 3000 feet into the ocean. Of course, you'd be easily crushed.
This really thick atmosphere results in very low wind speeds, since such a thick
atmosphere can't move easily, though it would easily carry material in it. If the poison,
acid, and heat don't kill you, you'll get crushed by the weight of the atmosphere.
Venus is not high on anybody's vacation list.
The cloud coverage of the planet is so complete that it prevents us from looking
directly at the surface. In spite of this we have been able to explore it to some degree,
or at least we have sent some unlucky probes there. Most of the probes that have
visited Venus were sent by the former Soviet Union. These included Venera
9 and 10 (1975), 13
and 14 (1982),
and 15 and 16 (1984
).
Figure 6. The
surface of Venus as
seen by the Venera
13 spacecraft. The
picture may look a
bit distorted, but
that's intentional.
The top picture
shows the actual
color of the surface. This orangish tint is due to the cloud cover. The bottom picture
has been altered to remove the coloring by the clouds, so this is how Venus would
look under normal lighting. Due to the severe conditions on the surface of the planet,
a spacecraft may have only one chance to take a picture before it is destroyed, so the
lens is shaped in such a way to get as much information as possible. That is why there
is such a distorted view, since it not only shows the ground directly below the
spacecraft but also how things look in the distance.
367 | P a g e

These probes not only landed on the surface but also did some radar mapping of the
surface as they orbited around Venus. As they descended through the atmosphere,
they took samples of it and provided information on its composition, temperature,
pressure, etc. Once they reached the surface they obtained as much information as
they could before the nasty conditions destroyed them. All images from the surface of
Venus are from these Soviet probes and they show a surface that looks very dark -
rather similar to the lunar mare. Some of the rocks look rather volcanic (basaltic) in
origin.
The U.S. did not send as many probes to Venus but has done extensive work on
investigating this nasty planet as well. The Pioneer 1 & 2 spacecrafts (1978) did a
variety of things, including radar mapping of the surface and ultraviolet imaging of
the clouds (to see weather patterns). The Pioneer 2 also dropped a few probes into the
atmosphere. The most data-rich study of Venus was performed by
the Magellan Orbiter. This spacecraft was in orbit around Venus for several years
(1990-1994) using a very high resolution radar system to map out the surface features
of Venus better than had been done previously. You can't see the surface directly, so
radar is the only way to look through the clouds to "see" the surface. You have to be
careful, since a radar image isn't the same as a photographic image. The radar shows
mainly the texture of the surface. Areas that are smooth appear to be very dark on a
radar image, while areas that are very rough appear very bright. By combining this
information with elevation data, the Magellan spacecraft was able to provide us with a
high resolution global view of the planet.
Figure 7. (Click on the image to get a larger version
showing more sides of the planet.) An image of the
surface of Venus as seen by the Magellan spacecraft.
Remember, the light-dark areas are not the colors of
those regions but indications of the roughness of the
surface, with the smooth areas being dark and the
rough areas being very bright. If you would like to
see a movie showing how the globe of Venus looks as
it rotates, just click here. Images from NASA,
animation from Calvin J. Hamilton.
Currently the Venus Express space craft is in orbit
about Venus, mainly studying the atmosphere and cloud layers in great detail. This
mission was launched by the European Space Agency (ESA), and it was pretty quick -
only 3 years between the time the mission was approved until the launch date. The
spacecraft arrived at Venus in April 2006 and is still in orbit about it. In general
Venus appears to have a rather boring cloud layer, since it usually appears to the eye
to be uniformly colored. However it is possible for the Venus Express(2005-present)
368 | P a g e

to image different layers of the atmosphere using infrared and ultraviolet cameras.
This has allowed astronomers to understand the temperature structure and wind
circulation patterns in the atmopshere. The atmosphere structure of Venus has
variations like the Earth, with the temperature being very high at the surface, cooling
with elevation, and then getting hotter again as you get further up. By the time you get
100 km above the surface, the temperature is down to 170 k (-148 F). Since Venus
rotates much more slowly than the Earth, the air flow patterns are not as complex,
however it is still possible to have some good sized storms present. One
such "storm" was first observed over the planet's south pole years ago, and it is not
clear how it formed or why it exists.
Figure 8. Image from the Venus
Express spacecraft showing the cloud
layers of Venus. The left side is the
"day" view taken with a visible light
camera, while the right side shows
the "night" side of the planet taken
with an infrared camera.
What is the surface like? There are
quite a few similarities between the
features seen on the surface of Venus
and those seen on the surface of the
Earth, though the way that these
features are arranged and how
common they are tells us something
about the nature of Venus's surface
history.
First of all, there are the Lowlands, which cover about 27% of the surface. They often
appear as large circular areas and are often designated as Plains (Planitia)on maps or
globes. They are pretty smooth and flat, so these are the mainly dark features that you
see on the Magellan images of Venus. They also tend to be located in regions of low
elevation, which tells us something about their formation. It is possible that they could
have been formed by large impacts (circular impact basins), the relaxation (slumping)
of the surface, or due to lava flows that filled in circular basins. They do look
suspiciously similar to the mare of the Moon.
The most common type of surface features on Venus are the Rolling Plains, which
cover about 65% of the surface. These are not entirely smooth like the lowlands but
instead are rather hilly. When you look in these regions, you find such features as
impact craters, volcanic mountains and volcanic openings (calderas or craters). In fact,
369 | P a g e

there are many volcanic features seen on Venus. The volcanoes tend to have very low
slopes (they aren't steep) and their structures are similar to the basaltic volcanoes on
the Earth, like those seen in Hawaii. This would be in line with the images of the
surface rocks that the Venera spacecraft obtained, which showed dark rocks that look
very basaltic - quite a bit of volcanic activity on Venus, eh?
The last major surface features are the Continents. There are many small features that
may be viewed as small continents (raised land masses), but there are only two really
big ones.
Aphrodite Terra is about the size of Africa and is found along the equator. It is a
generally raised up region, about 5.6 km above the surrounding plains. There are
mountainous regions on it and evidence of volcanic activity.
Ishtar Terra is the other large continent and has many mountains on it. It is found far
in the north and is only about the size of Australia. Located on Ishtar is the highest
mountain on Venus, Maxwell Montes, which is part of a high mountain range that
extends up to 11 km above the surface. For comparison, Mt. Everest on the Earth is
about 9 km above sea level.
Actually, since there is no sea on Venus, an average surface elevation is used to define
a type of "sea level." This is the level that is used to define the average surface
elevation.
370 | P a g e


Figure 9. The surface elevation of Venus is shown. The lowest regions are marked in
purple, the mid-elevations in green, and the highest parts in yellow. The gray parts
are where the mapping by the Magellan spacecraft was incomplete. To see another
global view of Venus's topography, click here. Actually, this little movie is pretty
good, since the elevation differences between the rolling plains and the continents are
more obvious. Image and animation from NASA.
Are there active volcanoes on Venus? If you look at the surface (mainly
in Magellan images), you see many volcanic structures, not only the usual sorts of
things like hot spot volcanoes (similar to Hawaii) and basaltic volcanoes (similar to
those in Iceland) but also all sorts of unique volcanic structures that aren't seen
elsewhere in the solar system. Some of the structures are so unusual that we are
having a hard time explaining their origins. This link will take you to some of the
features that are seen.
As with most objects in the solar system, there are a number of impact craters.
However, the number of craters is much lower than that on the Moon or Mercury, and
like on the Earth, there are no large, obvious impact features (the lowlands may be
large impact features, but they also may not be). This tells us that there has been
371 | P a g e

erosion at work covering up the big impacts that occurred early in the planet's history
(as happened on the Earth). It is sort of strange that there are really only medium sized
craters on Venus. Actually, the largest crater is Crater Margaret Mead, which is only
275 km across. There are no small craters, since objects that would normally produce
them are destroyed as they travel through the atmosphere and get burned up.
The lack of any old craters and very few craters overall indicate that the surface of
Venus has been redone (resurfaced) within a span of a few hundred million years.
This is a relatively short time on geological time scales. Venus seems to have a very
active surface if it can redo the landscape on such a short time scale. It is likely that
there is current volcanism going on that helps in the erosion process.
While there appears to be volcanism, there does not appear to be evidence of
organized plate tectonics or volcanic mountain chains like those seen with subduction
on the Earth. There are strange volcanic structures such
ascoronae, arachnoids and pancakes located all over the place. It seems that the
volcanism on Venus is more like the hot-spot type of volcanism associated with
Hawaii or Yellowstone park, not that associated with plates. This sort of makes since,
because it seems that the volcanoes are like those of the Hawaiian islands - rather flat
volcanoes that are likely rather basaltic. The cause for this lack of plate tectonics isn't
known for certain, but it may have something to do with composition differences or
the thickness of the crust. Like the Earth, Venus is still a pretty active
world.
Figure 10. Click on the image to see a larger version. A comparison of
the surfaces of the Earth and Venus. The two planets are almost the
same size, so the images are on the same scale. Notice how the
continents on the Earth are well defined and organized, while those on
Venus appear to not be as clearly defined. The volcanism on Venus is
also quite a bit more random with mountains scattered all over the
place, not in nice, neat chains as on the Earth. This indicates that plate tectonics are
probably not in operation on Venus. Images from NASA/NOAA.
Obviously, Venus is a really fun place to visit! It has it all - a deadly atmosphere, acid
rain, high enough temperatures to melt metal, air pressure strong enough to crush a
truck and a lot of volcanism to cover everything with lava. This place is number one
on my vacation list! (NOT!)

Now that you've read this section, you should be able to answer these questions....
How did astronomers measure the rotation rates of Mercury and Venus?
372 | P a g e

What is unusual about Mercury's rotation and orbit?
What is the surface of Mercury like?
What object is Mercury most similar to?
How is Venus similar to the Earth?
How is Venus quite different from the Earth?
What unusual surface features exist on Venus?
What evidence exists that indicates that Venus may still be volcanically active?
Mars

Mars Symbol


What's covered here:
What surface features of Mars are visible from the Earth?
In what ways is Mars similar to the Earth?
Is there water on Mars?
Is there life on Mars?

Mars has always been of interest to astronomers and the public, in part due to its
proximity. Unlike Venus, surface features can be seen even with a small telescope.
You can see variations in the planet over time - during the change of Martian seasons.
Astronomers (and non-astronomers) noted that light and dark areas on the surface
appeared to change their sizes over the course of a Martian year. People speculated
that these variations could be due to vegetation, just like how the surface of Iowa
changes as plants develop and grow over the seasons. Other things that can be seen
from the Earth are the polar ice caps, and it is easy to see them change in size over the
course of the Martian year, just like Earth's own polar ice caps. Also, the length of the
Martian day is very close to that of the Earth's (24 hours and 37 minutes). Also, the tilt
of Mars's rotation axis is about 24 degrees, which is very similar to the Earth's. With
all of these features that were seen on Mars that could be identified with similar
373 | P a g e

features on the Earth, people were rather curious about just how much Earth and Mars
were alike - for instance, could there be life on Mars? Could there be Martians? Most
importantly, do they have digital cable tv? Perhaps that wasn't one of the big
concerns, but you can see why people have long been interested in Mars..
Things became very complicated in 1877 when the Italian astronomer Giovanni
Schiaparelli described many lines criss-crossing the surface of Mars. He thought that
they might be water channels, which in Italian is written as canali (even though there
are dry water channels on Mars it is unlikely that he did observe them, since even with
the best telescopes today these features can't be seen from the Earth). Now you must
remember,channels can be naturally occurring things, while canals are built. People
looking through Schiaparelli's work came across his mention of canali and thought
"gosh, this fellow has seen canals on Mars!" He never said he did, but people who
didn't know what Schiaparelli actually intended to say (channels) just jumped the gun
and made the conclusion that this fellow observed canals on Mars.
Of course, if there are canals then someone (or something) must have built them! Oh,
dear, many people thought, there may actually be Martians! One person who thought
there was life on Mars was Percival Lowell, a rich Bostonian with too much free
time and money on his hands. He actually had an observatory built to study Mars (and
other things - Lowell observatory in Arizona). In the 1890s he spent a great deal of
time making detailed maps of the surface of Mars showing features such as lines,
canals, cities, farm lands (he mapped up to 160 canals). His publications and theories
behind the Martians, their canals and their cities were widely read by the public,
though professional astronomers didn't think he was correct. But who ever listens to
professional astronomers? Lowell's ideas had enough of an influence that people
sometimes mistook his theories for proven facts.
When science fiction writers like H.G. Wells and Edgar Rice Burroughs wrote their
stories, they often used the "information" derived by Lowell and others who believed
in the existence of Martians. The general public as well had perceptions that life
existed on Mars, or at least it seemed reasonable to most people. This was made very
obvious in the panic that occurred when Orson Welles dramatized in a radio show a
Martian invasion of the Earth in 1937. Many people actually thought it was going on
(whether this had to do with the quality of Orson Welles's work or the gullibility of
people, you can decide). I have to wonder about some people, since in Orson Welles's
dramatization the Martians invaded New Jersey, which is probably not very high on
many Earth invaders' lists of first targets.
Why did people really think that there were Martians? In part, because the ideas
presented by Lowell and others made so much sense. The basic idea was that Mars
was a desert world that relied on the elaborate canal system to bring water from the
374 | P a g e

icy polar caps to lower latitudes to grow food. This would explain the changing
surface features over the Martian year. As crops grew the color of the land changed.
Of course, if the Martians were so advanced that they could build canals that crossed
the surface of the planet, then they must be advanced in other respects as well.
Perhaps they have telescopes and with those telescopes they might observe our world
- our blue, water covered world. If the Martians are desperate for water, wouldn't they
love to come here and get our water (and, of course, destroy major cities and
landmarks along the way, until someone like Will Smith or Jeff Goldblum saved the
day - oops, that was a different set of alien invaders - never mind)? At least, that is
what some people thought. You also have to remember what the world was like
around the early 1900s - things like canals were very common on the Earth and in the
lives of many people, and things like rockets were part of popular literature.
Is this really true? How do we know what Mars is really like? The only way to really
study Mars is to go there. Over the years, the US and the former Soviet Union have
sent quite a few probes to study Mars. However, unlike the success that they had with
studying Venus, the Soviet probes were not so successful - pretty much all of them
failed for a variety of reasons. While the US probes haven't been entirely successful,
we have been pretty good at exploring Mars. Amongst the more prominent projects
are -
Viking1, 2, which landed on the surface of Mars in 1976. Each spacecraft
included an orbiter to obtain pictures of the entire surface and landers which
worked up until 1980 and 1982. The data from the surface were valuable,
particularly the weather data and some analysis of the surface soil.
Pathfinder lander (renamed the Carl Sagan Memorial Station) and the rover
named Sojourner landed July 4, 1997. The rover was able to explore and
analyze many rocks and soil samples. The weather data from the station were
also valuable. This project is now over, since both parts have stopped operating
(this was expected).
Mars Global Surveyor, which started orbiting around Mars in 1997 and only
stopped working in November 2006 took many detailed images of the surface.
This mission has produced a great deal of information, not only through the
images, but also with other detectors such as elevation and magnetic field
detectors.
2001 Mars Odyssey, has been orbiting Mars since October 2001. Its primary
mission is to analyze the surface features, in particular to try to determine the
chemical composition of the soil and rocks and obtain information about
radiation and climate as well. This information can help us determine if it is
possible for life to exist on Mars (not big Marvin the Martian type of life, but
375 | P a g e

small bacterial stuff). Its primary mission ended in August of 2004 and its work
has been extended. It is still obtaining images of the surface.
Mars Express, this mission is very similar to the above two missions, but this
was launched by the European Space Agency. They also had a rover (Beagle 2)
with this mission, but it appears that the rover was lost in December 2003. It is
currently orbiting Mars and using cameras and other instruments to study the
atmosphere, surface features, and look for signs of water. The mission has been
extended to last through 2012.
Spirit and Opportunity, which both landed in January 2004 and started
investigating the surface of Mars. Each is equipped with instruments to
investigate the properties of the soil, rocks and terrain and look for evidence of
water. Both rovers have detected evidence that hints that Mars once had a
substantial amount of liquid water in the past. The performance of these rovers
it exceptional, given the fact that the original mission was only for 90 days. By
early 2010, both had been working on Mars for over 2000 days! Unfortunately
the Spirit rover has become stuck in the sand and may no longer be able to
explore the surface, but so long as it has a communication link with the Earth it
will provide valuable information. So even though some think it may
be abandoned, that's not the case.
Mars Reconnaissance Orbiter arrived at Mars in 2006, but is currently having
some problems. It is designed to look for traces of water using high quality
cameras and spectrometers. So far it has obtained about 11 terabits of data -
enough to fill over 2000 CDs (or over 135 iPods). After finishing its primary
mission it now serves as a communication relay system for other spacecraft and
the rovers.
Mars Phoenix landed in the northern part of Mars and started to sample the
surface in May 2008. By November 2008, it stopped communicating with the
Earth. This was expected since the lander is located at a very northerly location
which limits the amount of solar energy it can get for its solar panels. Before
we lost contact, the lander was able to dig in the ground and find frozen water
under the surface. This was viewed as a major finding, though it was long
suspected to exist. No signs of life were discovered.
Curiosity is the most advanced rover to be placed on the surface of Mars. It
arrived on August 2012, and is still going strong. The landing site was selected
since it appears to show many physical features associated with water flow. So
far it has found all sorts of evidence of past liquid water on the surface.
376 | P a g e





Figure 1. Spacecraft and rovers that have studied and are currently studying Mars. From left to
right, Pathfinder (with the rover Sojourner analyzing a rock), the Mars Global Surveyor, the
rover Spirit rolling off of its spacecraft, and the Curiosity rover ready to examine the surface of
Mars.
The missions sent in the 1960s and 1970s were rather elaborate, for there was enough
money in those days so that two spacecraft could be sent (like the Viking mission).
Now the policy for NASA is to send small, simple spacecraft to study planets
(the faster and cheaper policy). While this can provide some rather dramatic results,
at times the limitations of the systems are annoying. It has actually been a bit of a
problem, since in the late 1990s two such spacecraft failed upon arrival at Mars.
While some people think faster-cheaper might be good for the budget, it is bad for
science. Unless the budget changes, this policy probably won't change. That's enough
political discussion; let's get back to the science!
What is Mars REALLY like? While it might be nice to think that you could handle the
conditions on Mars, that will depend upon which part of Mars you are on. This is
because the orbit of Mars is much more elliptical than the Earth's, and this has an
effect on the severity of the seasons in the northern and southern hemispheres. When
Mars is at perihelion (near the Sun), the southern hemisphere is tilted toward the Sun
while the northern hemisphere is tilted away. At this time it gets pretty hot in the
south (where it is summer), while in the north it isn't such a harsh winter. The
situation is reversed at the Martian aphelion (when it is far from the Sun), when the
north is experiencing summer and the south is experiencing winter. Again, the
distance from the Sun has an influence, in this case making the southern winter pretty
cold and the northern summer fairly mild. Obviously, you would want to be located in
the northern hemisphere unless you really like extreme weather conditions, like
Minnesotans. The surface temperature on Mars is most like that at the Earth's south
pole, ranging from very cold to being rather toasty during the summer -
often very toasty in the southern hemisphere during summer. The temperature
variations would not be too painful for humans.
377 | P a g e

Figure 2 The great variations of
Mars's seasons is due to its
extreme orbit. The ice caps
change size during the seasons
as can be expected. This image
is not shown to scale.
There is another side effect from the change in the distance between Mars and the
Sun. During the time of perihelion, the temperature differences between the day and
the night time sides of the planet cause the atmosphere to become more unstable.
Basically, there are a lot of winds at this time, and since there is a great deal of dust on
the surface, this is picked up by the winds. The result is that dust storms will develop
first in the southern hemisphere, but they can get large enough to cover the entire
planet. These are what the astronomers saw changing the surface features during the
Martian seasons - not vegetation or life. It is also interesting to note that these storms
would develop during the time of perihelion - when Mars was close to the Sun. At
certain times, this is also when Mars was closest to the Earth, and there was a lot of
interest in viewing Mars at such times. If those views showed a changing surface (due
to the dust storms), it probably influenced the observers a great deal.
Figure 3. Images from
the Hubble Space
telescope showing the
changing appearance
of Mars due to global
dust storms. The image
on the left is the
"before" view - surface
markings and
colorings are clearly
visible. The image on
the left shows the
global dust storm's
affect on the view - this is the "after" picture. The surface details are washed out by
the dust in the atmosphere, making Mars look rather smooth and basically boring.
Image credit: NASA, James Bell (Cornell Univ.), Michael Wolff (Space Science Inst.),
and the Hubble Heritage Team (STScI/AURA).
Mars has dust storms, so it obviously has an atmosphere. Let's look at that in some
more detail. In terms of composition, here's the breakdown of the elements -
CO
2
- 95 %
378 | P a g e

N
2
- 3 %
O
2
- 0.15 %
H
2
O - 0.03 %
It is sort of interesting to note that the chemical composition of Mars's atmosphere is
most like that of Venus, but there the similarity ends. While Venus's atmosphere is
very thick and oppressive, Mars's is much more diluted. The atmospheric pressure is
about 1/100th the Earth's, so you would have a hard time getting a lung full of air (this
aspect of Mars's atmosphere was highlighted in the movie Total Recall, though it was
exaggerated a bit there). The atmospheric pressure does change drastically over the
seasons as CO
2
goes from an ice form into the atmosphere at the pole experiencing
summer, and at the same time it will condense out of the atmosphere and become ice
at the pole that is experiencing winter.
In 2009 scientists detected traces of methane in the atmosphere. What does that
indicate? Methane is a gas that can originate from biological organisms, or from
volcanism. While it looks like Mars appears to be geologically dead (no active
volcanoes), it is possible for very low level activity to exist. And the other possibility
also exists - that life may be the source for this, but this would be microscopic forms
of life below the surface. We'll just have to wait and see if there are more observations
of this in the future.
Even though there is so little water in the atmosphere, there is 100% humidity on
Mars. The atmosphere is full of water, but since it is so thin, there isn't really all that
much water in the atmosphere (compared to the Earth). If there was any more water in
the atmosphere, it would condense and form frost or snow. A rather interesting aspect
of Mars's atmosphere is that water cannot be in a liquid state. It can only be in the
form of a gas or a solid (ice or frost). Most of the clouds seen on Mars are water vapor
clouds and there is also some water ice seen in the ice caps. On really cold days, some
of the water vapor can form as frost on the ground.
Figure 4. Foggy weather on Mars is shown, with some of the
low lying clouds seen in a complex valley structure. Image
from NASA.
The internal structure of Mars is a bit different from that of the
Earth, in part because of its lower mass. Mars doesn't have as
much material as the Earth, so there is less internal heating. It
also has a much lower density, which indicates that there isn't
as much iron as in the core of the Earth. Mars is mainly made
of rock. Even though it spins around as fast as the Earth, there
isn't a lot of iron in the core. The basic upshot is that there is a
379 | P a g e

very weak and disorganized magnetic field. Actually, there are magnetic fields
detected on only certain parts of Mars.
Figure 5. The magnetic fields of the
Earth and Mars are compared. Earth's is
pretty simple, with one magnetic pole in
the north and the other in the south.
Mars has magnetic bumps all over the
place, with small fields existing in only
limited areas. The lack of a major
magnetic field structure points to a lower iron content in the interior of Mars as well
as less internal heat (so the core, if it exists,
can't be very mobile). Image from NASA.
Figure 6. The internal structures of the
Earth and Mars. The size of the core of Mars
is not well known; it may be larger than is
shown here, but it is probably not mobile or
as iron rich as the Earth's core.
Surface Features
All of the spacecraft that have visited Mars
have provided us with the chance to compare its surface features to those of the Earth.
In this respect, Mars is the most similar in appearance to the Earth. There are many
features seen on Mars that are also present on the Earth. These include
Impact craters of various sizes. There are also some relatively recent impact
features that have been found - in part because we've been observing the
surface so long, that we can see these appear in previously imaged regions.
Impact Basins (we don't have many of these on the Earth, but they are just like
the big ones that we see on Mercury and the Moon)
Dry River channels and other water flow features, as well as regions that show
evidence of large scale water flow
Volcanoes and volcanic structures
Large rift (a break in the crust)
Smaller fracture lines, as well as faults and folds
Polar caps, as previously mentioned they are mainly composed of carbon
dioxide ice and a little bit of water ice
Ice features as well as ice flow features (glacier-type)
Sand dunes, which change when the winds pick up
380 | P a g e

Of course the surface is red in color due to oxidation of the soil - basically a rusting of
the iron in the soil. Let's take a peak at some of the more interesting surface features.


Figure 7. Several surface features seen on Mars, including a
crater (upper left - this one is called the Smiley Face crater - can
you see it?), dry water channels (upper center), ice caps (upper
right) and volcanoes (left). Images from NASA/JPL/MSSS.
As with Venus, Mars does not appear to have any plate tectonics, though it does have
one structure that could be thought of as a continent - a region raised above the
surroundings. This is the feature known as the Tharsis Bulge. It is a large plateau
with several massive volcanoes located on top of it. Tharsis is raised about 10 km
above the surrounding terrain. The volcanoes on it have been extinct for some time, as
have all of the volcanoes on Mars.

Figure
8. Olympu
s Mons as
seen from
above and
from the
side.
Image
from
NASA.
381 | P a g e

The main volcano, and also the highest mountain in the solar system, is Olympus
Mons. This critter is about 600 km across, reaching an overall height of about 25 km.
For comparison, the volcano that makes up the big island of Hawaii, Mauna Loa, is
only about 10 km high. Olympus is very similar to Mauna Loa and the other basaltic
shield volcanoes on the Earth since it is very wide and spread out. Like the volcanoes
of Hawaii, it and the other volcanoes on Tharsis are hot spot volcanoes, created by a
plume of material that breaks through the surface.
As was already mentioned, Mars is a much smaller and therefore much cooler planet
than the Earth, so why is Olympus so big? You would only expect such large objects
on worlds that would be very active and have a great deal of internal heat, right? - not
quite. While it is a similar volcano to the Hawaii ones, there is the important
difference that there does not appear to be any plate motions on Mars as on the Earth.
Due to this, the one hot spot would have continually fed its material to the same
location. Also, the low surface gravity on Mars allows larger structures to be built up.
On the Earth where gravity is about three times greater than on Mars, the weights of
mountains will tend to cause them to not get very large since they tend to sink down.
Figure 9. Valles Marineris.
Image from NASA.
Not too far from the edge of
the Tharsis Bulge is another
very significant feature,
the Valles Marineris, a 4000
km long rift feature. This is
about the same distance as
between New York and Los
Angeles. In places it gets to be about 6 km deep and 200 km wide. It should be noted
that even though it looks like a river channel (like the Grand Canyon), it is not a water
feature but a crack in the surface of the planet. Why such a large crack would be on
such a small planet is another mystery. If you'd like to fly through the valley, just
click here for a big quicktime video through it.
Figure 10. (Click to see larger version) The
elevation map of Mars is shown (from the Mars
Global Surveyor project). The two hemispheres
show a wide range of elevation, with the north
having a lower elevation and the south having a
higher elevation. The Tharsis bulge and Valles
Marineris stand out easily.
382 | P a g e

When you look at the overall surface features of Mars, there are some rather
interesting ways that these features are arranged. First of all, both the Tharsis bulge
and Valles Marineris are located close to the equator. If you look at the northern
hemisphere, you'd see that it is relatively smooth with very little cratering. This is also
where you tend to see volcanoes. In the south there are many more craters and some
large impact basins. Also the elevation in the south is significantly higher than in the
north, which makes the planet a bit unbalanced. Mars is sort of like the Moon in that
one side of the world is quite different from the other, but in this case, it is the
northern and southern hemispheres that are different. Based on the numbers and sizes
of the craters in the north, it is thought that there were active volcanoes as recently as
1 billion years ago. That isn't what you might think of as recent, but it does support
the age difference between the two hemispheres. In early 2008 it was suggested that
the northern part of Mars was smoothed out due to a massive impact that may have
occurred 3.9 billion years ago. This would have also caused it to have a lower
elevation. You can read about this theory at this link.
It is also worth noting that, especially in the north, there are many dry river channels
and also indications of large scale flooding. Wait a minute! Didn't I say that water
can't exist as a liquid on the surface of Mars, that it would immediately evaporate or
freeze? If that's true, how could you have flood features and other water-related
structures?

Figure
11. Evidenc
e of water
on the
surface of
Mars both
in the past
and
currently.
The image
at the far
left is of
water
shaped
features,
comprised
of impact
craters and
then tear-
drop shaped
forms
around
383 | P a g e

them, most
likely due to
the flow of a
great deal of
water. The
image
directly to
the left is of
a crater rim,
which has a
bunch of
flow
features in it
where water
has recently
flowed out
of. Images
from NASA.
We need to look into Mars's past to answer that question. All we need to do is make
Mars's atmosphere thicker (increase the atmospheric pressure/density). If that were to
happen, then you could have water in a liquid form on its surface, since its
atmospheric temperature and air pressure would be suitable for that to occur. This
seems to be situation that must have existed when the large scale water features were
formed. If there was water on the surface at one time, where did it all go? We know
that there is some water in the polar caps, and there are also indications that there
must be a lot of water frozen under the surface in the form of permafrost. In fact,
recent observations by the Mars Reconnaissance Orbiter show evidence
of recent water features that are seen on the edges of craters that perhaps penetrate the
permafrost (frozen water in the soil). In places where the crater walls erode away and
expose the water layers, the water quickly flows out and leaves a rather obvious
discoloration on the terrain. The neat thing about these findings is that it is seen to
repeat according to the Martian seasons.
The Phoenix lander has uncovered frozen water under the surface of Mars, actually
quite close to the surface. If you click on this image you'll see a trench that the lander
dug and the white water ice in it. Also, look at the lower left corner of the trench in
the two images. In the left image you'll see a couple of chunks of ice, but they are
gone in the image on the right. What happened to the ice? It evaporated - technically
it sublimated similar to how dry ice on the Earth goes from a solid to a gas state.
Current studies suggest that a large part of the planet may have been covered with
water in the past, especially in the northern part. Observations by the rovers and
orbiting spacecraft have revealed the presence of minerals that are water-formed
384 | P a g e

features. This includes the formation of things like opals (yes, those gem
stones), gypsum, various silicates and grey hematite, something which may be
produced by water in a liquid form. Large areas of the flat northern hemisphere appear
to have terrain that suggest water features, even shorelines. This link leads you to a
recent study indicating that large oceans may have existed on the surface of Mars in
the past (and I would have to mention that one of the authors of this study, Brian
Hynek, is a UNI graduate in Earth Science!). Follow this link to see the edge of a
crater explored by the rover Opportunity, which shows extensive layering. However,
all of this would have occured a long time ago - a few billion years at least.
If Mars had a water-sustaining atmosphere in the past, why did it change? We're not
quite sure of the answer to this. There could have been several causes, one cause or
who knows what. A change in weather could have caused the atmosphere to freeze
out and that could have been the end of it - no greenhouse effect, no way to thaw the
atmosphere. You also have to remember that Mars has a lower gravity, so it is easier
for gases to escape out into space. In this case, if the atmosphere were overheated, it
may have been lost into space. An impact by a large object may have blown off a
large amount of the atmosphere. We do know that Mars, like the Earth, has had
epochs of large scale heating as well as times of sustained cooling (ice ages). The end
result is that we know at one time there was a substantial amount of atmosphere on
Mars, but now it is much thinner.
Here are some animations showing Mars. First there is a general view of Mars put
together from various views of the surface (animation from Calvin J. Hamilton). Then
there is a view as seen by the Hubble space telescope. In this case, the atmosphere has
a slight effect on the view. The last movie shows the topography (elevation) of the
surface with the purple areas being very low elevation (animation from NASA).
Satellites of Mars
Mars is rather lucky to have not just one but two satellites. This is mainly due to its
proximity to the asteroid belt, where wayward asteroids could be pulled into orbit
around Mars. This is probably where these two moons came from since they are very
small and asteroid like.
Asaph Hall (1829-1907). Photo courtesy of US Naval Observatory
Archive.
The moons were discovered by Asaph Hall (1877), who was working at
the US Naval Observatory. He tried to discover the moons of Mars for
quite some. After a while, he thought they didn't exist or that he wouldn't
be able to ever see them. His wife, Angelina, encouraged him to go back and try some
385 | P a g e

more. A few days later he did spot the moons. He and his wife are immortalized by
having craters on the satellites named after them (craters Hall and Stickney - her
maiden name).
As you can see, the too moons, Phobos and Deimos, are quite irregular in shape - sort
of potato shaped and small. They also appear to have very low densities (mainly
rocky), like the objects seen in the asteroid belt. There's not much more that you can
say about them.

Figure 12. Phobos (far
left) and Deimos (near
left). If you click on each
image, you'll see a movie
of each moon rotating
around. Images from
NASA, animations by
Calvin J. Hamilton.

Now that you've read this section, you should be able to answer these questions....
How can the observations of features on Mars be explained without relying
upon Martians?
What causes the dust storms?
What is the atmosphere like? Which planet has a similar composition for its
atmosphere as Mars?
Can humans live on Mars?
What types of surface features on Mars also exist on the Earth?
How are some of the Martian surface features unique?
Does water exist on the planet?
How are Mars' moons different from ours?
Jupiter

Jupiter Symbol

386 | P a g e


What's covered here:
What characteristics do the two largest planets in the solar system have?
How are they different and how are they similar?
What are their satellites (moons) like?

As it should be, Jupiter was named after the king of the gods of mythology. It
deserves this title because it is the largest of all of the planets. Its mass is about 318
times that of the Earth's. Its radius is about 11 times greater than the Earth's. In fact,
Jupiter is massive enough that you could compare its mass to that of the Sun (it's
about 1/1000 of the Sun's mass). If it were about 50-100 times more massive it could
be a star (a really faint and dim star, but a star nonetheless).
Jupiter is usually one of the brighter objects in the night sky. Amongst the planets
visible to the naked eye, it is usually second only to Venus in terms of brightness.
With even a modest sized telescope you can see all sorts of details of Jupiter,
including cloud features and its major moons. What is it really like? There have been
several missions that have explored Jupiter and the other planets in the outer solar
system. These include -


Figure 1. The exploration of Jupiter has included visits by the Voyager spacecraft (two of them)
seen on the left, the Galileo spacecraft (center) that was recently in orbit about Jupiter, and the
Galileo spacecraft probe (right) that descended into the upper atmosphere of Jupiter. Images
from NASA.
387 | P a g e

Pioneer 10 and 11 (1973, 1974), which discovered the extent of the large
magnetic field. Jupiter's magnetosphere is actually larger in size than the Sun,
and its magnetic field is much stronger than the Earth's.
Voyager 1,2 (1979). These were much more sophisticated spacecraft than the
Pioneer spacecraft and had much better imaging capabilities and more sensitive
magnetic field sensors. Many of the best images of Jupiter and its satellites
were obtained by the Voyager spacecraft.
In the cases of both the Pioneer and Voyager spacecraft, there were limited amounts
of data gathered, since the spacecrafts were just "fly-bys" - they were just going past
their targets on their way to the next target down the line. Right now (August 2004)
the Pioneer 10 spacecraft is about 86 A. U. from the Sun, traveling at about 27,300
mph. The Voyager 1 spacecraft is even further away, around 93 A. U. (14 billion km),
and it's going at about 38,400 mph (that's why it's so much further than the others; it's
going faster). Voyager 1 is currently the most distant human artifact out there.


Galileo. This is the most recent mission of exploration of Jupiter. It was
launched in 1989 and arrived at Jupiter in 1995 after traveling along a very
round-about path. It was in orbit around Jupiter for nearly eight years and
thoroughly explored the Jovian system. One of the most significant aspects of
its exploration of Jupiter was the deployment of a probe which entered the
upper atmosphere of Jupiter - the first direct analysis of the atmosphere of a
jovian planet. It also had higher quality cameras than were on the Voyager
spacecraft. During its mission, huge amounts of data were transmitted by the
spacecraft even before it got to Jupiter (as you'll see later). Every week, a new
item was posted on the Galileo website. But all good things come to an end,
and after the spacecraft had successfully completed all major phases of its
mission, the controllers at NASA had to do something with it. They were afraid
of losing control of it and having it crash into one of the satellites around
Jupiter. This may have caused contamination to the system and we didn't want
to risk that. So they decided (in September 2003) to crash the spacecraft into
Jupiter, where it was quickly destroyed in the upper atmosphere. So Galileo is
no more.
If you were to look at Jupiter in a telescope, what would you see? Clouds, clouds, and
clouds, but not all the same; in fact, the clouds come in various colors due to different
chemical impurities in them. Also, the clouds are not all at the same levels in the
atmosphere, so you are sometimes seeing deeper into the atmosphere in certain places.
The horizontal cloud features are divided into two groups, the belts and zones. The
belts are the darker and deeper parts of the atmosphere, while the zones are the lighter
388 | P a g e

and higher elevation regions. The belts and zones get their coloring from different
chemicals - the light colored zones have ammonia ice crystals in them, while the dark
belts have ammonium hydrosulfide crystals. The zones look white and the belts look
brown. There is also another cloud layer below the belts, one made of
mainly water ice crystals, and above all these layers is one of haze. While the colors
of the clouds are due to these various chemicals, the atmosphere is made mainly of
hydrogen and helium, which
is also what most of the planet
is made of.
Figure 2. The belts and zones
of the jovian atmosphere are
apparent in this image of
Jupiter taken by the Cassini
spacecraft. Image credit:
NASA.
When the Galileo probe went
into the atmosphere of Jupiter
it discovered some very
interesting things that were a
bit of a surprise.
The probe was able to
get down to a depth of
150 km below "sea
level" before it stopped
working. There is no
surface on Jupiter, so
you can define "sea
level" as the place in
the atmosphere where
the gas pressure equals
the Earth's atmospheric pressure at "sea level," or 1 bar. At this height in
Jupiter's atmosphere (pressure = 1 bar), the temperature was around 130 K (-
225 degrees F). The last data from the probe were from a location where the
pressure was up to 22 bars (22x normal Earth air pressure) and the temperature
was about 425 K (305 degrees F). The temperature goes up as you go further
in.
The cloud layers that we thought consisted of (in order from top to bottom)
haze, ammonia, ammonium hydrosulfide, and water were not all detected by
the probe. Only the layers of haze and ammonium hydrosulfide were seen.
389 | P a g e

While this may appear to indicate that the ideas we have about the cloud
compositions are wrong, you must remember, this is just one sample of a very
complex atmosphere. It is sort of like picking one spot on the Earth and saying
that is typical for the entire Earth.
The temperatures of the cloud layers were much hotter and the atmosphere was
denser than expected. Again, we don't know if this is typical for Jupiter.
Winds throughout the atmosphere were pretty high and didn't appear to vary
much with depth. Wind speeds in the atmosphere were on the order of 700 kph
(435 mph).
Generally, it is assumed that most of Jupiter is similar chemically to the Sun,
but this doesn't seem to be the case - at least according to the probe. Chemical
analysis seems to conflict with this in that much less oxygen and neon were
detected, while more than the expected amounts of carbon and sulfur were
seen.
The helium abundance was very close to that of the Sun.
It appears that we have more questions to figure out about Jupiter than we started
with.
One feature that affects how the clouds of Jupiter appear to us is the rate of its
rotation. In fact, Jupiter is the fastest rotator (spinner) in the solar system! Think about
that - an object much larger than the Earth spinning around about two times faster -
very dizzy! It only takes about 10 hours for Jupiter to rotate once. It is not a solid
object, so the rotation is different at different latitudes. The equator rotates the fastest;
the poles go the slowest. This variation in rotation, or differential rotation, we've
already seen in the Sun, and it is also present in the other jovian planets. There are
some consequences for spinning very fast and having differential rotation. One is that
the planet doesn't stay very spherical but gets a little squished, so that the diameter of
the planet measured from pole to pole is less than the diameter measured at the
equator. Basically, Jupiter is wider and shorter than a perfect sphere.
Another side effect of the fast rotation rate are the complex weather patterns. If you
look at the cloud patterns of Venus (1 rotation = 243 days), the cloud circulation
patterns are very long, stretching from pole to pole. For the Earth (1 rotation = 24
hours), there are complex wind patterns such that some latitudes have the winds
toward the east and others toward the west. With Jupiter, this effect is even more
extreme, with many more regions (latitudes) having winds going one way and many
others going the other way. One part of the atmosphere is moving slightly faster or
slower than an adjacent part or even going in the opposite direction, so this will
produce some rather interesting effects. The shearing winds and turbulence in the
atmosphere lead to the formation of circulation regions, or spots (storms). One thing
that has been seen over time is the disappearance and reappearance of some of the
390 | P a g e

major belts. This happened recently in 2010. For this reason, astronomers continue to
observe Jupiter to help understand why the features do change over time.
The largest of these storms is the Great Red Spot. This spot is so large that you could
fit two or three Earths inside it, depending on its size (it's a storm, so it can get bigger
or smaller over time). The Great Red Spot has been observed for about 300 years, but
it is probably much older (you can't see it without a telescope, so it wasn't discovered
until the telescope was invented). It also tends to tear up smaller storms that get too
close or just basically gobble them up completely. There are also other spots, which
appear as white, brown and red features, though they don't appear to be as long lived
as the Great Red Spot. In 1998 and 2000 several spots merged together to form a
larger spot, which again merged with other spots in 2005 to create what is called Red
Junior, a spot that is pretty good sized - not as big as the Great Red Spot, but one that
is holding up. Red Junior passed near the Great Red Spot in the summer of 2006 and
seems to have survived the passage quite well. Another spot also popped up in 2008,
but it doesn't have a cute name yet. We'll have to watch and see how things develop in
the future with these spots. The image that is at the link given above is a false color,
near-infrared image, which is why the spots don't actually look red. Such images show
the temperature variations within the atmosphere and help astronomers determine the
elevation of the various features as well as the temperature and composition structure
of the jovian planets.
Figure 3. Click on the
image to see an animation
of the motion of the Great
Red Spot. Images from
NASA.
Both
the Voyager and Galileo sp
acecrafts have been able to
see the dark side of Jupiter
and observe things that
aren't visible from the
Earth. This includes
lightning that is seen in the
cloud layers. Also, it is possible to see aurorae near the poles (this can be seen either
on the dark side of the planet or using UV telescopes from the Earth). What is causing
these features? It is thought that they are produced by charged particles in the
atmosphere, which interact with the magnetic field. This also indicates that the
magnetic field is very strong. Previously I mentioned that it was large (about 30
391 | P a g e

million km wide), and it is about 19,000 times stronger than the Earth's magnetic
field.
Figure 4. The aurorae of Jupiter as seen by the
Hubble Space Telescope using its UV camera. Image
Credit: J. Clarke and G Ballester (University of
Michigan), J. Trauger and R. Evans (Jet Propulsion
Laboratory) and NASA.
What produces the magnetic field? To answer that we
need to look at the internal structure of the planet. At
the center is a small core of rock and metal - this core
would be about the same size as the entire Earth,
though much more massive. Above that is a large layer
of liquid metallic hydrogen. Above that, there is a layer of mainly liquid hydrogen.
Above that layer is a layer of molecular hydrogen, and at the very top are the
atmospheric layers that are visible to us - water ice crystals, ammonium
hydrosulfide, ammonia and haze at the very top. The internal structure is shown in
Figure 5. As mentioned previously, the composition of Jupiter is thought to be very
similar to the Sun, so there is also a significant amount of helium in these layers as
well. The helium isn't so important here - it is the hydrogen that we should pay
attention to.
Now that you know what the interior is made up of, can you figure out what is
producing the magnetic field? Do you remember what you need in order to produce a
strong magnetic field? You need two things - fast rotation and an electrically
conducting layer. Jupiter certainly has a good rotation rate (more than two times as
fast as the Earth), so what about the electrically conducting layer? There is that pretty
large layer of liquid metallic hydrogen in there. As was stated in Terminator 2 - "it is
liquid metal." While it isn't iron or nickel, it does have some of the same properties as
these metals, and it can conduct electricity. This layer is so large that it helps give
Jupiter the very strong magnetic field that it has.
Figu
re
5. T
he
inter
nal
struc
ture
of
392 | P a g e

Jupiter is shown. The main element is hydrogen, but it is found in various forms inside
of Jupiter. Starting at the layer where the pressure equals the Earth's atmospheric
pressure, the first interior layer is mainly hydrogen gas in molecular form (H
2
). At
about a depth of 7000 km the hydrogen is under so much pressure that it is in a liquid
form. After another 7000 km, you run into the layer of liquid metallic hydrogen. The
core of rock and metal only extends about 10% of the way from the center, so it has a
radius similar to the Earth's but is about 10 times more massive than the Earth.
If you were to go out to Jupiter and measure the amount of energy (heat) that you
would be getting from Jupiter and compare that to the energy that you would be
getting from the Sun, you would note that Jupiter is radiating more heat, about two
times as much. Why is Jupiter hot? Actually, it has never been cool. This is heat left
over from the formation of the planet. There is so much mass that it will take a long
time for it to completely cool off. This is also something that helps keep the
atmosphere active. The heat coming from the center helps in the circulation of gases,
including the rising and sinking motions seen in the zones (rising) and belts (sinking).
Another surprising discovery was that of a very thin ring observed by the Voyager
spacecraft. The ring is rather small and composed of dark, dusty particles, so it cannot
be seen from the Earth. The best way to view such rings is to let sunshine highlight
them, sort of the same way that you can see dust particles in the air if there is sunshine
or another bright light shining through them. As you will see, all of the Jovian planets
have at least one ring. A recent image of the ring obtained by the New
Horizons spacecraft can be seen here.
Satellites
I really hate this discussion of the jovian planets, since the numbers keep changing. It
seems that every few months someone discovers a bunch of new satellites (moons)
around a jovian planet, so I have to keep updating my notes - very annoying!
Currently, there are more than 60 known satellites around Jupiter, which makes it the
champion in terms of "most moons." I'm not going to give an exact number, since if I
do, someone will come along and discover a couple more, and I'll have to change my
notes again. It is not unusual for the jovian worlds to have so many moons; they all
have quite a few satellites. In a way, they seem to be mini-solar systems with a variety
of worlds around them. There are three categories of satellites. First, there are the
really puny ones, ranging from 1 to 300 km in diameter. These tend to be pretty
common and also rather irregular in shape. They are also prone to have retrograde
(backwards) orbits. The next batch has a range of sizes between 300 and 1500 km.
The really big ones are more than 1500 km in diameter, similar in size to our Moon.
Most of the satellites for the jovian planets are in the "puny" category, with only a few
in the really big group. It is thought that many of the puny satellites are actually
393 | P a g e

asteroids that the planets have captured. These are also the most likely satellites to
have unusual orbits or shapes (non spherical shapes). We will be looking only at the
big ones for the most part.
The jovian planets are located far from the Sun, so their satellites tend to be made of
common materials found at these distances. These include a large amount of ice in
various forms, including not only water ice but also CO
2
ice, methane ice and
ammonia ice. There is also some rock and metal, but it is generally much rarer at this
distance from the Sun. Remember, low density material was a much more common
commodity in the early solar system, especially far from the Sun.
The names of the satellites come from mythology and are named after people that
Jupiter was associated with, in some cases in a rather, ahem, "inappropriate manner,"
especially for a married deity (which, of course, did not make his wife happy).
Obviously, he was a rather busy fellow. Not all of the satellites have been officially
named - those that were recently discovered get numbers assigned to them before they
get their official names.
The four big satellites around Jupiter were first seen by Galileo (the guy, not the
spacecraft) when he turned his telescope to the sky in 1610. They are collectively
known as the Galilean Satellites.

Figure 6. The four Galilean satellites. Furthest to the left is Io, followed by Europa,
next is Ganymede and finally Callisto. The satellites are shown to scale. Images from
NASA.
The physical characteristics of the Galilean satellites vary with increasing distance
from the planet. The one closest to the planet is Io, which has the highest density of
the four. It is composed mainly of rock, probably with layers of magma in the interior
(molten rock). Another interesting feature is that Io has virtually no impact features,
394 | P a g e

indicating that it has a very young surface. The next satellite out is Europa, which is
less dense, has less rock and more ice in it, and more impact craters and thus a slightly
older surface. Next is Ganymede, which has a lower density, less rock and more ice,
and more craters. Finally, there is Callisto, with the lowest density, the most ice, and
the most impact features. Of the Galilean satellites only Europa is smaller than our
own Moon. You can think of these as mini-planets. The images in this section are
from Calvin J. Hamilton.
Io looks very different in appearance from any other object in the solar system.
Nothing comes anywhere close to looking like this thing. When the Voyager images
from this object were first seen, someone mentioned that it looked like a pizza. The
scientists knew that there was something strange about Io, since they noted a great
deal of sulfur in the area of Io's orbit. They weren't quite sure why the sulfur was
there. Basically, Io is covered with volcanoes, which were discovered by
the Voyager spacecraft. Currently, there are mainly sulfur and sulfur compounds
given off by these volcanoes. Why are there no other gases? Don't volcanoes give off
other gases? The low weights of the other gases, such as CO
2
and H
2
O, would allow
them to easily escape from the satellite, while the heavier sulfur falls back to the
surface. The sulfur gets recycled over and over, since the volcanoes are continually
active. Click here to see images of the eruptions.



Figure 7. (left) A view of Io as seen by the Galileo spacecraft. The center image shows several
volcanoes erupting at the same time (click on the image to see a larger view). (right) A view of
the interior of Io shows a large iron-nickel core (gray), covered with a rocky layer. Images and
artwork from NASA.
If you go to the surface of Io, you would find sulfur in liquid, solid, frost and gas
forms. It is a pretty stinky place. Eight volcanoes were seen erupting during the
relatively quick Voyager fly-by. About 100,000 tons of material are expelled every
second at a velocity of about 1 km/s. There is so much activity on Io that large scale
395 | P a g e

variations in its surface features can be seen in only a few years. If you added up all of
the lava that comes out of Earth's volcanoes over 100 years, then you'd have the
amount of material blasted out of Io's volcanoes in 1 year.
What is Io's problem? Io is getting heated by the gravitational pull that it gets from
Jupiter and the other moons. Very strong tidal forces are basically pulling and tugging
the planet back and forth - remember, the other satellites are pretty big and therefore
have a good deal of pull on Io as well. This explains why there are no impact craters
on the surface. The continual volcanism will cover up any craters that would appear.
Also, theGalileo spacecraft was able to detect a magnetic field around Io. To produce
the magnetic field seen by Galileo, Io would need a very large iron core, one that
would extend over 1/2 the radius of Io!



Figure 8. (left) The surface of Europa is shown. (center) A view of the interior of Europa shows
a small iron-nickel core (gray), covered with a rocky layer (brown) and a layer of liquid water
(blue). This is all under the visible ice layer of the surface. (right) The various surface features
on Europa are visible - click on the image to see up close views. Images and artwork from NASA.
Europa - Europa, like Io, is also a rather unique place. It is ice covered (which isn't so
unusual), but it has very few craters; only tiny ones are seen and there aren't very
many of those. Why? Like Io, Europa is heated by the tidal pulls of Jupiter and the
other moons, but since it is further from Jupiter than Io, it doesn't get as hot. However,
it is hot enough for the ice to easily flow over the surface and cover up most
distortions (craters). Europa is the smoothest place in the solar system, with very few
craters.
396 | P a g e

Figure 9. The ice cracks on the surface of
Europa. Notice how the cracks are made up of
several parallel groves. These indicate a
freezing, melting, and cracking cycle that allows
the cracks to get filled up time and time again.
One of the more intriguing aspects of Europa is
what is under the icy crust. Based upon the
amount of heating it gets and its composition,
there could be a liquid water ocean under the ice
layer. Europa is considered a likely location
where life could exist in the solar system due to
the presence of liquid water. Of course, we
would have to go there and find it before we
could be sure.
Ganymede - There are many more craters on
this world, but there are also many ice flow features (similar to what was seen on
Europa). This is a bit confusing, since Ganymede is so much further away from
Jupiter than Io and Europa. It is possible that due to its large mass there is a great deal
of internal heating from radioactive decay (Ganymede is the largest satellite in the
solar system). There should also be some tidal heating from Jupiter but nowhere near
the amounts that Io and Europa experience. Data from the Galileo spacecraft indicate
that about 1/2 of the surface has been resurfaced by ice flows or ice tectonic
(volcanic) activity. The Galileo spacecraft has detected a magnetic field around
Ganymede. What produces it? We're not entirely sure, since Ganymede is less dense
than the other two satellites mentioned (remember, density decreases as you get
further from Jupiter). It is possible that there is a small iron core. A large amount of
radioactive material trapped in the interior there early in the history of the planet could
help to explain the ice flow features on Ganymede.
397 | P a g e




Figure 10. (left) The surface of Ganymede is shown. (center) A view of the interior of Ganymede
shows one of the possible models for the interior - a small iron-nickel core (gray), covered with a
rocky layer (brown) and a layer of soft ice (bluish). This is all under the rigid, visible ice layer of
the surface. (right) An up close view of the surface showing the ice flow features along with some
cratering. Images and artwork from NASA.
Callisto - This place is much more typical of the type of surface that is seen in the
outer solar system. There are many craters, including some really big ones. This is
what you would expect since there are fewer tidal forces and less tidal heating due to
its greater distance from Jupiter. One surprising result from the Galileo spacecraft is
the possible presence of a thin, salty ocean under the icy surface. This idea is based
upon the way that the magnetic field of Callisto seems to change over time. The
surface is much older than the surfaces of the other satellites. Some of the really big
impacts could date from the early days of the solar system. The surface is icy, so the
large impact features tend to have a rippled look to them, like the effect of dropping a
rock in the water. It isn't liquid water, but the ice is elastic enough to show ripples.



398 | P a g e

Figure 11. (left) The surface of Callisto is shown. (center) A view of the interior of Callisto
shows a possible model for the interior - a large rocky core, which may be mixed with ice
(mottled gray), covered by a very thin ocean (light blue), all of which is under the thick icy crust
(white). (right) A view of the surface showing some of the craters, including a large impact
feature (Asgard). Images and artwork from NASA.
As mentioned previously, there are more than 60 known satellites around Jupiter, and
you can see a list of them by going here. Some of the names may be familiar to you if
you read any stories from classical mythology, but some names are rather obscure -
and hard to pronounce. Except for the I four I discussed above, the rest of these are
tiny and not really too exciting.

Saturn

Saturn Symbol

In terms of mythology, Saturn was the father of Jupiter (he was also called Cronus).
Saturn is the last of the planets that have been known since ancient times and the most
distant planet that can be seen easily with the eye. In most respects Saturn is very
similar to Jupiter. They have similar compositions and internal structures, so most of
the items mentioned about Jupiter can be directly applied to Saturn. There are some
important differences, though.
While Saturn has a radius that is about the same as Jupiter's (though a little smaller), it
has only about 1/3 the mass of Jupiter. This gives Saturn a much lower density (less
material filling the same volume of space makes it rather fluffy). In fact, Saturn has
the lowest average density of all of the planets. You may have seen cartoons in
elementary school science books showing Saturn sitting in a bathtub of water, not
because it needs to wash behind its ears, but because the average density is less than
that of water, and in that sense it could float in water. That is the average density.
There are, of course, parts of Saturn that have much greater densities than water and
399 | P a g e

parts with much lower densities. The bathtub depiction of Saturn is not quite right (oh,
dear, another childhood illusion destroyed - just don't ask me about Santa Claus).
Another side effect of Saturn having a lower mass than Jupiter is that it has a lower
surface gravity. This causes the clouds layers to be spread out over a larger range. The
cloud layers stretch to greater depths than they do on Jupiter, so the cloud features are
less distinct. Saturn has the same cloud layers (ammonia, ammonium hydrosulfide and
water ice), but they get sort of fuzzed out of view by the haze layer on the top. Even
though the clouds have the same compositions as those on Jupiter, they are not as
prominent, so its storms and spots aren't as obvious as Jupiter's. There are times when
the storms do become pronounced, as was the case of a storm that started in December
2010, which was going strong into 2011. This image from the Cassini spacecraft
shows the 2010-2011 storm.
Saturn also has a slightly slower rotation rate than Jupiter, but it is still going fast and
has differential rotation just like Jupiter, so there are some spots and storms, though
they aren't as easy to see. Also, due to the lower amount of mass and its slower
rotation rate, it has a weaker magnetic field (it's only 1000 times stronger than the
Earth's). The effects of the magnetic field are visible, though. Aurorae can be seen
near the magnetic poles, just like on Jupiter. If you would like to see a movie showing
the motion of the aurora, just click here.

Figure 12. To the left
is an image of the
aurorae visible in the
atmosphere of Saturn.
This is an image from
the Hubble Space
Telescope (UV
image). The image to
the right is also from
the Hubble; this one
shows one of the rare
large storms in
Saturn's atmosphere.
This storm did not
last very long and
was gone in a few
months. Image
credits: J.T. Trauger
(Jet Propulsion
Laboratory), Reta
Beebe (New Mexico
State University), D.

400 | P a g e

Gilmore, L. Bergeron
(STScI), and NASA.
While the planet itself doesn't look so spectacular, it makes up for this by having an
extensive ring system. Actually, Saturn's ring system totally kicks a... well, you know
what I mean. These rings are so impressive that Galileo sort of saw them with his tiny
telescope. A weird thing happened after people read about Galileo's discoveries a few
years later. When others tried to observe the rings at this time, they could not see
them! Why?Saturn is tilted on its axis so that the orientation of the rings changes over
time. There are times, about every 16 years, when the rings are tilted edge on with
respect to the Earth so that they are virtually invisible.
Figure 13.Saturn's rings are
not visible in this view from
August 1995, when the tilt of
the planet (and the rings) was
such that they were virtually
invisible from the Earth. Some
of Saturn's satellites are visible,
though. Image credit: Phil
Nicholson (Cornell University)
and NASA.
The rings are very thin but very spread out (a few tens of meters thick and about
270,000 km wide). When they are edge-on, they are very difficult to see. However,
this is a great time to see the satellites and to look for undiscovered ones. This sort of
alignment doesn't happen very often, so you will usually be able to see the rings to
some degree when you look at Saturn through a telescope.
From the Earth we can see the
following rings - A, B, and C. (This
and other ring images from NASA).
The A-ring is the outermost, while
the B and C rings are closer to
Saturn. While there is an obvious gap
between the A and B rings, the way
you distinguish the B and C rings is
by how they look. The C ring is sort
of thinner or more transparent
looking than the B ring. As
mentioned, there are gaps in the
rings. The two gaps in the rings that are visible from the Earth are the Cassini division
401 | P a g e

and the Encke division. The Cassini division (between the A and B rings) is easily
seen, but you would need a very powerful telescope to see the Encke division. The
Encke division is actually contained within the A ring. Even though the divisions look
rather dark, there is stuff in them, as you'll see.
To understand what Saturn is really like you must explore it, and, of course, there
have been several spacecraft fly-bys of the Saturnian system. The Pioneer 11(1979)
and the two Voyager spacecraft (1980, 1981) flew by after they visited Jupiter and got
better pictures of the rings. Also, by going behind the planet on the dark side, they had
a better view of the rings. Images of the rings showed that there weren't just a few
rings but hundreds of them, each only a few kilometers wide. You could almost think
of the rings as the grooves on a phonograph record. The Voyager spacecraft
discovered several more rings as well, so now there are the D, E, F, and G rings. All
of these new rings except for the E ring are pretty small - the E ring is wide but very
sparse. In 2006, the Cassini spacecraft had discovered another ring, which will no
doubt be the H ring. The only problem with all of these rings is that they are arranged
out of order when you go from the planet outwards. Again, with new technology
comes new discoveries, and things can get a bit messed up with such discoveries. Just
click here to see the rings and their relative locations to the satellites orbits.
Right now the Cassini spacecraft is in orbit around Saturn, just like
the Galileo spacecraft was in orbit around Jupiter. It got to the Saturnian system in the
Summer of 2004, and has already made several discoveries, including another ring
(not yet named) and some new moons (possibly). The Cassini spacecraft also had a
probe. which was named Huygens, not into Saturn, but onto the surface of the satellite
Titan (more about this below). Obviously I'm going to have to keep updating my notes
with this spacecraft making many more discoveries - that's just part of the job.
The rings,
especially the
thin outer ones,
are kept in tidy
order
by Shepherd
Satellites. These
are small
satellites that are found near thin ring sections that use their gravitational pull to keep
the particles in the rings in their places. Actually, most of the ring structure can be
related to the various satellites of Saturn, either as sources for ring material or as
objects that influence the ring structures. Various gaps and inward/outward motions of
the rings are related to the motions of other, larger satellites that move the ring
402 | P a g e

particles around slightly. Click here to see an image showing how the same section of
the rings appears at different locations around the planet.
Figure
14. A
false
color
image
from the
Cassini
spacecra
ft of the
A ring of
Saturn.
This
shows
the
individua
l rings
that
make up
the
larger ring structure and also the material that is found in the dark appearing Cassini
and Encke divisions. The Cassini division is on the left side of the image and is
colored redish, while the Encke division is much narrower, towards the right side, and
is a darker red.
As mentioned previously, the divisions were found to not be completely empty. They
appear to be dark and empty when viewed from the Earth, but when looked at from
behind the planet with the sunlight shining through them, they appear to have particles
in them. Such particles are generally rich in dark material like dust, so they only show
up well when light shines through them, just like the dark, dusty ring around Jupiter.
Most of the major ring sections are made up of particles that have more ice in them, so
they reflect light better and show up more easily. The particles in the rings can be as
small as dust specks and as large as houses.
Where did the rings come from? When in doubt, blame gravity. In this case, gravity is
guilty of the crime. It has to do with a gravitational boundary around a planet known
as the Roche Limit. Objects that are a certain distance from the planet (within the
Roche Limit) can be easily broken up if they are not made of very strong material.
Chunks of lightweight stuff like ices are very easily broken apart. These objects will
continue to break up until the particles are too small to break apart.
403 | P a g e

Why should they break up? It can be related to Kepler's 3rd law. If you don't
remember this one, I'll remind you. It says that the closer you are to an object, the
smaller your orbital period (the faster you go). If you have a big chunk of material that
gets into the Roche limit, the part of the object that is closest to the planet will want to
move faster as it orbits around, while a part that is further away will want to move
more slowly. The differences in speed will cause the chunk of material to break apart.
The material will keep breaking apart until the speed difference across the object is
not significant enough to tear it apart anymore.
Any icy object, like a moon or a comet, cannot remain as a large solid if it is too close
to the planet. This helps to explain why the massive jovian planets all have rings,
since they have large Roche limits (since they have large masses).
In 2009 the Spitzer telescope discovered a rather unusual ring around Saturn. This ring
is much further away than the others and it is tilted significantly relative to the other
rings. It is also rather spread out, stretching from 6 - 12 million km from Saturn.
Satellites
Saturn has a large number of satellites (62 as of August 2011), and there are more
objects that are suspected satellites, but they have yet to be confirmed. It is sort of
difficult to find a satellite around Saturn because the rings are so bright they tend to
outshine any objects near them. When the rings are tilted so that they appear to be
gone, astronomers will try to find new moons around Saturn. Most of the satellites are
pretty standard for the outer solar system - icy, low density, small worlds. Of course,
there are a few exceptions.
IR images from the Cassini spacecraft showing
the surface features on Titan. Image from NASA/JPL/Space Science Institute.
404 | P a g e

Titan (click on it to see an IR movie of Titan's surface based upon Hubble
Telescope infrared images). Titan is a very large satellite; it's even bigger than
Mercury. This isn't why it's important. Titan has a thick atmosphere! The atmospheric
pressure is about 1.6 times the pressure on the Earth's. Unlike Venus or Mars, the
atmosphere isn't too thick or too thin compared to the Earth's. The composition of the
atmosphere is also mainly nitrogen (again, just like the Earth). There is also other stuff
like methane and ethane just to make it annoying and stinky. The atmosphere contains
a haze layer that prevents us from easily seeing the surface. The Cassini spacecraft
has been able to give us a view of the surface though, using an infrared imaging
camera.

Figure 15. Images of Titan from Cassini. Above are three
views of Titan. The first, a visible light image, the second, an
infrared image, and the third, a false-color infrared image. In
the false-color image, the red areas actually do correspond to
layers in the atmosphere where methane absorbs sunlight, and
green areas are surface layers that are visible. To the right is
another infrared image obtained by Cassini. Due to the haze
layer on Titan, the only way to see surface features is to use
other than visible wavelengths. Images from NASA/JPL/Space
Science Institute

405 | P a g e

Is Titan really just like the Earth? No, it is much further from the Sun, so it gets pretty
cold. The temperature is close to 95 K (-290 degrees F), so that makes it rather
unpleasant for humans. The really freaky thing about Titan is that the temperature and
the pressure at its surface are such that methane can exist as a gas, solid or liquid
there, sort of the same way that water can exist on the Earth. But there has been quite
a bit of speculation about Titan as a place for "different" forms of life - those that may
use methane rather than water as a basis. This is still just guess work at this time, but
it is does keep our interest rather high for this little world.
Due the presence of the atmosphere, Titan was the target for the probe
(named Huygens) that the Cassini spacecraft dropped off. The probe floated through
the atmosphere for a few hours and then landed on the surface and transmitted data for
about 1 hour. To see a movie of what the probe "saw" as it descended, just click here -
it does get a bit dizzy as it goes down, but of course it was pretty windy there. It was
able to make a thorough analysis of the clouds, atmosphere and some surface features.
Amongst the more interesting things that have been discovered by the Cassini mission
are -
Lakes or large bodies of icy methane, ammonia, water ice or other
hydrocarbons (click here to see an image of one of these lakes compared to
Lake Superior)
"Rivers" of methane or hydrocarbons, possibly as long as 1,500 km (930 miles)
Erosion features
Large dunes - these are actually pertty common on the surface, and much larger
than Earth dunes.
Possible volcanism and tectonic activity
Craters (well, that's not very exciting..)
Wind speeds up to 430 kph (265 mph) in the upper atmosphere, down to a mild
breeze at the surface
Nearly constant drizzle, not of water, but rather liquid methane drizzle, making
the surface rather "muddy"
Methane "fog" has also been observed
No doubt there will probably be more discoveries as the Cassini mission continues.
406 | P a g e

The other moons can be divided into the usual categories of really puny and sort of in
between sized. The really small ones tend to be the ones found around the rings (they
tend to be shepherd satellites). All of the moons except for Enceladus show
considerable cratering.
Enceladus has some ice flow features and some rather unusual surface features. Apart
from Titan, it is the most interesting world around Saturn. This is due in part to the
fact that it has a large amount of resurfacing going on, just like on Europa and
Ganymede. And what could cause that? Again, the best cause would be internal
heating, similar to the heating of the Galilean satellites. Recent calculations of the
amount of tidal heating indicate that the cracks (as seen in the images below) may
actually open and close while Enceladus orbits around Saturn once every 1.3 days.
Another item to support the heating was the existence of an atmosphere around
Enceladus, something that Cassini also discovered. This was a surprise, since
Enceladus is rather small to hold onto an atmosphere. The atmosphere is made
primarily of water vapor with some nitrogen, carbon dioxide and other carbon-based
molecules mixed in.



Figure 16. Images of Titan from the Huygens probe. The left and
middle images were taken during the descent of the probe. These
show features that look very much like water features on the Earth,
though of course with the conditions on Titan, it can't be water! On
the right is a true-color picture from the surface of Titan taken by
the probe. The rocks around it aren't very big, only a few inches in
size. Images from NASA/JPL/Space Science Institute/ESA/University
of Arizona
407 | P a g e

Another recent discovery were active geysers erupting from the surface of Enceladus.
Observations of the material indicates that it is mainly water vapor, with some carbon
dioxide, carbon monoxide and some organic compounds. The presence of water vapor
indicates that there are likely pools of liquid water just under the surface of
Enceladus! In 2009 the Cassini spacecraft discovered ammonia in the vapor of
Enceldus, which provides even more evidence that water may exist as a liquid below
the surface. How does the presence of ammonia tell us about liquid water? Ammonia
acts as a natural "anti-freeze" and will allow water to stay in a liquid state even at the
extremely low temperatures on Enceladus. The eruptions from Enceladus also appear
to be the source for the moon's atmosphere as well as a major contributor to one of
Saturn's rings. A simple measurement of the surface temperature shows that the cracks
on Enceladus are significantly hotter than the surrounding areas, again, showing that
Enceladus is currently hot and active. Enceladus is located right in the middle of one
of the outer, fainter rings (the "E" ring), and if it is blowing a bunch of material off of
its surface, this material would likely end up in orbit around Saturn as part of the ring
system. The Cassini spacecraft has also detected grains of salt in the "E" ring around
Saturn, along with the regular ice particles. If this ring is produced by Enceladus'
eruptions, then the salt is another indicator of a liquid interior for the moon - the
evidence just keeps on piling up!
In 2011 it was determined that the water vapor not only contributes to the ring around
Saturn, but some of it can actually get to Saturn itself! This is an important discovery
since it shows that material can go from the satellite and get into the upper layers of
Saturn's atmosphere. It is fairly rare for there to be a transfer of material observed
from one object to another, but with so much water being blasted out (current
estimated rate of water expulsion is 250 kg/sec), some of it can travel down to Saturn.
Astronomers have also determined that the eruptions seem to be linked to the orbit of
Enceladus around Saturn, with more activity when the moon was far from Saturn. It is
thought that when it is close, the gravitation effects of Saturn tighten up the cracks and
prevent the fissures from erupting, while when it is far from Saturn, there is less
gravitational contractions and the material can erupt more freely.
408 | P a g e


Figure 17. Images of Enceladus. On the left is the view from Cassini, with Saturn in the
background. Note the rather cracked surface and the relatively few craters. A close up view is
seen in the middle showing the flow features, nicknamed the "tiger strips". On the right is a view
of an eruption of material from the surface. Images from NASA/JPL/Space Science Institute.
While Enceladus has the most dramatic geyser activity, it appears that a couple of other satellites
are also geologically active. In 2007, evidence of eruptions were linked to Tethys and Dione, two
relatively moderate sized objects. But of course with further studies we might find even more
activity in the Saturn system, courtesy of the Cassini spacecraft.
The named satellites were given names related to the Titans of mythology, but now there are so
many that some other mythologies are included, such as Norse and Inuit. Some of the names are
possibly familiar to you - Pan, Daphnis,
Atlas, Prometheus, Pandora, Epimetheus, Janus, Mimas, Methone,
Pallene, Enceladus, Tethys, Telesto, Calypso, Dione, Helene,
Polydeuces, Rhea, Titan, Hyperion, Iapetus, Kiviuq, Ijiraq, Phoebe, Paaliaq, Skathi, Albiorix,
Bebhionn, Erriapo, Siarnaq, Skoll, Tarvos, Hyrrokkin, Mundilfari, Narvi, Bergelmir, Suttungr,
Hati, Bestla, Farbauti, Thrymr, Aegir, Kari, Fenrir, Surtur, Ymir, Loge, Fornjot and about nine
more that have yet to be named. Images from NASA/JPL/Space Science Institute and Calvin J.
Hamilton.
Recently the Hubble Space Telescope was able to record a movie showing the motion of several
of moons in front of Saturn. It does look rather nifty.

Now that you've read this section, you should be able to answer these questions....
What are the cloud layers of Jupiter and Saturn made up of?
Why do they have so many storms?
What effect does a lower mass for Saturn have on its characteristics and its appearance?
What is the interior of Jupiter and Saturn like?
409 | P a g e

What produces the magnetic fields for these planets?
How are the characteristics of the galilean satellites like that of a mini-solar system?
What produces the features observed on Io?
What make Titan an interesting place to visit?
How does Enceladus influence the rings around Saturn?
What causes planetary rings?
Uranus, Neptune and Pluto
Uranus

Uranus Symbol


What's covered here:
How are Uranus and Neptune different from Jupiter and Saturn? How are they
similar?
Who discovered Uranus?
How was Neptune discovered?
What are the satellites of Uranus and Neptune like?
What are Pluto's characteristics?
What are the other objects found beyond Neptune like?

Sir William Herschel (1738-1822). Image Copyright Royal
Astronomical Society.
The first of the planets to be "discovered" was named after the
muse of Astronomy (Urania). Uranus is also the name of
Saturn's father. Uranus was discovered by William
Herschel in 1781, though it had been mistakenly mapped in
the past as a star. It is possible under the correct conditions to
see Uranus with your just eye (no telescope), though it is
410 | P a g e

pretty difficult to do, and you would need to have good eyesight and know exactly
where to look to see it. Of course, you have to pronounce the name correctly. If you
say something that sounds like "yer-anus," I don't think you'll make too many friends
that way, unless they are proctologists. You should instead say the name as "yer-ah-
nis" - not much of an improvement, but it may help.
Uranus is so far away that we can't get a great deal of information on it from this
distance. A few things we can see are its tilt and rotation rate. Uranus is pretty much
tipped over on its side. You know that the Earth is tilted over 23 1/2 degrees; Uranus
is tilted over 98 degrees! This means it is on its side (since being 90 degrees would be
exactly on its side). Due to this the seasons on Uranus are pretty extreme, especially if
you are located at one of the poles. It takes Uranus about 84 years to go around the
Sun, so the seasons last for about 21 years. If you were located at one of the poles,
you would have 21 years of constant daylight during summer and 21 years of
continual darkness during the winter. The temperatures would be pretty extreme. Not
only is the planet tilted over, but the rings and satellites are also in tilted orbits about
Uranus. It is likely that early in its history the planet was knocked over by an impact
and as the system formed (as the satellites and rings formed), they also ended up in
tilted orientations.
Figure 1. The orbit
and tilt of Uranus
are shown. Uranus
is tilted pretty
much on its side,
so there are times
when only certain
parts of it will get
any sunlight. This is especially true for the polar regions.
There are several very thin rings around Uranus. The rings were discovered from the
Earth by accident in 1977. Astronomers were trying to study the atmosphere of the
planet by watching how the light of a star would change (get absorbed by the
atmosphere) as the planet moved in front of it. Before the star was anywhere close to
the planet, the star's brightness faded several times. This happened again when the star
was located on the other side of the planet. Now it is possible that the star could have
been eclipsed by some moons, but that would have required some very precise
alignments of the moons with the distant star. Rings were the cause for the dimming
of the starlight! The rings are rather similar to Jupiter's ring in that they are made of
darker material as well as small microscopic particles - stuff like dust. Uranus actually
has 11 main rings, far fewer than Saturn.
411 | P a g e

Figure 2. The dimming of a
star's light as it approached
Uranus indicated that there are
rings around Uranus.
Viewing Uranus from the Earth
doesn't provide much
information, so let's send a
spacecraft out there. Actually,
we have - the only spacecraft to
visit Uranus was Voyager 2 in
January of 1986. The
other Voyager spacecraft did not
visit any other planets after its
encounter with Saturn,
while Voyager 2 continued on to
Uranus and Neptune. All of the up close images we have of them are from that one
spacecraft. What did it see?
When Voyager 2 flew by Uranus, there were no cloud features (spots) as had been
seen on Jupiter and Saturn. In fact, there are no markings at all; the atmosphere of
Uranus appeared to be very dull. This was really surprising, since at the time that
the Voyager spacecraft flew by Uranus, one of the poles was pointed toward the Sun.
Astronomers thought that this would have caused the atmosphere to heat up a great
deal and therefore produce a bunch of storms. It was a bit confusing when all
the Voyager images were so dull. When we look at Uranus today with the Hubble
Space telescope, there are cloud features visible. The atmosphere appears to change as
it orbits the Sun.
Figure 3. The view of Uranus as seen by the Hubble
Space Telescope. This differs from the Voyager view
from 1986, which showed no cloud features. The
clouds seen here as the pinkish features have been
observed since the mid-1990s. Image credits: Erich
Karkoschka (University of Arizona) and NASA.
What is the inside of Uranus like? (You'll want to
make sure you pronounce Uranus correctly if you ever
ask anyone that question.) The discussions of Jupiter
and Saturn showed how similar those two objects are
to one another. The same is true for Uranus and
Neptune, so most of the stuff I say here about Uranus
412 | P a g e

also applies to Neptune. From the observations of the Voyager spacecraft we know
that Uranus is similar to Jupiter and Saturn in some respects, though there are some
drastic differences. First of all, while Saturn and Jupiter are 100s of times more
massive than the Earth, Uranus (and Neptune) are only about 15 times more massive
than the Earth and their radii are only about 5 times greater than the Earth's. Their
internal structures are also different. Just check out Figure 4.
Figure 4. The
internal structure of
Uranus is shown. The
main difference
between Uranus (and
Neptune) and Jupiter
(and Saturn) is the
large liquid water
ocean that exists
under the upper
atmosphere.
Uranus is rather cool; unlike Jupiter and Saturn, there is not any excess heat coming
from its interior. Remember, Jupiter and Saturn are both hot because they haven't
cooled off. Uranus, being much smaller, has had the time to cool down completely.
When you look at the top of the atmosphere of Uranus, it has a rather nice blue-green
color. This is due to methane in the upper atmosphere. While this is only a small part
of the overall composition of these layers, it is enough to provide the color. It is sort
of interesting that most of the inside of Uranus is not made up of just regular hydrogen
but is comprised of hydrogen compounds (remember, water, methane and ammonia
all have hydrogen in them).
Is there a magnetic field? Yes, but it is very strange. The magnetic axis does not go
through the center of the planet and it is not located anywhere near the rotation axis. It
is both off center and off axis. This produces a rather unusual wobble in the magnetic
field as the planet rotates once every 16 hours. In terms of strength, the field is about
50 times stronger than the Earth's. Check out this animation to see what is happening.
413 | P a g e

Figure 5. The alignments of the Earth's and the jovian planets' magnetic fields are
shown. The top row shows each planet with no axis tilt - so that their rotation poles
are all oriented the same. The bottom row shows each planet with its normal tilt. The
magnetic axes are indicated by the red lines, while the rotation axes are shown with
green lines. There is only about a 10 degree difference between these two axes for the
Earth and Jupiter, while Saturn's axes are completely in line. Both Uranus and
Neptune have a huge difference in the alignments of their magnetic and rotation axes.
Also, note how the "magnet" doesn't go through the center of Uranus or Neptune but
is off to the side. Of course, there really are no giant magnets in the cores of planets,
this just shows the alignment of the magnetic fields with the rotation axis.
What produces the magnetic fields of Uranus and Neptune? Unlike the terrestrial
planets, there are not large amounts of iron and metal, and unlike Jupiter and Saturn
there aren't layers of liquid metallic hydrogen. Take a look at the image showing the
interior of Uranus - what material in there could conduct electricity? If you know that
you aren't supposed to have electrical appliances near water, then you should guess
that the water would be the most likely one to produce the magnetic field. It appears
that the magnetic field is produced by electrically charged water in the huge water
layer. The high pressures on the water cause it to be electrically charged.
Satellites - Like other Jovian planets, most of the satellites around Uranus fall into the
puny category (less than 300 km in diameter). They are, for the most part, dark, small,
icy worlds. The ices seen on their surfaces are in various forms including water,
ammonia, and methane ice. As has been seen on other satellites, they have quite a few
craters and there are some ice flow features. Voyager flew by when one of the poles of
Uranus was pointed toward the Sun, so the satellites where also in that same
orientation. Therefore we have pictures of only 1/2 the surfaces of these worlds.
414 | P a g e

Five satellites were known of before the Voyager 2 probe flew by. These were
Miranda, Ariel, Umbriel, Titania, and Oberon. The Voyager 2 spacecraft discovered
10 more, and they were named Cordelia, Ophelia, Bianca, Cressida, Desdemona,
Juliet, Portia, Rosalind, Puck, and Belinda. In case you haven't figured it out, the
moons are named after characters from William Shakespeare and Alexander Pope.
Twelve more moons have been discovered since the Voyager mission bringing the
total to 27. It's a good thing that Shakespeare and Pope had a lot of characters or else
we'd run out of names.
Figure 6. The surface of Miranda.
There is always one satellite that is unusual. In
the case of Uranus, the unusual satellite
is Miranda. While this is not a very large
object, only about 470 km wide, it does show
some very unusual surface features. These
include some large circular patterns on its
surface and abrupt color variations. There is
also a very large cliff visible (Verona Chasm,
5 km high). Why is it so screwy? It is possible
that after the moon formed, it was broken up
and then reformed in a jumbled mess. It is also
possible that there is a great deal of tidal
heating and this may have rearranged the features on Miranda, or at least allowed
them to be a bit mobile. It should be noted that Miranda is the closest of the major
moons to Uranus, so it would have experienced the most tidal heating.
Several of the other moons are a bit mysterious. Ariel and Umbriel both appear to be
the same size, yet Ariel's surface shows evidence of ice volcanism and tectonics,
while Umbriel's surface is just full of craters. Titania and Oberon also appear to be
very similar in size, and Titania appears to have had much more activity on its
surface. It was thought that worlds of similar size should have similar natures, yet this
small sample of worlds shows quite drastic differences - just more of those annoying
mysteries we have yet to solve.

Neptune
415 | P a g e


Neptune Symbol

The last of the jovian planets was named after the god of the oceans. It was discovered
in 1846 by people using mathematics. What? You mean math is actually good for
something? Of course it is. Any ways, here's what happened - once Uranus was
discovered, people started to observe it and plot its orbit. They noticed that its motion
wasn't exactly correct; that it deviated slightly from Kepler's and Newton's laws of
how things move. This would happen if there was another object out there pulling on
it. This led John Adams, an English astronomer, to calculate where a planet should
be based upon the unusual motions of Uranus. A little bit later,Urbain Le Verrier, a
French astronomer, did pretty much the same calculations, and his results were
published (Adams's calculations were pretty much ignored). The published results of
Le Verrier were used by Johann Galle, a German astronomer, who used the
information from Le Verrier to search for the planet. He found it (quite easily)! Who
actually discovered Neptune? Credit for the discovery of the planet is given to both
Adams and Le Verrier, since without their work Galle would have never seen it.

Left to right, John Couch Adams (1819-1892), Urbain Jean Joseph Le Verrier (1811-1877) and
Johann Gottfried Galle (1812-1910). Adams and Le Verrier are credited with the discovery of
Neptune.
416 | P a g e

It is also possible that Galileo saw it hundreds of years before Galle, since some of his
diagrams indicate an object where Neptune would have been at the location he was
observing. He would have probably mistaken it for a star.
The only spacecraft to go by Neptune was Voyager 2, which flew by in August 1989.
Neptune is quite a long ways out there and the fact that the spacecraft lasted long
enough to get there was quite amazing (it was launched in 1977). Also, the path that it
was on as it flew by Neptune was very precise. If you were to scale the flight path
down to a golf putt, the accuracy of the path was the equivalent of sinking a putt from
3630 km (2260 miles) away. That's accurate! The images from the Voyager
2 spacecraft look a little strange, in part because the spacecraft is traveling so fast and
it has to take fairly long exposures, since there is not a lot of sunlight out at that
distance. Now the spacecraft is on its way to the outer parts of our solar system, which
it won't get out of for a few thousand more years (even though it is traveling at about
38,500 mph)
While Neptune is pretty similar to Uranus in terms of its composition, structure and
density, there are some subtle differences. One area where it is quite different is in the
atmospheric features observed by Voyager 2. TheGreat Dark Spot was observed on
Neptune. This is a storm like feature about the size of the Earth. It appears to have
been similar in many respects to the Great Red Spot on Jupiter (proportionally about
the same size). Current observations by the Hubble Space Telescope indicate that the
Great Dark Spot is no longer present (I guess it wasn't so great) and other spots have
appeared. It appears that Neptune's atmosphere is much more active compared to
Uranus's.

Figure 7. To the left is an
image of the Great Dark Spot
as photographed by the
Voyager 2 spacecraft in 1989.
On the right is an image of
Neptune obtained by the
Hubble Space Telescope in
1996. The Great Dark Spot is
no longer visible, but other
spots appear that weren't seen
by Voyager. Images from NASA
and Lawrence Sromovsky
(University of Wisconsin-
Madison).

The Voyager images also showed other higher altitude clouds as well as smaller spots.
Many of these can't be seen from the Earth with the Hubble Space Telescope, but
417 | P a g e

large changes in the atmosphere of Neptune are visible. As with Uranus, the color of
Neptune is due to methane in its atmosphere, even though the main components of the
atmosphere are hydrogen and helium.
As mentioned previously, it appears that the internal structure is much like that of
Uranus, and the magnetic field is also like that of Uranus (tilted 47 degrees from the
rotation axis and off center). The strength is a bit weaker than Uranus's, though, only
about 35x the Earth's magnetic field strength. You have to remember that Neptune
itself is not tilted over like Uranus. A rather major difference between Uranus and
Neptune is that Neptune has an internal heat source (Uranus doesn't). This heat may
be left over from the planet's formation, but that seems unlikely considering that
Neptune is not that large (like Jupiter and Saturn). We are still not quite sure why
there is still so much heat in Neptune and none in Uranus, but the heat in Neptune
does help to explain the greater activity in its atmosphere, since heat helps to produce
turbulence in the atmosphere, which leads to the formation of storms and spots. One
rather surprising feature of Neptune's atmosphere was the high wind speed observed -
it can go up to 2000 km/h (1200 mph).
Like the other jovian planets, Neptune has rings. This wasn't known for certain
until Voyager 2 went out there. Neptune isn't tilted like Uranus, so the rings could not
be seen in the same way they were around Uranus. However, when astronomers did
try to find rings using the variation of star light around Neptune, they tended to get
conflicting results. Sometimes it looked rings were there; other times it looked as if
there were no rings. This inconsistent result is due to the presence of not only rings
but also ring arcs, which are thicker parts of rings. The ring arcs would block light
very well, so they would at times give the indication that there were rings, but since
they didn't extend all of the way around, the expected ring effects (dimming of
background stars) weren't seen everywhere. Like the rings of Uranus, the rings of
Neptune are made up of dark material.


418 | P a g e

Figure 8. The rings of Neptune are seen above in a
very long exposure image. The rings are made up of
dark material, so it is easier to see them in over
exposed pictures - but you have to remove Neptune
from the pictures. The image on the right shows an
up close view with the ring arcs clearly visible.
Image from NASA.
Neptune has the fewest number of satellites amongst the jovian planets. Before
the Voyager 2 spacecraft visited, we knew of two satellites Triton, and
Nereid. Voyager 2 discovered six new ones, which are named Naiad, Galatea,
Thalassa, Larissa, Proteus, and Despina. Since then, five more were discovered but
have yet to be named. Either way, this gives us a grand total of 13 satellites. The
theme here is to name them after water deities (which fits in well with the whole
ocean god thing). As usual, most are pretty dull - just small, icy worlds with lots of
craters. As usual, there is an exception - one strange world.

Figure 9. Triton, the
largest of Neptune's
satellites, is shown (left).
Note the strange color of
its surface features. An up
close view of the surface
is seen on the right. Note
the dark streaky parts that
are thought to be due to
ice geysers. Image from
NASA.

The largest of the Neptunian satellites is Triton. Not only is it large, but it doesn't act
normally. It is in a rather bizarre orbit, since it actually goes clockwise (rather than
counterclockwise) around Neptune, which is usually only seen in satellites that are
small, puny and thought to be captured. To have a large satellite traveling in
the wrong direction is a bit of a mystery. The really freaky thing about Triton, though,
is its surface. There are not very many large craters and there does appear to be
evidence of resurfacing. One of the more interesting items observed by Voyager
2 were cracks and streaks in the ice, possibly produced by ice volcanoes - ice
volcanoes?! It seems that there are active volcanoes (or perhaps we should
say geysers) that are erupting mainly nitrogen and methane gas and also methane ice
onto the surface. The very thin atmosphere of Triton causes these particles to drift
down wind and leave long, streaky marks on the surface. Triton's thin atmosphere is
419 | P a g e

composed mainly of nitrogen and methane. Triton also has the coldest surface
temperature yet measured on a solar system object, only 37 K (-393 degrees F).
What is Triton's problem? Why does it have activity? What do you need to have such
action? HEAT! For the most part, little worlds shouldn't be hot unless something is
heating them. It is thought that the heating could be due to the tidal forces exerted by
Neptune. Triton's orbit is not only backwards but also rather elliptical. This means that
it gets pretty close to Neptune and gets pulled pretty strongly by Neptune's tidal
forces. This can cause some major internal problems in poor little Triton. Of course,
when it gets far from Neptune, the heating is diminished so that it freezes up its
surface again until the next heating episode. As with many things, this is just a theory.

Beyond Neptune
In 1930 an object was discovered beyond Neptune and was named Pluto. In
1951, Gerard Kuiper proposed that there was a region beyond Neptune where
objects that could become comets may originate from (we'll get to comets in the next
set of notes, so don't get all excited just yet). In 1992 the first object to fit Kuiper's
theory was found in what is now called the Kuiper Belt. Some could argue that Pluto
was the first Kuiper belt object, but initial findings found that the objects discovered
there were much smaller than Pluto, so Pluto was still considered a planet. To date,
about 1000 of these objects (sometimes also refered to as Trans-Neptunian Objects)
have been found. It is thought that there may be millions of these objects out there, but
most are pretty small.
The Kuiper Belt is thought to extend from about 30 AU (the distance of Neptune) out
to several hundred AUs. Most KBOs found so far are within the 30-40 AU range, and
appear to be rather similar to the satellites of the outer solar system. The orbits of
KBOs is not at circular as that of the planets and the orbits tend to be a bit more tilted
relative to the ecliptic.
Clyde Tombaugh (1906-1997) with the telescope he built.
This was before he was hired to look for a planet beyond
Neptune.
There are some very famous KBOs, and of course the most
well known one is Pluto, which was named after the god of
the Underworld. Pluto is so far from the Sun that it takes
about 249 years to complete one orbit. It was discovered in
1930 by Clyde Tombaugh. At the time people thought that
420 | P a g e

Neptune's orbit was being altered by another planet and Clyde Tombaugh was hired
by Lowell observatory to search for the unseen planet. As you'll see, Pluto is too small
to alter Neptune's orbit, and the unusual motions have been explained without the
need for another planet. Pluto is so far from us that it wasn't until 1978 that its satellite
was discovered. The satellite is named after the ferryman of the underworld, Charon.
And in 2005 two more very small satellites for Pluto were discovered, while another
satellite was discovered in 2011, so this is now a rather busy little system. For the
story of Pluto's discovery, follow this link.
Not much is known about these Pluto and Charon due to their great distance. A
spacecraft is currently heading for them, but until it gets there our information about
them is limited by our telescopes. We do know that the Plutonian system is tilted over
like the Uranian system - Pluto is on its side and its satellites are in a tilted orbit. This
is actually a good thing, since recently Pluto and Charon were orientated in such a
way that they would eclipse each other. Why is that important? The sizes (radii) of the
objects can be learned from eclipses, and using Kepler's laws, we can get their masses.
This information can give clues as to the densities and likely compositions of not only
Pluto and Charon, but also of other Kuiper Belt Objects. We know that Pluto and
Charon are very similar in terms of their sizes. Pluto is only eight times the mass of
Charon (for comparison, the Earth is 80 times the mass of the Moon). Pluto is also
only about two times wider than Charon. Some people have referred to this system as
a binary system, since the two objects are relatively close in size, though with the
other satellites, this isn't really the case. They are rather different in density, though.
Charon's density is a bit lower than Pluto's, indicating that perhaps it formed in a
similar manner to how our Moon formed, as a big impact event. Astronomers using
the Hubble Space Telescope were also able to discern rough surface maps
of Pluto which showed variations in surface markings from previous maps. In 1989
Pluto was as close to us as it
gets for the next 249 years,
and even then it isn't very
close.
Figure 10. Hubble Space
Telescope image of Pluto and
its family of satellites. P4 was
only recently discovered and
has yet to get a formal name.
Image credit: NASA, ESA,
and M. Showalter (SETI
Institute)
421 | P a g e

A rather nifty aspect of Pluto-Charon is that both objects are tidally locked - one side
of Pluto faces Charon and one side of Charon faces Pluto. This also has the rather
nifty aspect that the time for a month on Pluto (the time for the satellite to orbit) is
equal to a day on Pluto (the time for the planet to rotate). This is about 6.4 Earth days.
If you were to stand on the surface of Pluto, the view of Charon would be great - it
would be about 10 times wider in appearance than our Moon - very romantic! Odds
are the other satellites are also locked to Pluto, but since they are so small it is difficult
to determine if that is the case. For an artist's rendition of the view from Pluto, just
follow this link. You'll see Charon appearing like a large object in the sky, and the
Sun looking rather small.
The densities of Pluto and Charon are similar to the ice/rock mixture that is seen in the
moons of the outer planets. Pluto has a very thin atmosphere made up mainly of
nitrogen and methane. The atmosphere is only present around the time of perihelion.
This is when it is close enough to the Sun that some of the surface ices can melt and
become gaseous. Most of the time, Pluto doesn't have an atmosphere. And it is also
pretty cold, on average around 40 K.
So far no spacecraft have visited Pluto, but one is one the way. The New Horizons
Pluto-Kuiper Belt Mission (I hope they give it a better, shorter name soon - that's too
much of a mouthful) was launched in January of 2006 and if all goes well, it will
arrive at Pluto by 2015.
Currently, the largest Kuiper Belt object known is Eris (formerly known as 2003
UB
313
), which is about 2600 km wide. Pluto, in comparison, is about 2,300 km wide -
large, but not larger. Eris is really the one that got Pluto kicked out of the planet club.
Not only does Eris have a larger diameter, but observations of its moon reveal that its
mass is about 27% greater than Pluto. So if Pluto is a planet, some people thought,
then Eris should also be a planet. But where does one draw the line? That's why there
was all of that debate in the summer of 2006 when the definition of of a planet came
about. So if you are upset that Pluto is no longer officially a planet, blame it on Eris -
who is appropriately named after the goddess of discord and rivalry.
Other very large KBOs have been discovered and given fun names like Quaoar,
and Orcus (formerly known as 2004 DW). Observations of the surface of Quaoar
indicate that the surface is icy, but that the ice is in a form that was created due to a
heating and freezing mechanism. It is possible that something like an impact may
have caused the heating, or perhaps some tidal forces, or maybe....who knows? The
presence of this form of ice also indicates that whatever formed it happened fairly
recently (only a few million years ago), so that's sort of exciting.
422 | P a g e

In 2004 the discovery of an object called Sedna was announced. This is an object that
is located beyond the Kuiper Belt, in a region that's called the "inner" Oort Cloud (the
animation shows a zoom-out from the inner solar system, to Sedna's orbit and finally
to the inner Oort Cloud). You'll learn more about the Oort cloud in the next section, so
stay tuned for that. Sedna's orbit is much more elliptical than a Kuiper Belt object and
it is located well beyond the edge of the Kuiper Belt. It is larger than Quaoar though.
But there are even more distant objects that we've seen. The one with the largest orbit
is 2006 SQ
372
(gotta love the name), which has an aphelion distance of 1570 AU,
while it only gets about 24 AU from the Sun at perihelion. 2006 SQ
372
is a relatively
small object, only 60 or so km in size.
Figure 11. To scale view of various objects, including the largest Kuiper Belt Object
(Eris), as well as other objects in the outer solar system, along with the Earth for
comparison. Only the image the Earth is accurate, while the others are artist's
renditions. The known satellites of these objects are also shown.
423 | P a g e

So are Kuiper Belt Objects planets? If they are, or if some of them are, where do you
draw the line? Pluto was considered a planet for more than 60 years and some still
consider it a planet, but if you include it, then you would have to include a whole
bunch of other KBOs. If you look back at the characteristics of Terrrestrial and Jovian
planets you'd see that none of the KBOs fit into that classification scheme. Sort of as a
compromise, the category of dwarf planet was invented. These are objects that meet
the first two criteria of "planet-hood" - they orbit the Sun and are large enough to be
fairly spherical. In this group we have Pluto, Eris, two other KBOs (Haumea and
Makemake), and the asteroid Ceres. So you could say we have 8+5=13 planets which
can be further broken down into "regular" and "dwarf" groups, or you could just say
we have 8 planets and a bunch of small stuff. Since the two types of planets,
Terrestrials and Jovians, only work with 8 of these, perhaps that's the best way to
distinguish those in the planet club from those not in the club.
Another name you may hear is plutoid. That refers to a dwarf planet that is located
beyond Neptune, which means Eris, Makemake, Haumea and Pluto are in that group.
Sort of silly, but I guess it will make the people who want to return Pluto to "full
planet status" sort of happy, or maybe not.
In the next (and last) section of notes you'll learn about the rest of the stuff in the solar
system. This is mainly small stuff, but as you'll see, it can have a big impact.

Now that you've read this section, you should be able to answer these questions....
What are the season on Uranus like?
What is unusual about Uranus' and Neptune's magnetic fields?
How are the internal structures of Uranus and Neptune different from Jupiter
and Saturn?
How do the ring systems of Uranus and Neptune compare to that of Jupiter and
Saturn?
What is unusual about Triton?
What is a possible cause for Triton's unusual behavior?
What characteristics of Pluto make it unlike the planets in the outer solar
system?
What are the Kuiper Belt Objects?
Comets, Asteroids, Meteorites and Impacts

What's covered here:
424 | P a g e

What exactly are comets?
Where do comets come from?
What are asteroids?
Why aren't they planets?
What types of asteroids are there?
What's the difference between meteor, meteoroid and meteorite?
What causes "meteor showers"?
What types of meteorites are there?
What sort of objects have hit the Earth?
How much damage can these objects do?
When is the next chance for an Earth impact?

Finally, we come to the last bit of detail involving the solar system - the junk. This is
really just all of the little bits that can't be put into the categories of planet or satellite.
We'll start off with the very spectacular comets, then the elusive asteroids, followed
by the meteoroids-meteors-meteorites and end up with the problems that occur when
worlds collide. Get ready for a bumpy ride!
Comets
Image of comet Hale-Bopp taken by Dr. Morgan, 1997.
Comets are quite different from planets in how they move
around the solar system. While the planets tend to have
fairly circular orbits, the orbits of comets are very
elliptical, so that they are stretched out from near the Sun
to the very edges of our solar system. Some of the orbits
aren't even elliptical - the comets just do one passage in
toward the solar system and then are gone forever. The
orbits are also rather randomly oriented to the ecliptic;
they could come in toward the Sun at pretty much any
angle relative to the ecliptic. We know that comets are
composed primarily of many varieties of ice, including
water, carbon dioxide, methane and ammonia ice. There is also a bit of dirt mixed in,
usually in the form of carbon. This makes them appear asdirty snowballs, which is
actually the name for the model that is proposed for their compositions. The best way
to think of a comet is that it is like a big chunk of ice, dirt and slush that gets stuck to
the wheel well on your car during the winter.
425 | P a g e

Figure 1. The orbital paths of comets
are very elongated (elliptical) and
randomly oriented to the ecliptic.
Comets may appear as huge objects
in the sky, but they are typically only
about 10 km in diameter, much
smaller than many other objects
going around the Sun.
This core or nuclei is how most
comets appear when located in the
outer solar system (beyond Jupiter's
orbit). However, they don't always
remain as small, dirty, frozen
icebergs. Comets change as they
orbit around the solar system,
especially when they get into the inner solar system. As a comet moves closer to the
Sun, the heat from the Sun will start to evaporate the ices that make up the core of the
comet. The material is then in a gaseous state and will form around the core of the
comet as a coma, or head of the comet. As the comet gets closer to the Sun, the gas
starts getting blown off by the solar wind. Not only is the gas blown off, but also the
heavier, dusty material gets blown away. Due to the motion of the comet, which is
pretty fast, and the force exerted by the solar winds, the trail that this evaporated
material leaves can grow quite large and will develop into tails. The coma can be
thousands of times (or more) larger than the cometary nuclei, while the tails can be up
to 1 A. U. in size (remember, 1 A. U. is about 100 million miles!).
426 | P a g e




Figure 2. To the left is Comet Ikeya-Zhang, which passed by in 2002. The two tails are not
distinct in this image; you are seeing the gas tail for the most part. I should also mention that
each image is separated by one day, indicating how fast the comet moves from day to day. Notice
also how the tail changes its appearance from one night to the next. This is mainly due to
changes in the gas tail; the dust tail stays fairly consistent in its shape. On the right is an image
of Comet Hale-Bopp, taken in 1997. The color differences of the two tails are obvious. Image
copyright 2002 Korado Korlevic et al.
Two tails are usually seen. These include the gas tail (also called the ion tail), which
is made up of material that is blown straight back by the solar wind. This is generally
made of the really lightweight gases. Within the gas tail you find stuff such as water
vapor, CO, CO
2
, N
2
, ammonia and methane gases and particles. The gas tail has a
rather ragged appearance and is sometimes rather bluish. It is always pointed directly
away from the Sun. The other tail, the dust tail, is made up of heavier particles and is
not as greatly affected by the solar wind. It has a very fuzzy appearance, often looking
rather yellow-ish or whitish. This is, of course, made up of mainly dust (rocks and
silicates). This is much heavier material, so it is not pushed into a straight line like the
gas tail but often has a curved shape that is sort of symbolic of comets. Both tails get
longer as the comet gets closer to the Sun. Actually, the tails start developing when
the comet is still quite a ways from the Sun, well beyond the orbit of the Earth.
Another thing about comets that people are confused by is how fast they move.
Comets orbit the Sun, so they obey Kepler's laws, just like the planets. Over the
course of an evening it is possible to see tiny motions of a comet relative to the stars,
especially if it is close to perihelion, but they don't go streaking across the sky as is
often portrayed in cartoons. Think of it this way - you know that the Moon moves
427 | P a g e

relative to the background stars, right? Can you actually see its motion if you sit there
looking at the Moon for some time? No, because from our view point it looks like it is
hardly moving along, but if you look at it the next evening, you'll see it is in a
different location relative to the stars. The same is true for comets; their motions may
be apparent from one night to the next, but to see motion with your eyes over the
course of a few minutes during one evening would be difficult, if not impossible.
A comet is made up of material that gets evaporated easily by the Sun, so comets lose
mass with each passage around the Sun. As much as one percent of their masses can
be blown away. Comets may start out very icy, but this is not how they'll look for
long, since the ice is the first thing to go. After a while, their nuclei will look very
dark and dirty, since the dark, dirty material (mainly carbon) will not get blown away
as easily. Comets that can't withstand the strong solar winds can also shatter apart.
This has been observed recently in the case of Comet LINEAR (C/1999 S4),
and Comet Schwassmann-Wachmann 3, both of which broke apart into smaller
pieces. Other comets that pass very close to the Sun can either completely disintegrate
or actually hit the Sun!
428 | P a g e


Figure 3. Image
from
the SOHO satelli
te showing two
comets near the
Sun. The Sun is
not visible in this
image, but is
blocked by the
central disk in
the image, but
the size of the
Sun is indicated
by the white
circle. The two
comets are being
destroyed by the
strong light and
heat from the
Sun. To date, the
SOHO satellite
has discovered
more than 1000
comets since
1996. Many, like
those shown
here, don't
always make it
past the Sun, but
are destroyed by
their close
passage. To see
a movie of
comets passing
near the Sun,
just click here.
Images and
movies courtesy
of
SOHO/LASCO
consortium.
SOHO is a
project of
international
cooperation
between ESA
429 | P a g e

and NASA
Where do comets come from? Comets were originally thought to come from the Oort
Cloud, a spherical region that extends about 50,000 AU from the Sun where the cores
of comets reside. Every once in a while, the comets are perturbed by a passing star or
collision with other comets and some fall in toward the solar system. These comets
tend to have orbits that are very elongated. In some cases the comet will only pass
near the Sun once. In other cases the comet will have its path altered, usually by going
too close to Jupiter, and it will become trapped into a shorter period around the Sun.
The existence of the Oort Cloud is based on the characteristics of comets. Many
comets tend not to be aligned with the ecliptic, so it makes sense that they originate
from a place that surrounds the solar system in all directions. Also, the very long
period (greater than 1000 years) comets have paths that stretch out to such great
distances that it is logical that they originated at great distances from the Sun. It is
estimated that there are about 100 billion comet cores out in the Oort cloud.
There is also evidence that many of the short period comets do not come from the
Oort Cloud but from a closer reservoir of cometary material that was previously
mentioned (in the last set of notes), the Kuiper Belt. This region is much closer than
the Oort Cloud, extending from about 40 AU out to a few hundred AU at most. The
characteristics of Kuiper Belt Objects can help explain the shorter period comets (less
than 1000 years), especially those with orbits that tend to be closer to the ecliptic.
Remember, if the distance from the Sun is smaller, the time for an orbit is smaller -
Kepler's Third Law is action.
Figure 4. The two
comet sources are
shown. First is the Oort
cloud, located much
further from the solar
system, and the second
is the Kuiper belt,
located just beyond the
orbit of Neptune. Image
from Calvin J.
Hamilton.
It seems that there are
really two types of comets, with the difference in them evident in the periods of the
orbits. Those with long periods (thousands to millions of years) and large orbits (1000
to 30,000 AU) also tend to have very elliptical orbits and to have orbits randomly
430 | P a g e

oriented relative to the plane of the ecliptic. These are likely from the Oort Cloud.
Those with shorter periods (a few years to a few hundred years) and smaller orbits (a
few AU to 50 AU) could have originated in the Kuiper Belt.
Throughout history comets have scared, inspired or awed people around the world.
There are some comets that are well known for various reasons. Also, comets are a
popular target for amateur astronomers, since if you find one, it is named after you.
This is one of the few things in astronomy that you can have named after you and
many amateur astronomers have been lucky enough to find a few. However, this
would require years of work, and even if you search for years you might not discover
one.
Comet Halley was named after Edmund Halley, who first determined that comets are
objects that return to the inner solar system on periodic bases and are therefore
predictable. He did not discover the comet that is named after him, but because he
"demystified" comets, the comet whose orbit he determined and whose return he
predicted is named after him. The comet came by the inner part of the solar system in
1986 and won't come back again until 2061. Its passage in 1986 wasn't very
spectacular, and the 2061 trip will probably not be very good either. This comet has
gone around the Sun quite a few times, so it has been getting smaller and less
spectacular with each passage.

Figure
5. To the
left is Sir
Edmund
Halley,
the
person
who
determin
ed that
comets
are
objects
that orbit
about the
Sun, just
like
planets.
To the
right is
an image
of the

431 | P a g e

comet
that
bears his
name as
seen
during its
1986
passage
through
the inner
solar
system.
AAO/RO
E, photo
by David
Malin.


Figure 6. Halley's comet seen throughout history. The
image on the upper left is from a tapestry showing the
comet. This is a portent of doom for the king, in this case
Harold of England, who was to be killed later that year
(1066) in the Norman conquest of England. On the upper
right is an image of the fresco by the Italian painter Giotto
with a comet shown over the adoration of the Magi scene.
It is not known for certain whether this was Halley's
comet, but it is likely that Giotto did see the comet during
his life. To the right is an image of the core of the comet
from the Giotto spacecraft. Comet image from
NASA/ESA.

Due to its fame, Halley's comet was the target of several spacecraft during its passage
through this part of the solar system in 1986-1987. One of the spacecraft was
named Giotto after the Renaissance painter who incorporated a comet in one of his
frescoes. Giotto flew close to the core of the comet, only about 600 km from it. An
432 | P a g e

animation of the approach can be seen here. At that distance it was able to obtain
many images of the comet nucleus, showing it to be a very dark object. This makes
sense, because humans have observed Halley's comet for more than 2000 years, and
with each passage by the Sun it looses more and more of its ice, leaving behind the
dark dirt. Other features observed on the comet were strong venting events, where
material was evaporated in a rather explosive manner. These are seen as the bright
features in the Giotto image shown above. The power of such vents is pretty great,
enough to even alter the comet's orbital path slightly.



Figure 7. Comet Hale-Bopp seen from the roof of McCollum Science Hall (the observatory is
seen in the foreground) and from near Hillside Observatory. In this image, the UNI-Dome is
visible in the lower right. The image at right is from a location outside of the city and shows
more of the comet's tails.
Comet Hale-Bopp passed by during the Spring of 1997 and came within 1.315 A.U.
of the Earth. Even though this is not a really close passage, it was very bright due to
its unusually large size, with a core of about 40 km wide. Hale-Bopp is easily the most
photographed comet in history. Millions of people observed it since it was easily
visible to the naked eye for several months. It was also the third brightest in recorded
history. If you missed it, then you must have had your eyes closed, since this was one
of the most prominent features in the sky during the spring of 1997.
433 | P a g e

Figure 8. Comet McNaught, as seen in early 2007. Image courtesy of S. Deiries/ESO
Another recent spectacular comet was McNaught, which passed by in the winter of
2006-2007. Actually the best view was from the southern hemisphere, and boy was
that a view. The tail of McNaught was much longer than other comet tails, which led
many astronomers to speculate that it came from a very large comet nucleus, perhaps
over 200 km in size. Unfortunately McNaught is one of those "one-time" comets
which will not return again.
The way that we currently study comets is changing quite a bit. Previously we just
were able to only sit and watch them pass by. But recently we've sort of gotten "in
their face". The first close encounters were with Comet Halley, when the Giotto space
craft flew close to the core. This was followed in January 2004, when
the Stardust flew very close to the core of comet Wild 2. During the close encounter
with the core, it was able to obtain a sample of cometary material. The spacecraft then
flew past the Earth in 2006 and successfully dropped off the collected samples.
Currently astronomers are analyzing the material, which is microscopic, to determine
the conditions under which comets formed. In general we had assumed that they
434 | P a g e

formed out of only "cold" material, but scientists were surprised to discover several
different minerals that only form under conditions of very high temperatures! Also,
some icy particles were found that form only in condistion with very low
temperatures. How is that possible? It would appear that early in the history of our
solar system there was mixing of material from the inner to the outer solar system, so
that some of the high temperature material was incorporated into
comets. Stardust continued on its journey after dropping off the sample from Wild 2,
and visited the comet Tempel 1 (see below) to check on the "damage" caused by
another spacecraft. They actually renamed the spacecraft Stardust-NExT since this
was a "new" mission. After the encounter with Tempel 1, the spacecraft was sent off
into an orbit that would prevent it from hitting the Earth.
The most dramatic comet encounter occured in July 2005 when the Deep
Impact spacecraft sent a probe slamming into the surface of comet Tempel 1.
Astronomers wanted to break "into" the comet to see what sort of material was below
the surface, since the surface of a comet is altered by constant exposure to the Sun and
radiation from other sources. Is the ice under the "crust" different? How thick is the
surface layer? Results from the impact are still being studied, but there have already
been some surprises, such as the how the dust on the surface of the comet is very fine
and powdery, something that wasn't expected. The material also seems to be rich in
organic matter, material that is rich in carbon. There isn't a great deal of water ice
visible on the comet's surface, which means it looks more like a dirt ball with some
ice on it, rather that a dirty snowball. It is thought that there is still quite a bit of ice
inside of the comet, below the surface. The impact caused the comet to brighten
slightly, but that didn't last very long. And like the Stardust spacecraft was re-used to
visit Tempel 1, the Deep Impact spacecraft also was sent on a journey to another
comet, in this case Hartley 2. To see the view from the spacecraft as it flew by the
comet, just follow this link.
435 | P a g e



Figure 9. Image on the top left shows comet Tempel 1 before the impact of the Deep Impact probe. The image in the top right
is the view approximately 16 seconds after the impact. The image on the bottom shows the area of the damage before the
436 | P a g e

impact (2005) and the same area as seen in 2011 by the Stardust-NExT spacecraft. The probe hit the comet at a speed of about
10 km/s (6.3 miles per second). You can view a little movie of the impact here and here - one movie is from images taken by
the probe as it heads in, the other from the rest of the spacecraft that flew past the comet. Image credits: NASA/JPL-
Caltech/UMD/Cornell
Many people speculate that comets may have been a source of much of the water on
the Earth or even the source for life on the Earth, since very complex molecules have
been found in space. As with so many other aspects of astronomy, expect our
understanding of comets to change in a few years.

Asteroids
Figure 10. A view of an asteroid (named Braille) from the
Earth. The asteroid appears as a short streak on the image,
while the stars appear to be fixed in their locations in this long
exposure, since asteroids move enough so that their motions
reveal them to be solar system objects orbiting the Sun. Click on
the image to see how far the asteroid moves in half an hour.
Asteroids are quite different from comets. Even without looking
at them in great detail there are some obvious differences. One thing about asteroids is
that they weren't discovered until the invention of the telescope. The first asteroid to
be discovered wasn't found until 1801 and was named Ceres. This is also the largest
of the asteroids, about 940 km wide. The discovery of more and more asteroids over
the years has us believing that there are likely billions of asteroids out there but only a
couple hundred larger than 100 km. The vast majority are so small you could easily
hold one in your hand. The number of known asteroids changes every month with
new discoveries, and there are currently more than 100,000 known asteroids. Unlike
comets, asteroids are found relatively close to the ecliptic plane and their orbits are
not as stretched out as comets' but are more circular. This is why asteroids are also
called Minor Planets. The vast majority of asteroids are found in the asteroid
belt between the orbits of Mars and Jupiter. This gives them a typical distance from
the Sun of about 2-3.5 A.U. To see the latest plot of asteroids, just follow this link -
more than 100,000 objects are plotted.
You may also notice that there are some asteroids located along the orbit of Jupiter,
both in front of and behind Jupiter. These are the Trojan asteroids. They are in those
locations because of Jupiter's strong pull on them. They orbit the Sun with the same
orbital period as Jupiter, so they are always located in the same positions relative to
Jupiter, either always before or always after Jupiter. Jupiter has the most trojan
437 | P a g e

asteroids, but Mars, and Neptune also have them, but only a few. In 2011, the first
trojan asteroid of the Earth was announced (it has the lovely name 2010 TK7). It is
trapped in a location that causes it to always remain in front of the Earth as both travel
around the Sun.
There are also asteroids that are found closer to the Sun, and included in this group are
the N.E.A. - Near Earth Asteroids. These are the ones we want to pay close attention
to since they can cross the Earth's orbit - sometimes they get quite close. Some of
these have already been labeled as PHA- Potentially Hazardous Asteroids. Don't
astronomers have such fun abbreviations?

Figure 11. A variety of asteroids. On the far left is Ida, an asteroid that has a satellite - Dactyl is
the moon's name. Next is Gaspra, the first up close asteroid image we ever obtained. The third
image is the very dark asteroid Mathilde. And the right most is Itokawa. Gaspra is an s-type
asteroid; Ida's composition is likely an s-type, Mathilde is a c-type asteroid and Itokawa is an s-
type. Gaspra and Ida were photographed by the Galileo spacecraft on its way to Jupiter, while
Mathilde was photographed by the NEAR spacecraft and Itokawa was imaged by the Japanese
spacecraft Hayabusa. Click on an image to see an animation of the asteroid. Images from NASA,
animations A. Tayfun Oner., ISAS / JAXA
Studies of asteroids have been spotty at best. As you'll see, one of the best ways to
study them is after they fall to the Earth and are picked up as meteorites, but I'll get to
that later. Only on a few occasions have we observed asteroids in their native
environment. The Galileo spacecraft obtained images of two asteroids, Gaspra and
Ida, on its journey to Jupiter. At rare times, N.E.A.s get close enough to Earth for
astronomers to map them out with radar, which provides information on their shapes
and surfaces. On February 14, 2000 the NEAR spacecraft went into orbit about the
asteroid Eros. It thoroughly mapped and scanned the surface of Eros from that time
until February 2001. At the end of the mission, the NASA folks thought it would be
cool to see if they could land the spacecraft on the surface of the asteroid - and they
did! The darn thing landed fairly safely on the surface, even though it wasn't designed
438 | P a g e

to land. By the way, NEAR stands for Near Earth Asteroid Rendezvous - another one
of those cute astronomy abbreviations.
The japanese spacecraft, Hayabusa was able to land on, and take off from an asteroid
named Itokawa with a sample of material from the surface. The Dawn spacecraft is
currently (Summer 2011) in the asteroid belt in orbit about the large asteroid Vesta.
There are likely going to be more details about asteroids known in the future as results
from these missions are made
available.
Figure 12. The distribution of
asteroids from the Sun shows
prominent gaps, called Kirkwood
gaps. These are caused by Jupiter's
gravitational pull.
After data of the orbits of many
asteroids were collected, some
rather interesting aspects were seen
in their distributions in the asteroid
belt. There are a bunch of locations
where few asteroids are found.
These gaps were first noted by a
fellow named Kirkwood, so they
are called the Kirkwood gaps.
Why are they there? The gaps are
seen at places in the asteroid belt
where asteroids have orbital periods that are a fraction of Jupiter's, like 1/2, 1/3, 1/4
and so on. When an asteroid has an orbital period that is one of these fractions of
Jupiter's, that means that the asteroid and Jupiter will be lined up on a regular basis
and Jupiter will be able to pull the asteroid a bit more than normal. Jupiter tends to
pull such asteroids out of these resonant orbits and produce the gaps.
Astronomers have only recently been able to get up close views of asteroids, but even
without that data we have an idea about what they are made of (or at least what their
surfaces are made of) based upon how light reflects off of them and by looking at their
spectra. It appears that the compositions of asteroids vary with their distances from
the Sun. Those near the orbit of Mars are S-Types (silicate rich, rocky objects), while
those further out (about 2.7 A. U. and beyond) are designated as C-Types (carbon
rich). There are also some M-Types (metallic composition) mixed in with the S-types.
The general breakdown of the various types is that about 75% are C-types, 15% are S-
Types, and most of the remaining 10% are M-types. It is possible that the really
439 | P a g e

distant asteroids have quite a bit of ice in them. You'll notice how the asteroid
compositions also vary with distance from the Sun and work in quite well with the
density variations of the planets - cool, eh? As mentioned, asteroids are very small in
size. When an object is very small, it tends to not be circular in shape. Why is that?
High gravity tends to make things circular so that all parts have an equal pull inwards.
If you don't have much mass to work with, then the gravity isn't really strong enough
to pull the stuff inwards, so small asteroids are generally not spherical in shape but
tend to be potato shaped, just like the satellites of Mars and the small satellites of the
jovian planets.
Why are there asteroids? For some time people thought that the asteroid belt is
actually the remains of a planet that got broken apart. This is probably not the case.
First of all, if you were to add all the masses of the asteroid belt together, you'd end up
with an object that is only a fraction of the size of the Moon! More likely, this is just
material that was never able to form into a planet or be incorporated into any other
objects. One of the reasons for this is Jupiter. Jupiter's strong gravitational pull would
tend to prevent any large object forming in this area, sort of for the same reasons that
there are Kirkwood gaps.
If you saw the movie The Empire Strikes Back (one of the early Star Wars movies),
you might remember a scene where they are zipping through an asteroid field. This is
totally unlike our asteroid belt. Typical distances between large asteroids are on the
order of millions of kilometers, not just a few feet. At such great distances the odds of
a collision between asteroids is rather low - though not impossible. We know that
asteroids have collided in the past (due to some suspicious shapes) and will collide in
the future; it's just not a very common event. Some asteroids are actually not a single
object but are in several pieces. There are even asteroids that have their own satellites,
probably piece of the asteroid that broke off in the past, or flew off due to high
rotation rates. In 2009 an asteroid (named 1994 CC) was discovered to have two
satellites! However, when an asteroid does collide with another, it is possible that
parts of the asteroid can be sent out of the asteroid belt and directed toward us. This
could also happen if one of the PHAs gets a little too close to the Earth. In that case,
we get into the domain of meteors and meteorites.
What do you think about the existence of swarms of rocky material (asteroid belt),
and icy material (Kuiper belt and Oort Cloud) around the Sun? Is it normal for a star
to have such variety? Possibly. It is very difficult to see planets around other stars,
much less asteroids or comets, but the presence of a belt may be observable.
The Spitzer telescope has detected a cometary material belt (like the Kuiper belt), as
well as two asteroid belts around the nearby star Epsilon Eridanii. So I guess we're not
that unusual after all.
440 | P a g e


Meteoroids, Meteors, Meteorites
What's the difference between these things? Here's a quick definition of the three
Meteoroid - a small object in space. This can be in many different forms and
could be a comet, an asteroid, or even something that is artificial (man-made).
It is just a general term for anything small that is out there.
Meteor - when an object from space enters the Earth's atmosphere and burns
up. It usually leaves a trail of "smoke" in the sky and appears as a bright streak.
This is a meteor - the burning up of a meteoroid. The phrase meteor
shower comes when a lot of these happen in a short time span. A meteor is also
known as a shooting star, though, of course, it has nothing to do with stars. As
in the case of a meteoroid, the origin of the object (what it originally was) is not
used to categorize it, since it is not always possible to determine what it was
made of or where it originally came from. A meteor could be a piece of a
comet, asteroid, spacecraft, etc. that is falling through the atmosphere.
Meteorite - an object that actually makes it to the Earth. Again, the origins of
such objects can vary, though it is generally easier to have things that are rather
rocky or metallic end up all the way down at the surface. Icy objects usually
don't make it.
We already went over the stuff that would be in the meteoroid category, we'll go to
the next one - meteors.
Figure 13. A meteor shower is
shown in a 15 minute
exposure. Some of the
constellations are labeled in
this view, which shows many
stars (smallish streaks) which
all appear to be moving in a
uniform direction, while the
meteors (longer streaks)
appear to be radiating from
the constellation of Leo (if you
were to trace their motion
back). This is actually a picture from the 1998 Leonid meteor shower, which makes
sense, because that's where it appears the meteors are coming from. Image copyright
Lorenzo Lovato.
441 | P a g e

It surprises many people that meteor showers are not due to asteroids but are instead
caused by comets. Comets are objects that travel through the inner solar system, so
some of their paths intersect the Earth's orbital path. The chance of a collision is not
that great, but because the trail of debris left by the comet's passage is rather large, the
Earth does pass through that. As the debris is swept up by the Earth and passes
through our atmosphere, we have all sorts of bits of ice and rock burning up - that's
what makes a meteor shower. This is also how we can predict when the showers are
going to occur, since we know where the comet's orbital path is located and when the
Earth will pass through it. It is possible to see individual meteors at just about any
time of the year, but the dates of showers are when people make an effort to view
them. The meteors tend to originate from one location in the sky (since that is where
the cometary debris is), so the names of meteor showers are based upon the
constellation they are centered upon. Here are some of the main showers, approximate
dates, intensity estimates and the comets that are thought to have produced them.
Shower Dates Meteors/hour Comet
Quadrantids January 3 40 unknown
Lyrids April 22 15 Thatcher
eta Aquarids May 4 20 Halley
delta Aquarids July 30 20

Perseids August 12 80 or more Swift-Tuttle
Orionids October 21 20 Halley
Taurids November 4 15 Encke
Leonids Nov 16 15 Tempel-Tuttle
Geminids Dec 13 50
Phaethon
(asteroid)
Ursids December 22 15 Tuttle
If you are interested in watching a meteor shower, then you should do a few things.
First of all, get far from city lights. You will also have to do some careful planning,
since the best observations of meteor showers take place after midnight. You should
set up a lawn chair or something comfortable to sit on, since you could spend quite
some time looking up at the sky. You may want to bring along something or someone
to keep you warm. You do not want to use things like binoculars or a telescope, since
you want to view as much of the sky as possible. You just have to sit back and watch
the show! Meteors are actually quite a ways up the atmosphere, typically dozens of
miles. Most of the objects that are meteors don't survive their passage through the
atmosphere but are completely vaporized. However, if objects are large enough and
442 | P a g e

dense enough, they can make it to the surface of the Earth, at which point they
become meteorites.
Figure 14. A large meteorite on a wagon.
This one weighs over 15 tons and was found
in Oregon. It is the largest meteorite ever
found in the US and the sixth largest in the
world. You can see it today in the Museum of
Natural History in New York. You definitely
didn't want to be under this one when it fell.
Most pieces of cometary debris evaporate in
the atmosphere, so most meteorites are
thought to originate from asteroids. Just as there are different types of asteroids, there
are also different types of meteorites. There are actually three main types. These are -
Stony - Just basic stony, rocky material, often with a burned crust on the
outside. These are the most common type of meteorite (95%), but are hard to
find (it is hard to tell whether an object on the ground is a plain old rock or a
meteorite). One of the rare varieties of these is known as the carbonaceous
chondrite, an carbon rich object thought to originate from c-type asteroids. The
neat thing about these is that some have been found to contain things like water
and amino acids. In case you forgot, amino acids are the building blocks of
proteins - one of the basic cornerstones of life! Life from a meteorite? - it's
possible. However, most of the stony meteorites aren't so interesting, and are
very similar to the s-type asteroids.
Stony-Irons - As the name implies, these are a mix of stone and metals. These
are actually quite rare (1%).
Irons - While the name implies these are made of iron, they are actually a mix
of mainly iron and nickel. This often helps distinguish them from plain old iron
from the Earth, since often industrial by-products don't have nickel in them.
About 4% of meteorites come in this form.




Figure 15. Various meteorites. Starting on the left, a stony meteorite, followed by a stony-iron,
an iron meteorite and a sliced iron meteorite showing the Widmanstatten figures.
443 | P a g e

Meteorites have been floating around in space for so long that they contain material
that originates from the time when the solar system formed, and they have not been
altered significantly since their formations. Meteorites provide very important
information about the early solar system, and the processes that went into the
formation of the planets. Meteorites have a wide variety of minerals within them, and
the requirements for the formation of those minerals (temperature, density,
composition), help us understand the early solar system characteristics. Some of these
minerals are very ordinary, like calcite, sulfur and pyrite (fool's gold), but things like
diamonds (yes, real diamonds) have been found in meteorites. There are even some
minerals that are found only in meteorites. Meteorites also help us determine the age
of the solar system through the process of radioactive decay. There is even a group of
meteorites discovered in Antarctica that are thought to be from Mars. How is that
possible? It is possible that an impact in the past on Mars threw enough material up
and away from Mars that it could have eventually made its way to the Earth. The main
feature that distinguishes these meteorites from others is the amount of certain gases
found in them, which is similar to the atmosphere of Mars in many respects. A similar
process has brought some pieces of the Moon to the Earth as well.
How can you tell if you have a meteorite in your hand? For the stony ones it is
difficult. Often the composition has to be analyzed extensively and they should be
checked by someone who knows what they are looking for (geologist-chemist). In the
case of iron meteorites, there are characteristic patterns in the metal that can be seen -
Widmanstatten Figures. They are produced by the heating and cooling of the metal.
They are very distinctive and not found in Earth rocks or metals. It also helps if you
find your suspected meteorite in a place where there are not too many rocks. Places
like Antarctica and large deserts are good places for meteorite hunting, though they
are not very nice places to visit.
Figure 16. An image of many
micrometeorites from the Cold
Regions Research and
Engineering Laboratory of the
Army Corp of Engineers. In case
you didn't know, 300 microns is
0.3 millimeters in size, smaller
than the size of the period at the
end of this sentence.
It is also rather easy to find
meteorites by just setting out a
collecting dish in your back yard.
Huh? I didn't say they would be big meteorites. Pretty much any location on Earth
444 | P a g e

will have micro-meteorites fall on it. These are really small meteorites that are sort of
like dust specks that are continually raining down onto the surface of the Earth. If you
set out a bowl of water or a petri dish with petroleum jelly in it, you'll catch of a few
meteorites after some time - along with bugs, pollen, regular dust, etc. You'd have to
use a microscope to find the micrometeorites. They tend to very spherical.
Finding meteorites on the Earth may seem like a difficult task, just imagine how
difficult it is to find a meteorite on another planet. Or is it? The Mars
roverOpportunity has found a pretty good sized meteorite on the surface of Mars,
which is about 2 feet across. Since the rover has the ability to chemically analyze
rocks, it was able to determine that this meteorite (shown here) is an iron meteorite.
Figure 17. Meteorite
discovered in 2009, on the
surface of Mars. Image
from NASA/JPL -
Caltech/Cornell University.

Impacts
Obviously, if some meteors
can make it all the way to
the surface of the Earth and
become meteorites, it is
possible for even larger
objects to strike the Earth. This has happened quite a bit in the past and it will happen
in the future - we just don't know when. Even within the last 100 years, people have
had some near-misses with meteorites. In 1938 a woman in Illinois heard a sound in
her garage and went out to find a meteorite sitting on top of her car. A few years later
a woman in Alabama was hit by a ricocheting meteorite that gave her a really nasty
bruise on her hip. In 1971 a house in Wethersfield, Connecticut was hit by a
meteorite. Eleven years later, a house about one mile away was hit by another
meteorite. One of the best recorded impacts was of an object that smashed through the
trunk of a woman's 1980 Chevy Malibu. This happened in 1992. The woman was
offered $69,000 for her wrecked car, which was a real bargain (since the car was 12
years old and rather run down). The meteorite that did the damage was 30 pounds.
The path that this particular piece of space debris took was recorded by many
observers in the eastern US, and one person filmed it (they were actually at a high
school football game when they caught the images, which can be seen here).
445 | P a g e


Figure 18. At left, the
damage to the forest of
Tunguska by the impact
there. At the right is the
blast area, where trees
were knocked down,
superimposed on a map
of Iowa. The force of
the blast extended
much further than this
region, though. This
only shows where the
trees fell.

These are all really small things; what if a really big thing hit the Earth? If you want to
see the effects of very large impacts, you need only go to Arizona to see a large
impact crater. There are also craters visible in other parts of the world, like Australia,
Canada and Africa. However, these are pretty old impacts, most pre-dating human
history. A big impact that occurred not too long ago took place in remote Siberia. At
about 7:17 AM on June 30, 1908, an object struck the Tunguska region of Siberia.
The glow of the fire from the impact was visible 1000 km from the site. People 60 km
from the impact site were knocked down by the force of the shock wave. The night
sky was kept bright by the fires ignited by this event. Actually, people as far away as
England noted the overly bright night sky during that summer. Some records indicated
that the sky was so bright that you could read a newspaper at midnight. Barometers
around the world measured a change in air pressure as the blast wave traveled around
the Earth several times. It wasn't until the 1930s that a team of scientists finally
reached the site, since it was located in remote swamp lands. No meteorite fragments
were found, but the forests in the region were either flattened or burned. Also, based
upon the tree pattern, it appeared that the object did not actually hit the ground but
exploded in the atmosphere. It is most likely that the object was a comet, since it
would have shattered much more easily than an asteroid and would have been easily
dissolved. If the object was an asteroid, there should have been more pieces of
meteoritic material in the area. Either way, it was pretty devastating. The main
casualties in the area were herds of reindeer and one man who was knocked to the
ground by the blast and died of internal bleeding. The impact had the strength of
roughly a 40 megaton nuclear bomb (about 1000 times as powerful as the bomb that
fell on Hiroshima).
446 | P a g e



Figure 19. Earthly impacts. The crater in
Arizona looks pretty impressive but is
only considered a small impact. The
impact in Iowa, shown on the right, was
much more powerful and caused a
greater amount of damage. Arizona
crater picture courtesy of USGS, D.
Roddy. Manson impact site image from
the Iowa Geologic Survey.
What about other big impacts? Most other impacts are so far in the past that we only
see the damage after the fact. Some were pretty big. Probably the most famous impact
feature is the one in Arizona. However, that was produced by an impact producing
only 3.5 megatons of TNT - pretty puny compared to other impacts. One of the
biggest impacts on the Earth took place in Iowa, near what is now the town
of Manson (near Fort Dodge). That impact occurred about 74 million years ago, so
most traces of it are not visible. However, it was much more powerful than the
Arizona impact. It packed a punch of about 1/2 million megatons of TNT. The area of
damage was pretty extensive at the time. The image above shows just how nasty such
an impact would be if it happened today.
447 | P a g e




Figure 20. Shoemaker-Levy 9's damage on Jupiter is seen in these images. At left, the fireball
produced by the largest fragment of the comet exploding in the atmosphere is seen. This image is
infrared, and the collision is actually occuring at a location that is slightly out of view, on the
edge of the planet as we see it. This means that not all of the fireball is being seen, only part of it.
The image in the middle is from the Hubble Space Telescope's UV camera, showing the remnant
heat from the previous impacts. The view on the left is also from the Hubble Space Telescope,
showing the effects of the impact features in the clouds, this time in visible light. All of the dark
spots are from previous impacts. Images from MSSSO, Hubble Space Telescope Comet Team.
Of course, we have been seeing evidence of damage on other worlds. Only recently
have we been lucky enough to see a major impact on a world. In 1993, a comet
named Shoemaker-Levy 9 (SL9 for short) was discovered. There's nothing too
unusual about that, but when further observations showed the motion of the comet to
be in orbit about Jupiter (and not the Sun), interest increased. Interest in SL9
intensified when it was determined that this comet wasn't just in orbit about Jupiter
but was actually on a collision course with the planet. A previous passage around
Jupiter had caused the comet to break into about 20 chunks. This meant that for about
one week, we could watch 20 pieces of cometary debris, some about 1 km in
diameter, smash into the cloud layers of Jupiter from the safety of the Earth - and boy,
was it a cool show! Infrared cameras showedlarge explosions occurring as each
fragment hit, and as Jupiter rotated around, we could see dark cloud features, sort of
like cosmic black eyes, on the clouds. The largest fragment had an impact energy of 6
million megatons of TNT (think of that - it was 12 times more powerful than the
Manson impact!) The temperatures produced by each explosion was much greater
than the temperature of the Sun's surface. The spots that were left by the impact were
larger than the Earth in size. Jupiter wasn't damaged by the impact, and evidence of
the impacts was gone in a few months, but for a while it was the best show around.
If you missed it, odds are you won't be able to see such an event again. Such an
impact event has the chance of occurring only once every thousand years... Or so we
448 | P a g e

thought. In July of 2009 an amateur astronomer noticed a new spot on Jupiter that
appeared literally overnight and looked very similar to the spots left by Shoemaker-
Levy 9. So perhaps the "once-in-a-life-time" estimate that we thought wasn't quite
correct. The Hubble Space Telescope took some images of the new spot, which will
likely last only a few months. You can see those images here. And it also appears that
someone has caught an impact on video - not what you'd see on America's Funniest
Home Video, but still very
interesting. Here is a link to the
video.
Figure 21. Something that will
happen in the future. Let's just
hope we're not around when it
does happen....Painting by Don
Davis, for NASA.
Certainly impacts have happened
on the Earth, but are they really
that bad? Most of the craters we
see today, from relatively recent
impacts, are relatively small, so
we should have no problem
surviving an impact, right? Well
sometimes, as in the case of
Tunguska, we don't have an
impact feature and the effects can
be devastating. Some scientists
propose that 12,900 years ago, a
comet exploded over North
America, which caused global
cooling and the start of the Ice
Age. This may have led to the extinction of large North American mammals. A nice
theory, but there are many proponents that point out flaws in the data and evidence -
some of which may not be extra-terrestrial in origin. A major impact on the Earth
would be devastating, but do we really have to worry about it happening?
The odds of impacts by large objects are pretty rare. However, they can and will
happen. Here is a graph showing the level of devastation caused by such impacts and
their frequencies. Those that cause global destruction are pretty rare. However, they
are still possible. Right now (August 2011), there are many objects that have very,
very, very slight chances of hitting the Earth. One of these with the very small chance
of hitting the Earth is asteroid #99942, Apophis(also called 2004 MN4), which was
449 | P a g e

initially calculated to have a 1 in 45,000 chance for an impact on April 13, 2036,
however after further observations of its orbit, that chance has dropped to about 4 in a
million. In case you were wondering the name Apophis is the Greek name for
the Egyptian god of evil, destruction and darkness, and he's a character on the Sci-fi
show Stargate. Seems like they wasted a really cool name on an object that isn't going
to hit us. One thing that is worth noting is that Apophis will come very close to the
Earth in 2029, perhaps close enough that its orbit is changed enough to make a future
impacting object. So even though the immediate danger doesn't exist, things can
change in the future.
The object that currently has the best chance of hitting the Earth is asteroid 2011 AG5,
which doesn't have any cute name associated with it (looks like we want to be sure
before we give it a really nifty name). It has a 1 in 2300 chance of hitting the Earth on
February 5, 2040. Click here to see its orbit.
At this point, there is a growing list of objects that we know of with a small chance of
hitting the Earth. There are many other objects out there, the PHAs, that have even
lower probabilities of hitting the Earth. Now, you may notice that I wrote the words
"that we know of" with some emphasis. That's because the job of finding these critters
isn't easy and there are not that many people who are actively looking for them. In
fact, in February of 2011, an object came relatively close to the Earth (less than 6000
km from us, or about 0.85 Earth radii) and we didn't know about its existence until 14
hours before it flew by! In fact this one was so close that the Earth's gravity changed
its orbit. Just keep in mind that sometime in the future we'll get hit again, and whether
it is tomorrow or 1000 years from now, there is no avoiding the fact that it will
happen.

Now that you've read this section, you should be able to answer these questions....
What are the various components of a comet and how do they change as a
comet moves closer to the Sun?
What is the "dirty snowball" model?
Do comets actually zip across the sky quickly?
Why are there two tails?
Why is Halley's comet so famous?
What other famous comets are there?
What sort of explorations have we done with comets?
What are asteroids composed of?
Where are asteroids located?
Why aren't asteroids "planets"?
What causes the Kirkwood gaps?
450 | P a g e

What is the difference between a meteor, meteoroid and a meteorite?
Why are meteor showers predictable?
What are the different types of meteorites and how are they linked to the
different types of asteroids?
How devastating were past impacts on the surface of the Earth?
Is the Earth ever again going to get hit by a large object?

You might also like