You are on page 1of 22

Star

At first glance, stars pretty much all look alike. Twinkly dots, scattered across the sky. But as I
talked about in episode 2, when you look more closely you see differences. The most obvious is that some
look bright and some faint. As I said then, sometimes that’s due to them being at different distances, but
it’s also true that stars emit different amounts of light, too. If you look through binoculars or take pictures
of them, you’ll see that they’re all different colors, too. Some appear white, some red, orange, blue, and
for a long time, the reason for this was a mystery. In the waning years of the 19th century,
astrophotography was becoming an important scientific tool. Being able to hook a camera up to a
telescope and take long exposures meant being able to see fainter objects, revealing previously hidden
details. It also meant that spectroscopy became a force, say it with me now, for science. A spectrum is the
result when you divide the incoming light from an object into individual colors, or wavelengths. This
reveals a vast amount of physical data about the object. But in the late 1800s, we were only just starting

to figure that out. Interpreting stellar spectra was a tough problem. The spectrum we measure from a star
is a combination of two different kinds of spectra. Stars are hot, dense balls of gas, so they give off a
continuous spectrum; that is, they emit light at all wavelengths. However, stars also have atmospheres,
thinner layers of gas above the denser inner layers. These gases absorb light at specific wavelengths from
the light below depending on the elements in them. The result is that the continuous spectrum of a star has
gaps in it, darker bands where different elements absorb different colors. At first, stars were classified by
the strengths of their hydrogen lines. The strongest were called A stars, the next strongest B, then C, and
so on. But in 1901, a new system was introduced by spectroscopist Annie Jump Cannon, who dropped or
merged a few of the old classifications, and then rearranged them into one that classified stars by the
strengths and appearances of many different absorption lines in their spectra. A few years later, physicist
Max Planck solved a thorny problem in physics, showing how objects like stars give off light of different
colors based on their temperature. Hotter stars put out more light at the blue end of the spectrum, while
cooler ones peaked in the red. Around the same time, Bengali physicist Meghnad Saha solved another
tough problem: how atoms give off light at different temperatures. Two decades later, the brilliant
astronomer Cecelia Payne-Gaposchkin put all these pieces together. She showed that the spectra of stars
depended on the temperature and elements in their atmospheres. This unlocked the secrets of the stars,
allowing astronomers to understand not just their composition but also many other physical traits. For
example, at the time, it was thought that stars had roughly the same composition as the Earth, but Payne-
Gaposchkin showed that stars were overwhelmingly composed of hydrogen, with helium as the second
most abundant element. The classification scheme proposed by Cannon and decoded by Payne-
Gaposchkin is still used today, and arranges stars by their temperature, assigning each a letter. Because
they were rearranged from an older system the letters aren’t alphabetical: So the hottest are O-type stars,
slightly cooler are B, followed by A, F, G, K, and M. It’s a little weird, but many people use the
mnemonic, “Oh Be A Fine Guy, Kiss Me,” or “Oh Be A Fine Girl, Kiss Me,” to remember it — which

was dreamed up by Annie Jump Cannon herself! Each letter grouping is divided into 10 subgroups, again
according to temperature. We’ve also discovered even cooler stars in the past few decades, and these are
assigned the letters L, T, and Y. The Sun has a surface temperature of about 5500° Celsius, and is a G2
star. A slightly hotter star would be a G1, and a slightly cooler star a G3. Sirius, the brightest star in the
night sky, is much hotter than the Sun, and is classified as an A0. Betelgeuse, which is red and cool, is an
M2. Stars come in almost every color of the rainbow. Hot stars are blue, cool stars red. In between there
are orange and even some yellow stars. But there are no green stars. Look as much as you want, and you
won’t find any. It’s because of the way our eyes see color. A star can put out lots of green light, but if it
does it’ll also emit red, blue, and orange. And our eyes mix those together to form other colors. A star can
actually emit more green light than any other color, but we’ll wind up seeing it as white! How do I know?
Because if you look at the sun’s spectrum, it actually peaks in the green! Isn’t that weird? The Sun puts
out more green light than any other color, but our eyes see all the mixed colors together as white. Wait,
what? White? You may be thinking the Sun is actually yellow. Not really. The light from the Sun is
white, but some of the shorter wavelengths like purple and blue and some green get scattered away by
molecules of nitrogen in our air. Those appear to be coming from every direction but the Sun, which is
why the sky looks blue. The Sun doesn’t emit much purple, so the sky doesn’t look purple, and the green
doesn’t scatter as well as blue. That gives the Sun a yellowish tint to our eyes, and looking at the Sun is
painful anyway, so it’s hard to accurately gauge what color it appears. That’s also why sunsets look
orange or red: You’re looking through more air on the horizon to see the Sun, so all the bluer light is
scattered away. So we can use spectra to determine a lot about a star. But if you combine that with
knowing a star’s distance, things get amazing. You can measure how bright the star appears to be in your
telescope, and by using the distance you can calculate how much energy it’s actually giving off — what
astronomers call its luminosity. An intrinsically faint light looks bright if it’s nearby, but so does a very
luminous light far away. By knowing the distance, you can correct for that, and figure out how luminous
the objects actually are. This was, no exaggeration, the key to understanding stars. A lot of a star’s
physical characteristics are related: Its luminosity depends on its size and temperature. If two stars are the
same size, but one is hotter, the hotter one will be more luminous. If two stars are the same temperature,
but one is bigger, the bigger one will be more luminous. Knowing the temperature and distance means
knowing the stars themselves. Still, it’s a lot of data. A century ago, spectra were taken of hundreds of
thousands of stars! How do you even start looking at all that? The best way to understand a large group of

objects is to look for trends. Is there a relationship between color and distance? How about temperature
and size? You compare and contrast them in as many ways possible and see what pops up. I’ll spare you
the work. A century ago astronomers Ejnar Hertzsprung and Henry Norris Russell made a graph, in which
they plotted a star’s luminosity versus its temperature. When they did, they got a surprise: a VERY strong
trend. This is called an HR Diagram, after Hertzsprung and Russell. It’s not an exaggeration to call it the
single most important graph in all of astronomy! In this graph, really bright stars are near the top, fainter
ones near the bottom. Hot, blue stars are on the left, and cool, red stars on the right. The groups are pretty
obvious! There’s that thick line running diagonally down the middle, the clump to the upper right, and the
smaller clump to the lower left. This took a long time to fully understand, but now we know this diagram
is showing us how stars live their lives. Most stars fall into that thick line, and that’s why astronomers call
it the Main Sequence. The term is a little misleading; it’s not really a sequence per se, but as usual in
astronomy it’s an old term and we got stuck with it. The reason the main sequence is a broad, long line
has to do with how stars make energy. Like the Sun, stars generate energy by fusing hydrogen into helium
in their cores. A star that fuses hydrogen faster will be hotter, because it’s making more energy. The rate
of fusion depends on the pressure in a star’s core. More massive stars can squeeze their cores harder, so
they fuse faster and get hotter than low mass stars. It’s pretty much that simple. And that explains the
main sequence! Stars spend most of their lives fusing hydrogen into helium, which is why the main
sequence has most of the stars on it; those are the ones merrily going about their starry business of
making energy. Massive stars are hotter and more luminous, so they fall on the upper left of the main
sequence. Stars with lower mass are cooler and redder, so they fall a little lower to the right, and so on.
The Sun is there, too, more or less in the middle. What about the other groups? Well, the stars on the
lower left are hot, blue-white, but very faint. That means they must be small and we call them white
dwarfs. They’re the result of a star like the Sun eventually running out of hydrogen fuel. We’ll get back to
them in a future episode. The stars on the upper right are luminous but cool. They must therefore be huge.
These are red giants, also part of the dying process of stars like the Sun. Above them are red
SUPERgiants, massive stars beginning their death stage. You can see some stars that are also that
luminous but at the upper left; those are blue supergiants. They’re more rare, but they too are the end
stage for some stars, and again we’ll get to them soon enough in a future episode. But I’ll just say here
that it, um, doesn’t end well for them. But, on a brighter note, we literally owe our existence to them. And
this implies something very nifty about the HR diagram: Stars can change position on it. Not only that,
but massive stars versus low mass stars age differently, and go to different parts of the HR diagram as
they die. In many ways, the diagram allows us to tell at a glance just what a star is doingwith itself. This
difference between the way low mass stars like the Sun and higher mass stars age is actually critical to
understanding a lot more about what we see in the sky… so much so that they’ll be handled separately in
later episodes. I’m sorry to tease so much about what’s to come, but this aspect of stars — finally
understanding them physically — was a MAJOR step in astronomy, leading to understanding so much
more. Today you learned that stars can be categorized using their spectra. Together with their distance,
this provides a wealth of information about them including their luminosity, size, and temperature. The
HR diagram plots stars’ luminosity versus temperature, and most stars fall along the main sequence,
where they live most of their lives.
Brown Dwarfs

The sky, we now know, is full of stars AND planets. Stars are massive enough to fuse hydrogen into
helium in their cores, generating energy. The heat created by that process tries to expand them, but their
gravity balances that outward force, creating an equilibrium. Planets, even gas giants like Jupiter, are far
too small to generate fusion. The stuff inside them resists being squeezed, so their gravity is balanced by
simple gas pressure. Jupiter is only about 1% the mass needed to have fusion going on in its core. That’s a
pretty big gap between a big planet and a small star. It’s natural to ask what would happen if we dumped
more mass onto Jupiter. Eventually it would become a star -- the pressure in its core would get high
enough to initiate hydrogen fusion. But what if we stopped just short of that? What if we have an object
far more massive than a planet, but not quite massive enough to become a true star? What sort of thing
would we have then? What indeed. By the late 1950s, astronomers were starting to get a pretty good
handle on how stars worked. The mathematical equations that governed the physical processes of fusing
hydrogen into helium were being worked out, and applied to what we knew from observing the stars
themselves. In the 1960s the idea that you could have a star with a minimum mass was becoming clear; if
it had less than about 0.075 times the Sun’s mass, roughly 75 times the mass of Jupiter, it simply lacked
the oomph needed to squeeze hydrogen together hard enough to generate fusion. What would such an
object look like? Well, it might form like a star, collapsing from a gas cloud just like the Sun did 4.6
billion years ago, but instead of turning on and becoming a star, it would simply sit there, cooling. It
might start off pretty hot, due to the physical forces that made it, but it couldn’t sustain that heat. Like a
charcoal ember, it would radiate its heat away. After a few billion years it would be cold, black, and for
all intents and purposes dead. As people started working out what such an object would be like, they tried
to come up with a name for them. These things were black, and small, but the name “black dwarf” was
already being used for another type of object. Some people called them sub-stellar objects, but that wasn’t
terribly catchy. Really low mass stars are red, and these new objects would be so cool that they’d emit
light in the infrared, and almost nothing at all in the visible. So they’re somewhere between red and black.
Jill Tarter, then a young astronomer working in the field but who later made a name for herself looking
for aliens — and oh boy, we’ll get to that later — dubbed them “brown dwarfs.” he didn't mean it
literally; stars can’t be brown. But the name stuck. Work proceeded in figuring out what brown dwarfs
were like, and a lot of progress was made despite there not being any actual examples9o f them found.
But the hunt was on. Now as I talked about in Episode 26, astronomers classify stars by their temperature.
The hottest are O stars, then B stars which are slightly cooler, down through A, F, G, K, and with the
coolest stars being M. But then, in 1988 astronomers found a star that was so cool it was distinct from
even the M class stars. It was the first of a new, cooler class of stars, so it was given the letter L. Why L?
Because there wasn’t any other astronomical object that used that letter, so why not? Many more such L
stars were found, but still these weren’t true brown dwarfs; these stars were massive enough to initiate
fusion in their cores. Worse yet, when brown dwarfs are first born they’re very hot. They can mimic
higher-mass L stars for a while, looking just like them, making it hard to distinguish between the two. But
then a way out was found. A low-mass brown dwarf, it was determined, would have lithium in it, whereas
normal stars wouldn’t. Lithium is an element, the next one in the Periodic Table after hydrogen and
helium. It can be fused much like hydrogen can, and regular stars would quickly use up their supply of

lithium when they were still young. Brown dwarfs lighter than about 65 times the mass of Jupiter
wouldn’t fuse lithium at all. Very careful observations of an object would be able to detect lithium if it
were there. That would provide a test to distinguish brown dwarfs from regular stars! The lithium test
isn’t perfect, but it does work under a lot of circumstances. Astronomers began using it to look for actual,
real brown dwarfs.And so they found one. In 1995, a group of astronomers was observing the Pleiades, a
nearby cluster of stars that’s visible to the naked eye. They were trying to find the faintest stars they could
in the cluster to get a complete sample of its membership. The advantage of this is that the distance to the
cluster was pretty well known, so a faint star in it must be very low mass. They found an oddball object,
which they named Teide 1. It was very red and cool, and best of all, lithium was found in its spectrum.
The best models of stellar mass showed that it had about 50 times the mass of Jupiter, or 0.05t imes the
mass of the Sun. It was clearly sub-stellar. Huzzah! The very first true brown dwarf had been found. At
just about the same time, astronomers found that another nearby star, called Gliese 229, had an extremely
faint companion. Spectra showed that it was even weirder than Teide 1. It also had lithium, and so was
clearly a brown dwarf. But its spectrum showed it had METHANE in its atmosphere. Methane is a
delicate molecule, and would break down even in the mild heat of Teide 1’s atmosphere. This new object,

called Gliese 229b, was even cooler than Teide 1. It was looking like we needed yet another letter to
classify stars. And so T dwarfs became a thing. On a personal note, when I worked on Hubble Space
Telescope, Gliese 229b was one of our camera’s first targets after it was installed on Hubble in 1997. I
was lucky enough to work on the spectrum we took of it, and it was freaky. It emitted almost no light in
the visible part of the spectrum, and rocketed up in the infrared. I had seen a lot of stellar spectra before,
but nothing like this. Remember, Gliese 229b had only been discovered two years before! I became so
intrigued by it I wound up studying low mass stars and brown dwarfs for several years after. Well, it
didn’t take long before more brown dwarfs were found. In 2009, NASA launched the Wide-field Infrared

Survey Explorer, or WISE, an orbiting observatory designed to scan the entire sky looking at infrared
light. It found hundreds of brown dwarfs, and now at least 2000 are known, with more found all the time.
Some are so cool that they form yet another classification: Y Dwarfs. So now we have O B A F G K M L
T and Y. You’re on your own for an acronym here. So if brown dwarfs aren’t brown, what color are they?
Some are so cool they don’t emit visible light at all, so they’d be black. You could be right over one and
you wouldn’t see it. But some are still warm, and so give off some visible light, feeble as it might be.
What color would they look? Funny thing. They might be magenta. You’d think they’d be really red,
because of their temperature. But it’s a bit more complicated than that. Remember, they have molecules
in their atmospheres that absorb specific colors of light. In some brown dwarfs, there are molecules like
methane and even water—well, steam at those temperatures—that are pretty picky about what colors they
absorb. Some of these molecules block more red light than blue, so that messes with their colors, making
them look magenta. WISE takes pictures in the infrared, which our eyes can’t see. To make pictures,
astronomers map each infrared color to one our eyes can see. So an image using the shortest wavelength
infrared detector is displayed as blue, the medium wavelength one green, and the longest one red. Brown
dwarfs put out a lot of light in the intermediate wavelength WISE sees,so weirdly, they appear gree n in
WISE pictures. That does make them easy to spot in those images, even when thousands of other stars are
visible, too. The physical nature of brown dwarfs is just as weird as you’d expect. For one thing, they
have a very unusual characteristic: As they get more massive, they don’t get any bigger. Usually, if you
dump mass onto an object it gets bigger; take two lumps of clay and smush ‘em together and you get one
more massive, slightly larger lump. Same with planets and stars. But brown dwarfs are different. At their
cores the density is very high, and the physics is a bit different than what you’d expect. The details are
complex but the end result is that when you add more mass to them they actually get DENSER, not
bigger. This effect becomes important right around the mass of Jupiter, which means that a brown dwarf
twice as massive as our biggest planet won’t actually be a whole lot bigger. So what’s the difference
between a small brown dwarf and a really big planet? Well, not much. Nature isn’t as picky as we are
about having narrowly-defined borders between classes of objects. Some people say a planet forms from a
disk of material around a star, growing larger as it accretes stuff, while a brown dwarf collapses directly
from a cloud of gas and dust. But then you could have two objects the same mass, and which look exactly

the same, yet one would be a planet and one a brown dwarf, depending on how they formed. That strikes
me as… inconvenient. Astronomers are still debating this. And it gets worse. For example, as I said
before, brown dwarfs over 65 times Jupiter’s mass fuse lithium. It turns out that ones more massive than
about 13 times Jupiter can also fuse deuterium, an atom that’s very similar to hydrogen, except it has a
proton and a neutron in its nucleus. But neither of these fuses actual hydrogen, so they’re not considered
true stars. That’s still a little arbitrary, so again I don’t make too much of a fuss about it. I think it’s best
not to think of them as planets or stars, but something with characteristics of both. For example, the way
their atmospheres behave depends a lot on how hot they are. In some, iron is vaporized, a gas. In others,

they’re just cool enough that iron condenses out of the atmosphere… which means it literally rains molten
iron! One more thing. The nearest star to the Sun is a red dwarf called Proxima Centauri, which orbits the
binary star Alpha Centauri. It’s about 4.2 light years away. In 2013, astronomers announced the discovery
of a binary pair of brown dwarfs, called Luhman 16. They’re only 6.5 light years away, and became the

3rd closest known star system to Earth. You gotta wonder: Could there be an even fainter, cooler brown
dwarf closer to us? We know there’s none in our solar system, even out in the Oort cloud; it would’ve
been seen by now by one of several different sky surveys. But a light year or two out? Maybe. Is Proxima

Centauri REALLY the closest star, or will we find one even closer? It seems unlikely, but no more
unlikely than the existence of brown dwarfs themselves. Maybe sometime soon we’ll have to rewrite
astronomy textbooks. Again. Today you learned that brown dwarfs are objects intermediate in mass
between giant planets and small stars. They were only recently discovered, but thousands are now known.
More massive ones can fuse deuterium, and even lithium, but not hydrogen, distinguishing them from
“normal” stars. Sort of.
Low Mass Stars

INTRODUCTION:

Stars in the sky look pretty. Flickering, intense diamonds dotting the velvety night. But make no
mistake: They are churning cauldrons of violence, barely constrained thermonuclear generators, creating
enough energy to vaporize the Earth a thousand times over. Their lives depend on it. Heck, our lives
depend on it! But in their case, how they live, and how long they live, depends on that creation of energy.
And in this Universe, noting lasts forever.

DISCUSSION:

Very roughly speaking, we can divide stars into two groups: low mass stars and high mass stars.
That line lies around 8 times the mass of the Sun, and the reason for that is something we’ll get to shortly.
This get very exciting at the high end, but for now, let’s take a look at the low mass end of the scale.
Things are simpler there.

Stars make energy in their cores by fusing hydrogen into helium. This is actually a pretty
complicated process with lots of steps, but I the end 4 protons plus some other ingredients get turned into
one helium nucleus. Each time this happens a little bit of energy is released, but in the core of a star it
happens countless times every second. It adds up to a lot of energy: enough to power a star in fact.

The rate at which hydrogen fusion occurs depends on how much pressure the star has in its
core. A higher mass star squeezes hydrogen harder, so it fuses far more quickly. The rate is so much
higher, in fact, that even though they’re a lot bigger than low mass stars, high mass stars run out of
hydrogen fuel in their cores more quickly! The lower mass a star is, the longer it lives.

In the lowest mass stars, cool red dwarfs, the rate is so slow they last a long, long time. But
there’s more to it than that. They fuse hydrogen only in their centers, and their cores aren’t very big.
Outside of this small region the gas is convective, which means the hot stuff rises all the way to the
surface, cools, then falls back to the core. That’s important, because it means the entire star is available
for fuel. Both hydrogen and helium rise up through the star and then fall back down. The star can’t fuse
the helium, but the hydrogen that makes it back to the center is just more grist for the mill. Even after a
long, long time, theses stars can still shine as the hydrogen mixed throughout the star eventually makes it
down to the center. So how long can a really low mass red dwarf last? A trillion years. A trillion! The
Universe itself is less than 14 billion years old, so even the oldest red dwarf stars are basically infants.
They’re just barely getting a start in life, and they’ll look pretty much the same for the next, 990 billion
years or so. Eventually, though, they do run out of fuel. When one does, the star itself will be nearly pure
helium (plus whatever heavier elements it was born with), and fusion will cease. It’ll then cool, which
will take many more billion of years. Finally, it will be a cold, black, dead star. And for them, that’ll be
pretty much that. Or so we think; the Universe is far too young for any of these red dwarfs to have run out
of fuel yet.

That’s true for stars at the lower end of the mass scale, up to roughly a third of the mass of the
Sun. But stars with more mass than that, stars more like the Sun itself, work a little differently. Their
cores are bigger, and hotter, and denser. The different conditions inside them means the material in them
doesn’t convect, the hot stuff doesn’t rise away from the core. What’s in the core stays in the core.
Outside the core there is a very deep and active convection zone, but that doesn’t interact with the
material in the core itself. This means that only do stars near the Sun’s mass use up their fuel more
quickly than red dwarfs, they have relatively less fuel to work with as well. Depending on their mass,
their lives aren’t nearly as long.

For the Sun, that total lifespan is about 12 billion years, give or take. The Sun is nearly five
billion years old now, so it’s approaching middle age, its life is similar to other stars of similar mass, so
let’s take a look into the far future. Will our Sun go gently into that good night? Ever since the Sun was
born, it’s been fusing hydrogen into helium in its core. The helium is trapped in there, unable to be used
as fuel, and is building up in the core like ash in a fireplace. As is does, the density in the Sun’s core
slowly increases. When you compress a gas it heats up. That means, every day, the Sun’s core get a wee
bit hotter. That extra heat works its way out through our star, heating up the outer layers too. A hotter gas
shines more brightly, so the Sun has been steadily increasing in brightness too. It’s an incredibly slow
but inexorable process; the Sun’s increased in luminosity by about 40% since it was born.

And it’s not done. Helium is still building up in its core and eventually it’ll run out of hydrogen to
fuse. Without fuel, fusion in the core will stop. Helium can be fused into carbon dioxide (and a little bit of
oxygen and neon), but it has to be incredibly hot to do so, and the Sun’s core won’t be anywhere near this
conditions. Still, half the Sun’s mass is bearing down on the core, so even though the fusion stops, the
core will continue to contract and heat up. It’ll get so hot that the temperature outside the core will get
to hydrogen fusion temperatures. Like igniting a match by holding it near a flame, the hydrogen outside
the core will start to fuse. That’ll occur in a shell surrounding the core, adding its heat to the outer layers
of the Sun a well. When you heat a gas it expands. The Sun’s outer layers will do just that, and our star
will grow to well over twice its current size. Astronomers call a star like this a subgiant. Not only will it
grow, but its color will change! The Sun will be giving off more total energy than it does now, but its
surface area will increase so much that the amount of energy radiated per square meter will go down.
There’ll jut be a lot more square centimeters. This means ironically that the surface will actually cool; the
Sun’s temperature will go down. Cooler stars are red, and so the Sun will be too.

But the core still isn’t done. The details are complicated, but the core continues to contract and heat up. It
gets so hot that the outer layers swell even more, and the Sun can bloat up to a fantastic 10 to 150
times its present size! It will be then a red giant. Red giants are fantastically bright because of all the
energy percolating up from their interiors, coupled with their enormous size. When it turns into a red
giant, the Sun will increase its luminosity by an incredible 2000 times. That huge increase in luminosity
does more than just make the Sun bright, though. As a red giant, the Sun will be swollen that the gravity
at its surface will drop substantially from what it is now. The force of gravity holding it down is weak, but
the force from all that light blasting out from below is strong, easily overpowering gravity. Material on
the surface will get blown off like a leaf in a hurricane. Think of it as a super solar wind. That means the
Sun will shed mass, a lot faster than it does now. While it’s a red giant, it’ll lose fully a third of its mass.
At some point, the core contracts and heats up so much that the conditions will finally be primed for
helium fusion. Suddenly, helium is converted into carbon, releasing a lot of energy. In a weird twist, due
to a lot of very complicated physics, the core itself winds up expanding, absorbing most of that energy. In
the end, less energy is pumped into the outer layers, so the Sun’s outer layers contract. The Sun shrinks.
The outer layers get dense again, which warms up a bit, and the Sun turns from red to orange. It’s still
pretty big, more than 10 times its current day size, and puts out 20-50 times as much energy as it does
now. But then we see a sort of reboot of what’s happened in the core before. This time, though, helium is
fusing in the core, and the carbon is building up like ash. Once again, the core starts to contract and heat
up. This dumps more energy into the outer layers, they expand and cool, and the Sun turns into a red giant
AGAIN. This time it’s even brighter and bigger, and loses even more mass, shedding over half of what’s
left. In the meantime, helium fusion isn’t very stable, to say the least. It swings wildly, increasing and
decreasing its rate by huge amounts on very short timescales. The Sun will probably undergo a series of
tremendous paroxysms, epic eruptions as the helium fusion spikes, creating huge surges in energy
production. It’ll lose even more mass each time. Finally, though, we approach the endgame. The Sun
doesn’t have enough mass to squeeze carbon atoms enough to fuse them. They build up in the core until
there’s no more helium, and fusion stops. On top of that… well, there’s nothing on top of that. By this
time the Sun will have literally blown off all its outer layers. It’s essentially a naked core: a hot, intensely
bright super-compact ball, not much bigger than Earth. We call this a white dwarf. Its fate is sealed: It
can’t generate energy, so it’ll cool and fade over the next few tens of billions of years. At that point, we
can safely say the Sun is dead. BUT, that’s not always the case for

all stars. Some, with more mass than the Sun, still go through another phase before dying. They form
what are called planetary nebulae, and these are so amazing and beautiful that they rate their own episode,
which, along with more fun stuff about white dwarfs, we’ll get to next time. So that’s it for the Sun. But
what about US? Of course, the Earth won’t fare very well during all this. It’ll get cooked before the Sun
even becomes a subgiant, and by the time the Sun is a red giant we’ll be good and fried. The average
temperature on our planet will be so high that it’ll melt rocks, and in a sense it’ll look much like it once

did so many billions of years before, when it first formed: a molten ball of rock. But it gets worse. When
the Sun expands, it’ll reach a radius very close to the size of the Earth’s orbit now. Does that mean our
planet itself will be consumed by the Sun? Maybe not. As the Sun expands, remember, it loses mass. That
means its gravitational pull will weaken, and Earth will move out from the Sun, into a bigger orbit. If
things work out right for us, the Earth will back off faster than the Sun expands, and we’ll avoid finding
out what it’s like to be inside a star. Due to a quirk of physics, if the Sun loses more than half its mass, its
gravity will decrease so much the planets will no longer be gravitationally bound to it, and they’ll be
flung into interstellar space. I’m not sure that’s any better for us. Either way, the Earth will have been
long dead, and hopefully humanity will have figured out the trick to interstellar travel by that time. All
those Earth-like exoplanets around younger stars will look pretty nice in a billion years or two. Mind you,
this isn’t happening any time soon. The Sun won’t become a subgiant until it’s nearly 11 billion years old,
over 6 billion years from now. The first red giant stage happens when it’s over 11.5 billion years old, and
the second stage a half billion years later. This is a long, long time coming. If the Sun had more mass it
could avoid this fate. A star with about 8 times the Sun’s mass can: It has enough mass to squeeze those

carbon nuclei hard enough to fuse them together. Those high mass stars have a much different and far
more explosive fate than lower mass stars like the Sun…and trust me, we’ll cover that in lots of fun detail
soon. And while this all sounds very bleak, not all is lost. This is all part of the natural cycle of the
Universe, and in fact we wouldn’t be here at all if it weren’t for the way stars live and die. Be thankful for
what you have -- and that we can not only observe it, but understand it, too. Today you learned that very
low mass stars live a long time, fusing all their hydrogen into helium over a trillion years. More massive
stars like the Sun live shorter lives. They fuse hydrogen into helium, and eventually helium into carbon
(and also some oxygen and neon). When this happens they expand, get brighter, and cool off, becoming
red giants. They lose most of their mass, exposing their cores, and then cool off over many billions of
years.
White Dwarfs and Planetary Nebulae

Hey folks, Phil Plait here. In the last episode of Crash Course Astronomy, I talked about the eventual fate
of the Sun, and other low mass stars like it. After a series of expansions and contractions, they blow off
their outer layers, become white dwarfs, and fade away over billions of years. The end. Except not so
much. First, white dwarfs are pretty awesome objects, and worth investigating. And second, when a star
becomes a white dwarf it produces what is, quite simply, one of the most gorgeous objects in space To
recap, when the Sun ages, it undergoes a series of changes in its core. It’s fusing hydrogen into helium
now, today, and will eventually fuse helium into carbon, and it’ll make a dash of oxygen and neon too.
But when it runs out of helium to fuse, it’s in trouble. It doesn’t have enough mass to squeeze the carbon
nuclei together, so they can’t fuse. Fusion is the Sun’s energy source. Once the core is nearly pure carbon,
that power is switched off. By this time, nearly 8 billion years from now, the Sun’s outer layers are gone,
expelled by all the shenanigans going on in the core. What’s left of our star is just its core, exposed to the
dark of space. Over the next few billion years it’ll cool and fade to black. That might seem like the end of
the story. But I skipped a step, and it’s a beauty. When helium fusion stops, the Sun’s core will have
about half the mass of what the Sun does today; the rest will have blown away into space around it. What
remains is basically composed of electrons and carbon nuclei, mixed with a small amount of a few other
elements. So what kind of an object are we left with here? You may know that like charges repel;
electrons have a negative charge and repel each other. The tighter you squeeze them, the stronger that
pressure. There’s also a second force, called electron degeneracy pressure. It’s a result of some of the
weird rules of quantum mechanics (for you QM nerds, it’s The Pauli Exclusion Principle). This describes
how sub-atomic particles behave on teeny scales. One of these rules is that electrons really hate to be
squeezed together, above and beyond simple electric repulsion. This resistance is phenomenally strong,
and becomes the dominant force in supporting the core of a star once helium fusion stops. By the time
this electron degeneracy pressure balances the core’s immense gravity, the core is only about the size of
Earth, 1% the original width of the Sun. And it’s called a white dwarf. Listing its characteristics is enough
to melt your brain. Ironically, everything about it gets amplified as its size shrinks. It becomes
ridiculously dense; a single cubic centimetre of it, the size of a six-sided die, has a mass of a million
grams — one metric ton. An ice cream scoop of white dwarf material outweighs 60 cars. Because it’s so
dense, the gravity at the surface of a white dwarf screams up, easily topping 100,000 times the Earth’s
gravity. If you have a mass of 75 kilos, you’d weigh 7,500 tons if you stood on the surface of a white
dwarf. Not that you can. Stand there, I mean. You’d be flattened into a greasy smear. But not for long.
Newborn white dwarfs are hot; they can glow at a temperature of upwards of 100,000 degrees Celsius. If
you were on the surface, you’d be a vaporized and ionized smear. Their intense heat makes them white,
and they’re small. Hence their name. They’re so hot they also glow in the ultraviolet, even in X-rays.
Weirdly, though, because they’re so small, they’re actually quite faint. The closest one to us, Sirius B, can
only be seen with a telescope even though it’s nine light years away, one of the ten closest known stars!
Over 10,000 white dwarfs have now been found in our galaxy. Still, any gas near a newly formed white
dwarf is likely to be affected by the intense radiation pouring out of it. And hey, wait a sec. When a star
like the Sun is in its final death throes, it expels its outer layers as a gaseous wind. You don’t suppose…?
Yup. By the time that white dwarf forms, the gas blown off hasn’t gotten very far from it, at most a light
year or two. That’s plenty close enough to get zapped by the white dwarf radiation, causing that gas to
glow in response. What does something like THAT look like? Why, it looks like this. This object is what
we call a planetary nebula. It’s a funny name, and like so many other names it’s left over from when these
objects were first discovered. The astronomer William Herschel — the same man who discovered
infrared light and the planet Uranus — gave them that name, because they appeared like small green disks
through the eyepiece. The first planetary nebula was discovered in 1764 by the French astronomer
Charles Messier, who spent years scanning the skies looking for comets. He kept seeing faint fuzzy
objects that he mistook for comets, so he decided to make a catalog of them, a sort of “avoid these
objects” list. That list is now a staple of amateur astronomers, because ironically it contains some of the
best and brightest objects to observe. Among them is M 27 — the 27th object on Messier’s list — one of
the biggest planetary nebulae in the sky, and one of my favorites; I love seeing it through my telescope
when it’s up high in the summer. Planetary nebulae can be a bit tough to observe; most are small and
faint. On film, even with long exposures, they can appear to be little more than disks. For a long time,
they weren’t thought to be terribly complicated; when a star becomes a red giant and blows off its outer
layers it’s rotating very slowly, so the wind should blow away in a sphere surrounding the star. Many
planetaries (as we call them for short), are round and look like soap bubbles, pretty much what you expect
when you look at an expanding shell of gas. But with the advent of digital detectors, their fainter
structures became clearer, and the true beauty of these phenomenal objects was revealed. Some are
elongated. Some have spiral patterns. Some have jets shooting out on either side. Some have delicate
tendrils streaming away from them. In fact, only a handful of the hundreds known are actually circular!
Clearly, there’s more to planetaries than meets the eye. If the wind from a star blows off in a sphere, how
can planetaries come in all these fantastic shapes? It turns out the real situation is more complicated. As
usual. When a star is a red giant, it spins slowly, and blows off a dense but slow solar wind. If there’s
nothing else happening to the star, then that wind will blow outward in all directions, spherically.
However, as those outer layers of the star peel away, they expose the deeper, hotter part of the star. The
star starts to blow a much faster, though far less dense wind. That wind catches up with and slams into the
slower wind. When that happens, you get that idealized soap bubble nebula. But some stars are binary;
two stars that orbit very close together. We’ll go into detail on them in a later episode, but, if the dying
star has a companion, they may circle each other rapidly. That will shape the wind, forcing more of it
outward in the plane of the stars’ orbits due to centrifugal force. The overall shape of the expanding gas is
flattened, like a beach ball someone sat on. When the fast wind kicks in, it slams into that stuff in the
orbital plane and slows down. But there’s less stuff in the polar direction, up and down out of the plane.
It’s easier for the wind to expand in those directions, and it forms huge lobes of material stretching for
trillions of kilometers. That’s a very common shape for planetary nebulae. But to explain the shapes we
see, the two stars would have to orbit improbably close. Most binaries aren’t that tight. So what can cause
these shapes? When I was in graduate school, my Master’s degree advisor, Noam Soker, came up with a
nutty idea; maybe the stars had planets, like in our solar system. If the star expanded into a red giant and
swallowed them, it would take millions of years or more for the planets to vaporize. And for all that time
they’d be orbiting INSIDE the star, moving faster than the star itself. Like using a whisk to beat eggs, the
planets inside the star would spin it up, causing it to rotate faster… fast enough to explain the shapes of
planetaries. That was in the early 1990s. A few years later, the first exoplanets were found, and we saw
that massive planets orbiting very close to their stars were common. I suspect this is why we see so many
weird and fantastic shapes in planetaries; their progenitor stars swallowed their planets. So planetary
nebulae really may owe their existence to planets! And we come… full circle. The glow in a planetary
nebula is due to the hot central white dwarf exciting the gas. Most of the gas is hydrogen, which glows in
the red. However, a lot of the gas is oxygen. There’s not nearly as much oxygen as hydrogen, but kilo for
kilo oxygen glows more brightly than hydrogen due to the atomic physics involved. This oxygen glows
green, giving planetaries their characteristic hue. Funny: When this green glow was first analyzed
spectroscopically, astronomers couldn’t identify the responsible element making it. They dubbed it
nebulium, but eventually figured out is was just extremely tenuous oxygen. Other colors can be found,
too. Nitrogen and sulfur glow red, and oxygen can emit blue as well, all adding to the beauty of these
celestial baubles. But these aren’t just pretty pictures: The structure, color, and shape of a planetary
nebula tells us about the life of the star that formed it. We learn even more about stellar evolution by
studying how stars die. Mind you, the gas in a planetary nebula is still expanding, cruising outward from
its initial momentum of being thrown off the star. Eventually, the gas expands so much it thins out, and it
stops glowing. That takes a few thousand years, so when you see a planetary nebula you’re seeing a very
short snapshot of the life, the death, of a star. And that’s why we don’t see many; though there are billions
of stars in the galaxy that die this way, this phase is very brief. Enjoy looking at them while you can. And
what of the Sun? Will it, one day in the distant future be at the center of a planetary nebula it expels as it
dies? Ehh probably not. When the Sun becomes a white dwarf, it most likely won’t be energetic enough
to make the surrounding gas glow; most planetaries start off as stars more massive and hotter than the
Sun. When our Sun dies, it’ll go quietly and without a lot of visible fanfare. Alien astronomers, if they’re
out there in 8 billion years, may not even notice. But more massive stars do make quite the spectacle. And
if they’re really massive, more than about 8 times the Sun’s mass, they really and truly make a scene
when they die. They explode. But that’s for next week. Mwuhahahaha. Today you learned that when low
mass stars die, they form white dwarfs: incredibly hot and dense objects roughly the size of Earth. They
also can form planetary nebulae: huge, intricately detailed objects created when the wind blown from the
dying stars is lit up by the central white dwarf. They only last a few millennia. The Sun probably won’t
form one, but higher mass stars do.
High Mass Stars

Stars are in a constant struggle between gravity trying to collapse them and their internal
heat trying to inflate them. For most of a star’s life, these two forces are at an uneasy truce. For a star
like the Sun, the balance tips in its twilight years. For a brief glorious moment it expands, but then blows
away its outer layers, leaving behind the gravitationally compressed core. It goes out with a whimper.
Well, a whimper from a two octillion ton barely constrained nuclear powered fireball. But more massive
stars aren’t quite as resigned to their fate. When they go out, they go out with a bang -- a very, very big
bang.

In the core of a star, pressure and temperature are high enough that atomic nuclei can get
squeezed together and fuse. This releases energy, and creates heavier elements. Hydrogen fusion makes
helium, helium fusion makes carbon, and each heavier element, in general, takes higher temperatures and
pressures to fuse. Lower-mass stars like the Sun stop at carbon. Once that builds up in the core, the star’s
fate is sealed. But if the star has more than about 8 times the Sun’s mass, it can create temperatures in its
core in excess of 500 million degrees Celsius, and then carbon will fuse. There are actually a lot of steps
in this process, but in the end you get carbon fusing into neon, magnesium, and some sodium. What
happens next hearkens back to what we found goes on in the Sun’s core as it ages: fuse an element, create
a heavier one, then that heavier one builds up until the core contracts and heats up enough to start fusing
it. So carbon fusion makes neon, magnesium, and sodium, and those accumulate. The core heats up, and
when it reaches about a billion degrees, neon will fuse. Neon fusion creates more magnesium, as well as
some oxygen. These build up in the core, it shrinks, heats up to about 1.5 billion degrees, and then oxygen
fuses, creating silicon. Then THAT builds up until the temperature hits about 2-3 billion degrees,
whereupon silicon can fuse. Among a pile of other elements, silicon fusion creates iron. And that’s
trouble. Big, big trouble. Once silicon fusion stars, the star is a ticking time bomb. But before we light
that fuse, let’s take a step back. What’s happening to the outer layers of the star? What do we see if we’re
outside, looking back at it Because the star was born massive, it spent its hydrogen fusing days as a blue
main sequence star. Stars like this are extremely luminous, and can be seen for tremendous distances.
Like the Sun, though, a massive star changes when hydrogen fusion stops, its core contracts, and then
helium fusion begins. It swells up just as the Sun will, but instead of becoming a red giant, it generates so
much energy it becomes a red supergiant.These are incredibly huge stars, some over a billion kilometers
across! And they are luminous. For example, Betelgeuse in Orion is a red supergiant, and one of the
brightest stars in the sky despite being over 600 light years away. From that distance, you’d need a decent
telescope to see the Sun at all. And that’s nothing compared to VY Canis Majoris, the largest known star,
which is a staggering two billion kilometers across. We even have a special term for it: a hypergiant. As
the core switches back and forth from one fusion reaction to the next the outer layers respond by
contracting and expanding, so a red supergiant can shrink and become a BLUE supergiant. Rigel, another
star in Orion, is a blue supergiant, putting out over 100,000 times as much energy as the Sun! OK, let’s go
back to the core. It now looks like an onion, with multiple layers: iron is building up in the center,
surrounded by fusing silicon. Outside that is a layer of fusing oxygen, then neon, then carbon, then
helium, and finally hydrogen. You might think massive stars would last longer because they have more
fuel than lower mass stars. But the cores of these monsters are far hotter, and fuse elements are far higher
rates, running out of fuel more quickly. A star like the Sun can happily fuse hydrogen into helium for over
10 billion years. But a star twice as massive as the Sun runs out of hydrogen in just 2 billion years. A star
with 8 times the Sun’s mass runs out in only 100 million years or so. And each step in the fusion process
happens faster than the one before it. In an extreme case, like for a star 20 times the mass of the Sun, it’ll
fuse helium for about a million years, carbon for about a thousand, and neon fusion will use up all its fuel
in a single year! Oxygen lasts for only a few months. Silicon fuses at a ridiculously high rate; the star will
go through its entire supply in — get this — a day. Yes, one day. The vast majority of a star’s life is spent
fusing hydrogen; the rest happens in a metaphorical blink of the eye. Silicon fuses into a bunch of
different elements, including iron. That inert iron builds up in the core, just like all those elements did
before, and just like before the iron core shrinks and heats up. But there’s a huge difference here. In all
the previous fusion stages, energy is created. That energy transforms into heat, and that helps support the
soul-crushing amount of stellar mass above the core. But iron is different. When it fuses it actually sucks
up energy instead of creating it. Instead of providing energy for the star, it removes it. This accelerates the
shrinking, compressing the core, heating it up even more. Even worse, at these temperatures and pressures
the iron nuclei suck up electrons that are whizzing around, which are also helping support the core. It’s a
double whammy; both major means of support for the star are removed in an instant — silicon fusing into
iron is happening so fast this literally takes a fraction of a second once it gets started. The core gets its
legs kicked out from under it. It doesn’t shrink, it collapses. The gravity of the core is so mind-bogglingly
strong that the outer parts crash down on the inner parts at a significant fraction of the speed of light. This
slams down on the central core, collapsing from several hundred kilometers across down to a couple of
dozen kilometers across in just a few thousandths of a second! The star is doomed. Because all hell is
about to break loose. Now, at this point, one of two things can happen. If the star has less than about 20
times the Sun’s mass, the core collapse stops when it’s still 20 or so kilometres wide. It forms what’s
called a neutron star, which I’ll cover in the next episode. If the star is more massive than this, then the
collapse cannot be stopped by any force in the Universe. The core collapses all the way down. Down to a
point. The gravity becomes so intense that not even light can escape. A black hole is born. We’ll cover
black holes in a future episode as well. But for now, what happens when the core collapses and suddenly
stops? The core of the star, whether it’s a neutron star or a black hole, is now extremely small with
terrifyingly strong gravity. It pulls on the star’s matter above it, HARD. This stuff comes crashing down
at a fantastic speed and gets hugely compressed, ferociously heating up. At the same time, two things
happen in the core. While this stuff is falling in, a monster shock wave created by the collapse of the core
moves outward, and slams into the incoming material. The explosive energy is so insane it slows that
material substantially. The second event is that the complicated quantum physics brewing in the core
generates vast numbers of subatomic particles called neutrinos. The total energy carried by these little
neutrinos is almost beyond reason: In a fraction of a second, they carry away 100 times as much energy as
the Sun will produce over its entire lifetime That’s an incredible amount of energy. Now, these little
beasties are seriously elusive and hate to interact with normal matter; one single neutrino could pass
through trillions of kilometers of lead without even noticing. But so many are created in the core collapse,
and the material barreling down on the core so dense, that a huge number of them are absorbed. This vast
wave of neutrinos slams into the oncoming material like a bullet train hitting a slice of warm butter. The
material stops its infall, reverses course, and blasts outward. The star explodes. It explodes. This is called
a supernova, and it is one of the most violent and terrifying events the Universe can offer. An entire star
tears itself to shreds, and the expanding gas blasts outward at 10% the speed of light. The energy released
is so huge they can be seen literally halfway across the Universe; they outshine all the stars in the rest of
the galaxy combined. The expanding material, called the supernova remnant, forms fantastic shapes. The
most famous is the Crab nebula, from a star we saw blow up in the year 1054. The tendrils form as the
material expands into the gas and dust that surrounded the progenitor star. As remnants expand and age
they become more tenuous. Some have bright rims as they push into material between the stars; others
form complex webs of filaments. I’m often asked if there are any stars near enough to hurt us when they
explode. The quick answer is no. Even though supernovae are incredibly violent, space is big. A
supernova would have to be at least as close as 100 light years from us before we start feeling any real
effects. The nearest star that might explode in this way is Spica, in Virgo, and it’s well over 100 light
years away. I say “might” explode, because it’s at the lower mass limit for going supernova. It might not
explode at all. Betelgeuse will certainly explode some day, but it's too far away to hurt us. We're pretty
safe from this particular threat. I’ll note that after all this, there IS another kind of supernova involving
white dwarfs, which we’ll cover in a future episode about binary stars. Happily, we’re probably safe from
them too. Breathe easy. As terrifying and dangerous as supernovae are, there’s a very important aspect of
them you need to know. Supernovae are capable of great destruction, but they’re also critical for our own
existence. When the star explodes, the gas gets so hot and is compressed so violently by the blast that it
undergoes fusion, what astronomers call explosive nucleosynthesis: Literally, creating heavy elements
explosively. New elements are produced in quantities that dwarf the Earth’s mass. Calcium, phosphorus,
nickel, more iron… all made in the hellish forge of the supernova heat, and flung outward into the
Universe. It takes millennia or longer, but this material mixes with the other gas and dust clouds floating
in space. Sometimes, these clouds will be actively forming stars — sometimes the collapse of the cloud to
form stars may even be triggered by the supernova slamming into it! Either way, the heavy elements
created in the supernova will become part of the next generation of stars and planets. Supernovae are how
the majority of heavy elements in the Universe are created and scattered. The calcium in your bones? The
iron in your blood? The phosphorus in your DNA? All created in the heart of the titanic death of a star.
That star blew up more than 5 billion years ago, but parts of it go on in you. Today you learned that
massive stars fuse heavier elements in their cores than lower mass stars. This leads to the creation of
heavier elements up to iron. Iron robs critical energy from the core, causing it to collapse. The shock
wave, together with a huge swarm of neutrinos, blast through the star’s outer layers, causing it to explode.
The resulting supernova creates even more heavy elements, scattering them through space. Also, happily,
we’re in no danger from a nearby supernova.
Neutron Stars

When an 8 – 20 solar mass star ends its life, it does so with a bang: a supernova. And when it’s all over,
there’s a couple of octillion tons of superheated plasma expanding away from the explosion site at a
fraction of the speed of light, a whole mess of energy released in the form of light and neutrinos, and a
bizarre little ball of quantum nastiness in the center composed almost entirely of neutrons. The properties
of this neutron star are almost as bizarre as things get in the Universe. And if it all seems rather alien to
you, well, that’s OK. For a little while, astronomers wondered if aliens really were behind what they were
seeing. Now I’m not sayin’ aliens. When we last left the core of a high mass star, it was in a bad way:
Milliseconds ago it was fusing silicon into iron, but now it’s collapsing under its own immensely
powerful gravity. The collapse takes a fraction of a second, but a lot happens in that fraction of a second.
In lower mass stars, the core supports itself via electron degeneracy pressure, the result of a rule in
quantum mechanics that says electrons vehemently resist being squeezed together. But even electron
degeneracy fails to stop the collapse if the core has a mass more than about 1.4 times the mass of the Sun.
That’s just too much of a load to bear, and the collapse continues. Under these huge pressures, a funny
thing happens: Protons, electrons, and other subatomic particles get smashed together, and they merge to
form neutrons. And this happens to almost ALL of them. When the core collapses down to about 20 km
in diameter, it’s essentially a ball of neutrons with some protons and electrons here and there that
survived, and a crust of normal but highly compressed matter on top. When this happens yet another
effect comes into play: neutron degeneracy. Like electrons, neutrons resist being squeezed too tightly
together, but this time the strength of the pressure is far, far stronger than for electrons. If the core is less
than about 2.8 times the Sun’s mass, the collapse runs into a wall. It stops. This generates a huge shock
wave, which, along with a flood of energetic subatomic particles called neutrinos, blasts outwards,
blowing up the star. What’s left of the core after the metaphorical smoke clears is a neutron star, one of
the most bizarre objects in the Universe. Such a star would be extremely weird. Or really, just
EXTREME. Its mass would be more than that of the entire Sun, all packed into a sphere maybe 20 km
across. Now let’s just stop there for a sec and let that sink in. The Sun has a mass 300,000 times the Earth.
Imagine packing that all into a ball THE SIZE OF A SMALL CITY. Too mind boggling? OK, think of it
this way: You are mostly empty space. Every atom in your body has a nucleus made of tightly packed
neutrons and protons, and electrons whizzing around outside them. If you could magnify an atom to be
100 meters across, the nucleus would be roughly the size of a marble. Imagine all that empty space
between the nucleus and the electrons. That’s a normal atom. But in a neutron star, ALL OF THAT
SPACE would be filled with neutrons. All of it. Every nook and cranny inside the neutron star has matter
in it, all the way down to the scale of an atomic nucleus. This is what gives a neutron star its mind-
crushing properties. I’m now going to barrage you with very large numbers. So, take a deep breath, and
you might wanna sit down. Neutron stars are ridiculously dense. A single cubic centimeter of neutronium,
as neutron star stuff is usually called, has a mass of about 400 MILLION tons. Want some perspective on
that number? Well, very roughly, that’s the total mass of every single car and truck in the United States.
Imagine a couple of hundred million vehicles, crushed down until they could all fit inside this six-sided
die. That’s neutronium. It’s so dense that, as far as it’s concerned, normal matter is a slightly polluted
vacuum. If you set it on the ground it would fall right through the Earth. Now, anything that dense has a
HUGE gravitational pull. If you were on the surface of a neutron star… well, you’d be very dead,
obviously -- like immediately, flattened down to a thickness of just a few atoms. And that’s because a
typical neutron star has a surface gravity 100 BILLION times stronger than Earth’s. I have a mass of
about 77 kilos, and here on Earth I weigh about 170 pounds. On a neutron star, I’d weigh 17 trillion
pounds. That’s 23,000 times the weight of the Empire State Building. But wait! There’s more! In our
introduction to the solar system I mentioned that when you take a spinning object and shrink it the spin
will increase—the usual example is an ice skater drawing in his arms, increasing his rotation until he’s a
blur. The same is true for the star’s core whenit collapses into a neutron star. It may have had a very slow
spin before the supernova, maybe even taking weeks to spin once. But when it shrinks down to just 20 km
across and becomes a neutron star, that rotation will increase by a HUGE factor. A freshly minted neutron
star might spin several times per SECOND. The magnetic field increases as well. A star like the Sun has
an overall magnetic strength not too different from the Earth’s. But when that core collapses, the strength
of the field skyrockets, and a neutron star can easily have a magnetic field several trillion times stronger
than the Sun’s. That’s strong enough to erase your credit card from a hundred thousand kilometers away.
See? Ridiculous. All of these properties are brain-melting. But are they real? Could an object like this
really exist? Oh my, yes. The first neutron star was detected in 1965, though not recognized for what it
was at the time. A couple of years later another one was found, and this time was correctly identified as a
neutron star. But then things got...weird. In 1967, Jocelyn Bell was a graduate student helping build a
adio telescope. There was a persistent noise in their data they couldn’t seem to fix. Bell studied it night
after night, finally figuring out that the pattern wasn’t a problem with their data, it was from an actual
astronomical object. She had discovered the first known pulsar. What’s a pulsar, you ask? Pulsars are
neutron stars. In a nutshell, their rapid rotation coupled with their incredibly strong magnetic fields launch
twin beams of energy away from the star, like the beams from a lighthouse. The beams sweeps around as
the star rotates, and from Earth we see this as a pulse, a blip, of increased brightness. This pulse can be
detected in visible light, radio waves, and even X-rays! The spin of a neutron star is amazingly
stable,making these pulses act like a very accurately timed cosmic clock. In 1967, no one could believe a
natural object could do this, and this object was half-jokingly given the name LGM-1. Little Green Men
1. Now we know of over a thousand pulsars in just our galaxy alone, and we know they are the leftover
cores of massive stars that exploded. Some spin with periods many seconds or even minutes long. Some
are in binary systems; another normal star orbits them. If they’re close enough together the neutron star
can rip material off the other star and feed on it. This increases the pulsar’s spin, and we know of a few
that have incredibly rapid rotation rates; some spin hundreds of times per second! These are called
millisecond pulsars, and if they spun much faster the centrifugal force would rip them apart despite their
tremendous gravity! Even after a thousand years, a pulsar can still be a force to reckon with. There’s a
pulsar in the center of the Crab Nebula, the remains of a star that exploded to create that supernova
remnant. A substantial fraction of the light emitted from the nebula is powered by the pulsar itself; its
fierce output energizes the nebula, causing it to glow brightly even after a millennium. I’m telling ya,
thinking about neutrons stars makes the hair on the back of my neck stand up. And I haven’t even
mentioned magnetars yet. Neutron stars are more than just weird little balls of neutrons. They have a
crust, probably a few centimeters thick, made of highly compressed but more or less normal matter,
squeezed into a kind of highly stiffened crystal state. The magnetic field of the star penetrates this
crystalline crust and stretches out for quite a distance. In some neutron stars the magnetic field is even
stronger than usual, and can be — get this — a QUADRILLION times stronger than the Sun’s. These
über-powerful neutron stars are given the name magnetars. They’re relatively rare; maybe 10% of all
neutron stars are born as magnetars. And they have short lifetimes; the magnetic field is so strong it acts
like the brakes on a car, slowing the neutron star’s spin. That spin helps power the magnetic field, so the
field weakens as the star slows. But while they're around, magnetars are the most magnetic objects in the
Universe. And with great power comes great responsibility…if your responsibility is to be one of the
scariest objects in the Universe. Why? In a neutron star, the crust and magnetic field are locked together,
so a change in one affects the other. The crust of the star is under incredible strain due to the intense
gravity and rapid rotation. If the structure slips, it can snap, creating a star quake — like an earthquake,
but just a WEE bit stronger. In an earthquake, huge amounts of energy are released when the Earth’s crust

shifts and snaps, enough to destroy buildings and quite literally move mountains. But in a neutron star this
effect is multiplied. Hugely. Remember, the crust is phenomenally dense, and the gravity is enormous. If
the crust strains and snaps, dropping just a single centimeter, the resulting release of energy is vast
beyond imagining. This energy is released as a tremendous explosion in the crust, shaking it. This also
shakes the magnetic field, which reacts…poorly. When the Sun’s magnetic field throws a tantrum, we get
a solar flare, which can be as powerful as billions of nuclear bombs. A magnetar flare dwarfs that into
insignificance. It can be trillions of times stronger than a solar flare — in a fraction of a second, a
magnetar can release as much energy as the Sun gives of in a quarter of a million years. In 2004,
astronomers were stunned when a huge blast of X-rays slammed into orbiting satellites. One of these
satellites, named Swift, actually had its detectors saturated with X-rays, even though it wasn’t even
pointing at the source at the time! The X-rays literally came right through the side of the satellite with
such ntensity that Swift — which was designed to detect powerful X-ray sources — was momentarily

blinded by them. The source of this X-ray burst was quickly determined to be a magnetar called SGR-
1806-20, and the effects were incredible: It actually compressed the Earth’s magnetic field, and partially
ionized the Earth’s upper atmosphere. Oh, did I mention that this magnetar is 50,000 light years away?
That’s halfway across the galaxy. That’s INCREDIBLE. At a distance of 500 quadrillion km its effects
were felt more strongly than a powerful flare from the Sun! The good news is that there are very few
magnetars like this in the galaxy, probably fewer than a dozen. Also, catastrophic explosions like the one
in 2004 are rare; if one had had happened any other time in the past 40 or so years we would’ve detected
it. And frankly, it’s really cool that we had astronomical satellites orbiting the Earth which could study it!

We’ve come a long way in understanding neutron stars since they were first discovered, but there’s much
about them we don’t understand. Every time we find out more we find out they’re even weirder than we
first thought. And yet, for all that, they’re not the weirdest things in the sky. Not by a long shot. That
place is held pretty securely by the other type of object created in a supernova: a black hole. Stay tuned.
Today you learned that when a star between 8 and 20 times the Sun’s mass explodes, the core collapses to
form a neutron star. Neutrons stars are incredibly dense, spin rapidly, and have very strong magnetic
fields. Some of them we see as pulsars, flashing in brightness as they spin. Neutrons stars with the
strongest magnetic fields are called magnetars, and are capable of colossal bursts of energy that can be
detected over vast distances.
Black Holes

As we’ve seen over the past few episodes, a lot of really epic stuff happens when a star dies. If the star’s
core is less than 1.4 times the mass of the Sun, it becomes a white dwarf—a very hot ball of super-
compressed matter about the size of the Earth. If the core is heftier, between 1.4 and 2.8 times the Sun’s
mass, it collapses even further, becoming a neutron star that’s only 20 km across. The neutron soup inside
of it resists the collapse, and prevents the core from shrinking any more. But what if the mass is MORE
than 2.8 times the Sun’s? If that happens, the gravity of the core can actually overcome the tremendous
resistance of the neutrons and continue its collapse. What force can possibly stop it now? It turns out,
none. None more force. There is literally nothing in the Universe that can stop the collapse. The core of
the star is about to go bye bye. Way back in Episode 7 I talked about escape velocity, and it’s about to
become a major player in the unfolding events of the collapsing core of a high mass star. In brief, it’s the
velocity at which you need to fling something off the surface of an object to get it to escape. For the
Earth, the escape velocity is about 11 km/sec. Get something moving that quickly, and it’s gone; it’ll
never fall back. The Sun, which has much stronger gravity than Earth, has an escape velocity of over 600
km/sec. A neutron star, with its immense gravity, can have an escape velocity of 150,000 km/sec – that’s
half the speed of light! Keep that in mind, and let’s go back to the collapsing core of the star. As it
shrinks, its gravity gets stronger and stronger. That means its escape velocity gets higher and higher.
When it’s neutron star-sized the escape velocity is half the speed of light, but if it’s more than 2.8 times
the mass of the Sun, the core will keep collapsing. When its size drops just a little bit more, down to
roughly 18 km, an amazing thing happens: The escape velocity at its surface is equal to the speed of light.
And, well, that’s a problem, because in our Universe, nothing can travel faster than the speed of light. Not
a rock, not a rocket, not even light itself. Once the core of the star shrinks down smaller than that magic
size, nothing can escape. No matter can come out, so it’s like an infinitely deep HOLE, and no light can
come out, so it’s BLACK. We should come up with a snappy name for such an object. A black hole is the
ultimate end state for the core of a high mass star. Whatever happens in a black hole STAYS in a black
hole. That region of space, that surface around the black hole where the escape velocity is the speed of
light, is called the EVENT HORIZON for that reason. Any event that happens inside can’t be known. It’s
beyond the horizon for us. Black holes mess with our concepts of space and time. The math and physics
of black holes is incredibly complex, so much so that even after several decades of study, physicists still
argue over a lot of their properties. This has led to a lot of misconceptions about them, too. All right, let’s
get this out of the way right now: The Sun cannot become a black hole. It takes a stellar core at least
about three times the mass of the Sun to overcome neutron degeneracy pressure. That means the original
star must have something like 20 times the Sun’s mass or more. So we’re safe from THAT particular scifi
scenario. Here’s another misconception: A lot of people think of black holes as cosmic vacuum cleaners,
sucking in everything near them. But that’s not really true. They have powerful gravity, yeah, but only
when you’re very close to one. The power of a black hole comes from its mass, certainly, but just as
important is its SIZE. Or, really, its LACK of size. If you could turn the Sun into a black hole, which you
can’t, but let’s pretend you could, then the Earth would orbit it pretty much exactly as it does now. From
150 million kilometers away, the Earth doesn’t care if the Sun is big or tiny. We’re so far away that it
doesn’t matter. It gets to be a big deal when you get close. Remember, from episode 7 about gravity, the
strength of gravity you feel from an object depends on how massive it is and your distance from its center.
The closest you can get to the Sun is by touching it, being on its surface, about 700,000 km from its
center. If you get any closer to its center, you’re INSIDE it. The material OUTSIDE of your position is no
longer pulling you down and so the gravity you feel will actually decrease. But if the Sun were crushed
down to about 6 km across it would be a black hole. You could get much closer than 700,000 km to it,
and as you did you’d feel a stronger and stronger pull as you approached it. So from far away, a black
hole with, say, ten times the Sun’s mass would pull on you just as hard as a normal star with that same

mass. You can orbit a black hole, too, as long as you keep a safe distance between you and it. Orbiting a
ten-solar-mass black hole would be just like orbiting a ten-solar-mass star… except not so hot and bright.

Black holes are weird enough without the misconceptions. Black holes also come in different sizes. The
kind I’ve been talking about has a minimum mass of about 3 times the Sun’s, and might get as high as a
dozen or more times the Sun’s mass, if the parent star was big enough. We call these stellar-mass black
holes. If it happens to gobble down more matter, it gets more massive, and the event horizon grows as
well. The black hole gets bigger. The idea that huge black holes could form in the centers of galaxies was
first proposed in the 1970s, and it wasn’t much later that the first one was found, in the center of our own
Milky Way galaxy. We’ve measured its mass at a whopping 4.3 million times the Sun’s mass! And now
we think every major galaxy has one at its heart, too, and in fact may be crucial in the formation of
galaxies themselves. I’ll discuss those more in a future episode. Here’s a fun thought: What would happen

if you fell into one? Say, a stellar black hole with ten times the Sun’s mass? You’d die. But what happens
in the few milliseconds before you left the known Universe forever is actually pretty interesting. As
we’ve seen many times in our own solar system, tides are important. They arise because gravity weakens
with distance, so a big object like a moon gets stretched by its planet’s gravity; the far side of the moon is
pulled less than the near side. A black hole has incredibly intense gravity, so the tides it can inflict are
serious indeed. They’re so strong that if you fell into a stellar mass black hole feet first, the force of
gravity on your feet can be MILLIONS OF TIMES STRONGER than the force on your head. Remember,
even the meager tides of a planet can rip moons apart. When you multiply that force by a million, you’re
in trouble. As you fall in, your feet are pulled so much harder than your head that you stretch, pulled like
taffy. You’d become a long, thin, noodle, kilometers in length, but narrower than a hair wide.
Astronomers call this – and no, I’m not kidding – spaghettification. This would happen pretty close to the
black hole, just a few dozen kilometers out. If you fell in from a long distance, you’d be moving pretty
near the speed of light by that point, and you’d only have a millisecond or so before it killed you anyway,
so yay? Note that this is only for stellar mass black holes. Supermassive black holes are far bigger,
millions or billions of kilometers across. Compared to that size, the distance between your head and feet
is small, so the tides across you aren’t nearly as severe. You’d fall in pretty much intact -- if that makes
you feel any better. But compared to either flavor of black hole, a star still has substantial size, and one
that gets too close to any black hole can be disrupted via tides. In March 2011, astronomers witnessed just
such an event. In a distant galaxy, a star apparently got too close to a black hole, and was torn apart by the
ferocious tides. As the star was disrupted, it flared in brightness, momentarily blasting out a trillion times
the Sun’s energy! That’s how we were able to see it even though it was several billion light years away.
But I’ve saved the weirdest thing for last One of Albert Einstein’s biggest ideas is that space isn’t just
emptiness, it’s an actual thing, like a fabric in which all matter and energy is embedded. What we
perceive as gravity is really just a warping of this space, like the way a bowling ball on top of a bed warps
the shape of the mattress. The more massive an object, the more it warps space. Not only that, but space
and time are basically two parts of the same thing, what we now call space-time. You can’t affect one
without affecting the other. Einstein calculated that when a massive object warps space, it also warps
time; someone deep inside the gravitational influence of an object perceives time as ticking more slowly
than someone far away from that object. I know, it’s bizarre; we think of time as just… flowing, and
everyone should see it move at the same rate. But the Universe is under no obligation to obey our
preconceptions. Einstein was right (he was right a lot). This slowing of time is stronger the stronger the
gravity of the object is. So your clock ticks a bit slower than someone far away from Earth, for example.
The effect is tiny, but real, and we’ve actually measured it on Earth with extremely precise clocks!
However, if you get near a black hole, the effect gets a lot stronger. In fact, black holes warp space-time
so much that, at the event horizon, time essentially stops! You’d see your clock running normally, and
you’d just fall in — bloop, gone. But someone far away would see your clock ticking more slowly as you
fell in. And this isn’t a mechanical or perception effect; it’s actually woven into the fabric of space. To
someone outside looking down on you, your fall would literally take forever. But then, they wouldn’t be
able to actually see you. The light you emit would have to fight the intense gravity of the black hole to get
out, and to do that it would lose energy. This is very similar to the Doppler redshift I’ve talked about in
earlier episodes, and is called a gravitational redshift. When you’re right at the event horizon, just when
an outside observer would see your clock stop, they’d also see the light coming from you infinitely
redshift! Your light would lose ALL its energy trying to leave the vicinity of the black hole, and you’d be
invisible. And from your viewpoint? Buckle up, because this is...WOW. You’d see the universe speed up,
and just as you hit the event horizon, all of time would pass — all of it. And all that light coming at you
from the Universe would be blue-shifted, becoming such high energy that you’d be fried. But since you’re
about to fall into a black hole, you probably wouldn’t care. See? Like I said…WOW. Black holes are so
strange, with such fiercely complicated math and physics to explain them, that scientists are still trying to
figure out even basic things about them. For example, some scientists argue that the event horizon as we
understand it may not actually exist, and that when you apply quantum mechanics to black hole physics,
you find particles can slowly leak out. We’re still new at this, and struggling to understand what may be
the most complex objects in the cosmos. Black holes, as bizarre and counterintuitive as they are, keep
popping up from here on out as we poke our noses into more and bigger astronomical objects. While they
may seem scary and weird — and let’s be honest: they are — they have literally shaped most of the
objects we see in the Universe. Today you learned that stellar mass black holes form when a very massive
star dies, and its core collapses. The core has to be more than about 2.8 times the Sun’s mass to form a
black hole. Black holes come in different sizes, but for all of them, the escape velocity is greater than the
speed of light, so nothing can escape, not matter or light. They don’t wander the Universe gobbling
everything down around them; their gravity is only really intense very close to them. Tides near a stellar
mass black hole will spaghettify you, and time slows down when you get near a black hole — not that this
helps much if you’re falling in.

You might also like