Sei sulla pagina 1di 46

BIOLOGY

The Nervous System

Pretty much every animal, except for some really simple ones, have nervous systems, which is great, because it's what lets
things do things like: have behaviors. It makes you the sentient, living thing that you are. The whole set-up here: your
brain, your nerves, your spinal cord, everything is made up of specialized cells that you don't find anywhere else in the
body. Most of those are neurons, which, you've seen them before, they look kind of like a tree with roots, a trunk and
branches. Neurons bundle together to form nerves, pathways that transmit electrochemical signals from one part of your
body to another. So, when you bite into a piece of pizza- The receptor neurons in my taste buds recognize I'm eating
something salty and fatty and awesome. And they carry that information along a nerve pathway to my brain. And then my
brain can be like "Yeah! Pizza!" and then it can respond by sending back information through different nerve pathways
that say: "You should eat more of that pizza!" And despite what my brain is telling me, I'm going to try to not eat any more
of that pizza. You wouldn't think that it's terribly complicated to know that pizza tastes good and to tell someone to eat
more pizza. But it turns out that our brains and our nervous systems are crazy complicated. Your nervous system basically
has a big old bureaucracy of neurons, and it's divided into two main departments: the central nervous system and the
peripheral nervous system.

Central and peripheral.

The central nervous system, basically your brain and your spinal cord, is responsible for analyzing and interpreting all
those data that your peripheral nervous system, all of the nerves outside of your brain and spine, collects and sends its
way. Once the central nervous system makes a decision about data, it sends a signal back through to the peripheral
nervous system saying "Do THIS thing!" Which the peripheral nervous system then does. Both of these systems contain
two different types of neurons: afferent and efferent. Afferent and efferent are biological terms, and they're horribly
confusing, and I apologize on behalf of the entire institution of biology for them.

Afferent systems carry things to a central point, and efferent systems carry things away from a central point.
So afferent neurons carry information to the brain and spinal cord for analysis.
In the peripheral nervous system, afferent neurons are called sensory neurons, and they're activated by external stimuli
like the complex and glorious flavor of pizza and then they convert those data into a signal for the central system to
process.

The central nervous system has afferent neurons too, and there they bring information into special parts of the brain, like
the part of the brain that goes, "Mmmmmm, salty!" Efferent neurons carry information out of the center. In the
peripheral nervous system, they're called motor neurons because many of them carry information from the brain or
spinal cord to muscles to make us move, but they also go to pretty much every other organ in your body, thus making
them, like, work and do stuff to keep you alive. In the central system, efferent neurons carry information from special
parts of the brain to other parts of the brain or spinal cord. Of course if it ended there, it would be way too simple and no
good bureaucracy just has two departments.

So the peripheral nervous system is actually made up of two different systems with two very different jobs: the somatic
nervous system and the autonomic nervous system. The somatic system controls all the stuff you think about doing like all
the information coming through your senses, and the movement of your body when you want it to make movements.
But here's something interesting: Since we're totally in love with our brains as sort of the center of all being, of ourselves,
we think that all the information about everything going on in our bodies goes to our brains for some kind of decision.
Not so!
Sometimes, like when we touch a hot stove, the afferent neurons carry the signal "HOT!" to the central nervous system,
but that information doesn't even ever get to the brain the spinal cord actually makes that decision before it gets to the
brain, sends a message directly back to the muscle saying, "Get your hand off the freakin stove, *******!" This bit of
fancy nerve-work lets the spinal cord make decisions rather than the brain, it's called the reflex loop. So, the other branch
of the peripheral nervous system, the autonomic system, carries signals from the central nervous system that drive all of
the things your body does without thinking about them: your heartbeat, your digestion, breathing, saliva production, all
your organ functions. But we're not done yet here. We need to go deeper.

The autonomic nervous system has two divisions of its own: the sympathetic and parasympathetic. And the jobs that
these two perform aren't just different they're completely opposite, and frankly, they're always vying for control of the
body in some kind of nervous system cage match. The sympathetic division is responsible for, like, freaking out. You've
probably heard this talked about as the fight-or-flight response. In other words, stress.
But stress isn't all bad: it's what saves our lives when we're being chased by saber toothed tigers, right? The sympathetic
system prepares our body for action by increasing the heart rate and blood pressure, enhancing our sense of smell,
dilating the pupils, activating our adrenal cortex to make adrenaline, shutting down blood supply to our digestive and
reproductive systems so there will be more blood available for our lungs and muscles when we have to, like, RUN!
Even though you're not in a constant state of panic at least, I hope not, I kind of am that system is running all the time,
every day.

But right next to it is the parasympathetic division, working hard to make sure we take it nice and easy. It dials down heart
rate and blood pressure, constricts our lungs, makes our nose run, increases blood flow to our reproductive junk, our
mouths produce saliva, encourage us to poop and pee. It's basically what we have to thank for taking a nap, sitting in front
of the TV, going to the bathroom and getting it on. So, consider yourself lucky you've got both the stress response and the
chill-the-heck out response, working side-by-side because together they create a balance, or a homeostasis.
Now, that's what the nervous system does.

Next we have to talk about how it does it.


The neurons that make up our nervous systems make it possible for our bodies to have their very own little electric
systems. So to understand how they work you have to understand their anatomy. Like I said before, a typical neuron has
branches like a tree. These are called dendrites, and they receive information from other neurons. Neurons also have a
single axon the trunk of the tree which is branched at the end and transmits signals to other neurons. The axon is also
covered in fatty material called myelin, which acts as insulation. But the myelin sheath isn't continuous, there are these
little bits of exposed neuron along the axon, which have the sweetest names in this whole episode they're called the
Nodes of Ranvier. Which seems like an excellent working title for the 8th Harry Potter novel. Harry Potter and the Nodes
of Ranvier. Anyway, these nodes allow signals to hop from node to node, which lets the signal travel down a nerve faster.
This node-hopping, by the way, has a name. It's called saltatory conduction.
Conduction because it's electrical conduction and saltatory because saltatory means leaping. Finally, the place where an
axon's branches come in contact with the next cell's dendrite is called a synapse, and that's where neurotransmitters pass
information from one neuron to the next.

Now, think back to, or just go watch the episode we did on cell membranes, where we talked about how materials travel
down concentration gradients. Well, in much the same way, all neurons in your body have a membrane potential, a
difference in voltage, or electrical charge, between the inside and the outside of the membrane. You might also
remember that this buildup of voltage is handled in part by a sexy little protein called the sodium-potassium pump.
Basically, the pump creates a voltage differential, like charging a battery, by moving 3 positively charged sodium ions out
for every 2 potassium ions it lets in, creating a net negative charge inside the cell relative to the outside. When a neuron is
inactive, this is called its resting potential, and its voltage is about -70 millivolts. But in addition to the pumps, neurons
also have ion channels. These are proteins that straddle the membrane, but they're a lot simpler and don't need ATP to
power them.

Each cell can have more than 300 different kinds of ion channels, each tailored to accept a specific ion. Now, don't zone
out here, because all of this stuff has got to come into play when a neuron becomes active. This happens when an input
or stimulus creates a change in the neuron that eventually reaches the axon, creating what's called an action potential a
brief event where the electrical potential of a cell rapidly rises and falls. When action potential begins, like when a
molecule of sugar touches one of my sweet tastebuds, some ion channels open and let those positive sodium ions rush in,
so that the inside starts to become less negative. With enough stimulus, the internal charge of the neuron reaches a
certain threshold, which triggers more sodium channels to respond and open the flood gates to let even more ions in.
That's happening on one tiny little area of the neuron. But this change in voltage creeps over to the next bunch of sodium
channels, which are also sensitive to voltage, and so they open. That exchange triggers the next batch, and the next batch,
and so on down the line. So this signal of changing voltage travels down the neuron's membrane like a wave.
But remember, the myelin sheath insulates most of the neuron, and just leaves those little nodes exposed, so instead of
being a steady wave, the wave jumps from node to node, speeding up the travel time of action potential down a neuron:
That's your saltatory conduction at work! When the wave reaches the end of the neuron, it triggers the release of
neurotransmitters from the neuron through exocytosis, and those neurotransmitters then float across the synapse to the
next neuron where they trigger another action potential over there.

Now, by this time, so many sodium ions have gotten inside the first neuron that the difference between the outside and
the inside is actually reversed: The inside is positive and the outside negative. And it seems like neurons hate that more
than pretty much anything else, so it fixes itself. The sodium channels close and potassium channels open up. The positive
potassium ions rush down both the concentration and electrochemical gradients to get the heck out of the cell. That
brings the charge inside the cell back down to negative on the inside, and positive on the outside. Notice, though, that
now the sodium is on the inside of the cell and the potassium is on the outside they're in the opposite places of where
they started. So, the sodium-potassium pumps get back to work and burn some ATP to pump the sodium back out and
the potassium back in, and phew! Things are now back at the resting potential again.
So, that, my friends, is how action potential allows neurons to communicate signals down a whole chain of neurons from
the outer reaches of the peripheral nervous system, all the way up the spinal cord and to the brain, and then back out
agian.
So, let's zoom out, and look at the broad view here. All my tastebuds have neurons in them. Each of my taste buds
contains between 50-100 specialized taste receptor neurons. Chemicals from this beautiful pizza dissolve in the saliva and
then stimulate the dendrites on the afferent neurons. This generates a bunch of action potentials that travel from the
afferent neurons in my tongue all the way to my brain, which is like, "My goodness, I think that's pizza!

Let's have another bite!"


The brain then sends messages through the efferent nerve pathways to do all sorts of things:
1. Chew, which involves constricting the muscles in my jaw over and over again.
2. Lower my head down to catch another bite, which involves moving all kinds of neck muscles.
3. Swallowing, which involves constricting the muscles in my throat and esophagus.
4. Opening my mouth again to receive another bite.
That signal is also going to my jaw. And that's not even to mention what's going to go on with the digestion of this bad
boy, driven by the autonomic nervous system.

Circulatory and Respiratory

All members of the kingdom Animalia need oxygen to make energy. Oxygen is compulsory. Without oxygen, we die. But as
you know, the byproduct of the process that keeps us all alive, cellular respiration, is carbon dioxide, or CO2, and it
doesn't do our bodies a bit of good, so not only do we need to take in the oxygen, we also have to get rid of the CO2.
And that's why we have the respiratory and circulatory systems to bring in oxygen from the air with our lungs, circulate it
to all of our cells with our heart and arteries, collect the CO2 that we don't need with our veins, and dispose of it with the
lungs when we exhale. Now, when you think of the respiratory system, the first thing that you probably think of is the
lungs.
But some animals can take in oxygen without lungs, by a process called simple diffusion, which allows gases to move into
and pass through wet membranes. For instance, arthropods have little pores all over their bodies that just sort of let
oxygen wander into their body, where it's absorbed by special respiratory structures.
Amphibians can take in oxygen through their skin, although they also have either lungs or gills to help them respire,
because getting all your oxygen by way of diffusion takes freaking forever.
So why do we have to have these stupid lung things instead of just using simple diffusion? Well, a couple of reasons.
For starters, the bigger the animal, the more oxygen it needs. And a lot of mammals are pretty big, so we have to actively
force air into our lungs in order to get enough oxygen to run our bodies. Also mammals and birds are warm blooded,
which means they have to regulate their body temperatures, and that takes many, many calories, and burning those
calories requires lots of oxygen.

Finally, in order for oxygen to pass through a membrane, the membrane has to be wet, so for a newt to take oxygen in
through its skin, the skin has to be moist all the time, which, you know, for a newt, isn't a big deal, but, you know, I don't
particularly want to be constantly moist, do you? Fish need oxygen, too, of course, but they absorb oxygen that's already
dissolved in the water through their gills. If you've ever seen a fish gill, you'll remember that they're just sort of a bunch of
filaments of tissue layered together. This gill tissue extracts dissolved oxygen and excretes the carbon dioxide. Still, there
are some fish that have lungs like Lungfish, which we call Lungfish because they have lungs. And that's actually where
lungs first appeared in the animal kingdom. All animals from reptiles on up respire with lungs deep in their bodies basically
right behind the heart. So while us more complex animals can't use diffusion to get oxygen directly, our lungs can.

Lungs are chock full of oxygen-dissolving membranes that are kept moist with mucus. Moist with Mucus... another great
band name. The key to these bad boys is that lungs have a TON of surface area, so they can absorb a lot of oxygen at
once. You wouldn't know from looking at them, but human lungs contain about 75 square meters of oxygen-dissolving
membrane. That's bigger than the roof of my house! And the simple diffusion that your lungs use is pretty freakin' simple.

You and I breathe oxygen in through our nose and mouth. It passes down a pipe called your larynx which then splits off
from your esophagus and turns into your trachea, which then branches to form two bronchi, one of which goes into each
lung. These bronchi branch off again, forming narrower and narrower tubes called bronchioles. These bronchioles
eventually end in tiny sacs called alveoli. Each alveolus is about a fifth of a millimeter in diameter, but each of us has about
300 million of them, and this, friends, is where the magic happens. Alveoli are little bags of thin, moist membranes, and
they're totally covered in tiny, narrow blood- carrying capillaries.
Oxygen dissolves through the membrane and is absorbed by the blood in these capillaries, which then goes off through
the circulatory system to make cells all over your body happy and healthy. But while the alveoli are handing over the
oxygen, the capillaries are switching it out for carbon dioxide that the circulatory system just picked up from all over the
body. So the alveoli and capillaries basically just swap one gas for another.

From there, the alveoli takes that CO2 and squeezes it out through the bronchioles, the bronchi, the trachea, and finally
out of your nose and/or mouth. So inhale for me once! Congratulations! Oxygen is now in your bloodstream! Now exhale!
Wonderful! The Co2 has now left the building! And you don't even have to think about it, so you can think about
something more important like how many Cheetos you could realistically fit into your mouth at the same time!
So, now you're all, "Yeah, that's great Hank, but how do lungs actually work? Like how do they do the thing where they do
where they get moved to come in and out and stuff?" Well, eloquent question! Well asked!

Lungs work like a pump, but they don't actually have any muscles in them that cause them to contract and expand.
For that we have this big, flat layer of muscles that sits right underneath the lungs called the thoracic diaphragm At the
end of an exhalation, your diaphragm is relaxed, so picture an arch pushing up on the bottom of your lungs and crowding
them out so that they don't have very much volume. But when you breathe in, the diaphragm contracts and flattens out,
allowing the lungs to open up. And as we know from physics, as the volume of a container grows larger, the pressure
inside it goes down. And the fluids, including air, always flow down their pressure gradient, from high pressure to low
pressure.
So as the pressure in our lungs goes down, air flows into them. When the diaphragm relaxes, the pressure inside the lungs
becomes higher than the air outside, and the deoxygenated air rushes out. And THAT is breathing! Now, it just so happens
that the circulatory system works on a pumping mechanism just like the respiratory system. Except, instead of moving air
into and out of the lungs, it moves blood into and out of the lungs.

The circulatory system moves oxygenated blood out of the lungs to the places in your body that needs it and then brings
the deoxygenated blood back to your lungs. And maybe you're thinking, "Whoa, what about the heart?! Isn't the heart the
whole point of the circulatory system?" Well settle down! I'm going to explain. We're used to talking about the heart as
the head honcho of the circulatory system. And yeah, you would be in serious trouble if you didn't have a heart! But the
heart's job is to basically power the circulatory system, move the blood all around your body and get it back to the lungs
so that it can pick up more oxygen and get rid of the CO2. As a result, the circulatory system of mammals essentially
makes a figure-8: Oxygenated blood is pumped from the heart to the rest of the body, and then when it makes its way
back to the heart again, it's then pumped on a shorter circuit to the lungs to pick up more oxygen and unload CO2 before
it goes back to the heart and starts the whole cycle over again.
So even though the heart does all the heavy lifting in the circulatory system, the lungs are the home base for the red
blood cells, the postal workers that carry the oxygen and CO2. Now, the way that your circulatory system moves the blood
around is pretty nifty.

Remember when I was talking about air moving from high pressure to low pressure? Well, so does blood. A four
chambered heart, which is just one big honkin' beast of a muscle, is set up so that one chamber, the left ventricle, has
very high pressure. In fact, the reason it seems like the heart is situated a little bit to the left of center is because the left
ventricle is so freaking enormous and muscley. It has to be that way in order to keep the pressure high enough that the
oxygenated blood will get out of there.

From the left ventricle, the blood moves through the aorta, a giant tube, and then through the arteries, blood vessels that
carry blood away from the heart, to the rest of the body. Arteries are muscular and thick- walled to maintain high
pressure as the blood travels along. As arteries branch off to go to different places, they form smaller arterioles and finally
very fine little capillary beds, which, through their huge surface area, facilitate the delivery of oxygen to all of the cells in
the body that need it. Now the capillary beds are also where blood picks up CO2, so from there the blood keeps moving
down the pressure gradient through a series of veins. These do the opposite of what the arteries did: instead of splitting
off from each other to become smaller and smaller, little ones flow together to make bigger and bigger veins to carry the
deoxygenated blood back to the heart. The big difference between most veins and most arteries is that instead of being
thick-walled and squeezy, veins have thinner walls, and have valves that keep the blood from flowing backwards. Which
would be bad. This is necessary because the pressure in the circulatory system keeps dropping lower and lower, until the
blood flows into two major veins: The first is the inferior vena cava, which runs pretty much down the center of the body
and handles blood coming from the lower part of your body.

The second is the superior vena cava, which sits on top of the heart and collects the blood from the upper body. Together
they run into the right atrium of the heart, which is the point of the lowest pressure in the circulatory system. So, all this
deoxygenated blood is now back in the heart. And it needs to sop up some more oxygen, so it flows into the right
ventricle, and then into the pulmonary artery now arteries, remember, flow away from the heart, even though in this case
it contains deoxygenated blood, and pulmonary means "of the lungs," so you know this is the path to the lungs. After the
blood makes its way to the alveoli and picks up some fresh oxygen, it flows to the pulmonary vein, remember it's a vein
because it's flowing to the heart, even though it contains oxygenated blood and from there it enters the heart again,
where it flows into the left atrium and then into the left ventricle, where it does the whole body circuit again.
And again and again and again. And that is the way that we work!

Our hearts are really efficient and awesome, and they have to be, because we're endotherms, or warm-blooded, meaning
that we maintain a steady internal temperature. Having an endothermic metabolism is really great because you're less
vulnerable to fluctuations in external temperature than ectotherms, or cold-blooded animals Also, the enzymes that do all
the work in our bodies operate over a very narrow range of temperatures.
In humans that range is between 36 and 37 degrees Celsius. But the trade-off is that endotherms need to eat constantly
to maintain our high metabolisms and also create heat. And for that we need a lot of oxygen. Hence, the amazing,
efficient 4-chambered heart and our gigantic freakin' lungs. Ectotherms, on the other hand, have slow metabolisms and
don't need as much in the way of food. A snake is totally pumped if it gets a meal once a month. So, since ectotherms
aren't doing much in the way of metabolizing, they don't need much in the way of oxygen. So their circulatory systems
can be, you know, a little bit janky and inefficient: it's still cool.

Remember back when we were tracking the development of chordates? One of the signs of complexity was the number
of chambers in an animal's heart. Fish only have two chambers, one ventricle and one atrium. The blood gets oxygenated
as it moves through the gills, and then carries oxygen through the rest of the body, back to the heart where it's moved
through the gills again. But reptiles and amphibians have three-chambered hearts: they've got two atria but only one
ventricle. And what that means is that not all the blood gets oxygenated every time it makes a full pass around the body.
So oxygenated blood gets pumped through the body and mixed up with a little deoxygenated blood. Not super efficient,
but again, it doesn't really have to be. So there you have it. The how and why behind how oxygen gets to all the places it
needs to be!
The question is: What powers the diaphragm? What powers the heart? Where does that energy come from? Well, it
comes from the digestive system.

The Digestive System

The digestive system is so fundamental that it's basically step number one in the guide: How to Make an Animal. You
probably remember that during the embryonic development of most animals, the digestive tract is the very first thing that
forms. When the blastula, that little wad of cells that we all used to be, turns into a little wad of cells with a tube running
through it, that tube is your digestive system. And pretty much every animal has a digestive system of some kind, but
they're not all alike. Far from it. In fact, digestive tracts are specially adapted to animals' feeding behavior and diet. For
instance, a house fly eats mostly liquid or very finely granulated food, but before it does that, it's got to puke its digestive
juices all over its lunch and then let them digest it for a while before it sucks it up into its mouth. If we did it like that, first
dates would be...less common. Most vertebrates put food in one end of the tube and our digestive system processes it,
and then it gets rid of the waste out the other end of the tube. No muss, no fuss. Well, actually, there's a little bit of muss,
at the end.

You may have noticed. But the beauty of it is that this whole process is run by our autonomic nervous system, so we don't
have to think about it, until maybe the very last step when we're in traffic and just had two cups of coffee and a bran
muffin... then we have to think about it a little bit. Among vertebrates, the digestive tract might be short or long, or have
organs that do different things depending on what its feeding habits are.

For instance, dogs are mostly carnivores and also scavengers: They mostly eat meat, but sometimes that meat's been
dead for a while. So, the dog's digestive system has developed to take food in, absorb as many nutrients as possible, and
then deposit it on somebody's lawn, all in a period of about six hours. Dogs have an extremely short digestive tract,
because, if you're in the habit of eating rotten meat, you'd better be able to digest it fast. If you don't, the bad bacteria
that's probably living on that armadillo carcass is going to take up residence in your gut and put you in a world of hurt.
Cows, on the other hand, take a very very very long time to digest their food, around 80 hours, because they have to
process plants, mostly grass. Grass has a ton of cellulose in it, and evolution has yet to produce an animal that can
manufacture a stomach acid or enzyme tough enough to break down cellulose. So, cows have microorganisms in their
guts that break down the cellulose for them. This process takes a four-chambered stomach each one with a slightly
different microecology and a lot of cud-chewing, or regurgitating and re-chewing of grass before it passes all the way
through.

So, nature is full of crazy digestion stories, and I honestly wish that I had time to tell them all. But let's focus on human
digestion from now on, mostly because: You're probably a human, we don't assume anything here, and you'll be wanting
to know how YOUR body does all this stuff. And humans actually have a have a pretty good all-purpose digestive system:
We're omnivores, after all, we eat plants AND meat so our systems are generalized to handle all kinds of stuff. Like most
animals, humans have a bunch of different acids and enzymes in our digestive tracts that break food down so that it can
be absorbed and used by our bodies. But the secret to successful digestion is maximizing surface area. In more that one
way, actually. The first way we maximize surface area is on the food itself. Say I take a bite out of this apple.
Right now there's like, an apple boulder sitting there in my mouth. I've got enzymes in my saliva that immediately start
breaking it down, like, the outsides of the boulder. If I swallowed this chunk whole right now, not only would it hurt like
heck, the rest of my digestive system would have a really hard time dealing with it, because most of the enzymes and
acids would have the same difficulty working all the way through this big solid hunk. But, when I use my awesome teeth to
chew up this hunk of apple suddenly there's double, triple, quadruple the surface area on the food! I'm making up apple
gravel from the apple boulder. Maybe even apple sand.
For humans, chewing is key because breaking down our food into smaller and smaller bits allows enzymes and acids to get
at them. And after our teeth have made the pieces small enough, the chemicals break them down further until
they're fine enough for our bodies to absorb nutrients from them. But it's not just the surface area of the food that's
important, the surface area of the digestive system is key to the whole process as well.

Last time I talked about how we have a whole bunch of surface area in our lungs to absorb tons of oxygen all at once.
Well, our digestive systems work in much the same way. Most of the absorption of nutrients happens in our small
intestines, and the length of the average human adult's small intestine is about 7 meters! Plus, inside our small intestines
there are a bunch of little folds and little absorbing fibers with absorbing fibers on them, and no I didn't mis-speak, the
fibers have fibers.. that's how hard our intestines work to increase their surface area. Last episode I was all impressed that
lungs had a total surface area of 75 square meters... well the small intestine has a surface area of 250 square meters!
Blegh... It's kind of gross. I wouldn't want to see it spread out over a tennis court or anything But I'm getting ahead of
myself here.
Digestion does not start at the small intestine, people, it starts at the mouth. Now, as you can see, this hot pocket is
surrounded by some kind of bread, if you can call it that. Bread is a starch, which breaks down into glucose. When I start
gnawing on a piece of bread, because the outside here is mostly bread. the glands in my mouth start secreting saliva,
which contains salivary amylase, an enzyme designed to break down starch into glucose. The more I chew, the more
amylase will get to all the different sides of the bread, and that's why the more you chew bread, the sweeter it tastes.
Amylase doesn't really do much to the meat or the cheese in this thing. I've got other enzymes and acids that are going to
work on them later on in the system, but I am gonna chew all that stuff up real good right now so that those other
enzymes can do their jobs later. I'm gonna swallow all this.
So now the masticated hot pocket has passed down my pharynx, or throat, and into my esophagus, which leads to my
stomach. There's actually this little cool flap of tissue called the epiglottis that blocks the trachea when I swallow,
so that the food doesn't end up in my respiratory system. This ball of food that I just swallowed actually has a scientific
name, it's called a bolus and it rides a kind of wave of muscle action down the esophagus into the stomach.
This wave-like contraction of the smooth muscles around the tube of the esophagus is called peristalsis, and it's basically
how most of the movement in your digestive system is accomplished. Now my hot pocket bolus is in my stomach now
which is where the food really starts getting manhandled. The stomach basically takes a scorched earth approach to
digestion. It's not messing around. It's like a churning cement mixer that can contract and expand with these big,
accordion-like folds of muscle called rugae. Your stomach's job is to turn everything over and over, smooshing and mixing
all the pieces up with its cocktail of acids and enzymes called gastric juice.

Gastric juice is mainly made up of hydrochloric acid, an enzyme called pepsin, and some mucus and water. Hydrochloric
acid has a Ph of about 1 which is strong enough that, if you got it on your hand, it would give you a chemical burn. So the
acid breaks things down and hopefully kills most of the bacteria that you might find on your food. The pepsin starts
breaking down proteins into amino acids. Now, that mucus is important. It's there to protect your stomach, so that it
doesn't digest itself. When you don't have enough of that mucus you get peptic ulcers, which happen when your stomach
lining comes in direct contact with your stomach acid. And the water's just in there to make everything all soupy, because
what you want by the time your food leaves your stomach is chyme, which is a kind of liquidy slop that you might be
familiar with from the last time you had a stomach virus. You knew this conversation was going to have to get a little bit
gross and I didn't want to bring diarrhea into it too much because, you know, I've been eating. But when something bad is
going on in your digestive tract, your body doesn't worry too much about absorbing nutrients, it just wants to get the
chyme out of there. So, chyme is what you see when... You get the picture. Anyway, there's a little valve, or sphincter
between the stomach and the small intestine that regulates how much chyme gets into the small intestine
and when it gets in there.
The very beginning of the small intestine is called the duodenum, this is where a lot of the small intestine action happens,
by which I mean, lots of things get absorbed and also secreted, like bicarbonate, which neutralizes the gastric acid
before it goes any further. Now, the coolness of the small intestine can't be overstated. It's ground zero for cellular
exchange of nutrients and the breakdown of fats. And again, the reason it's so good at absorbing is because of all the
surface area it's got going on. A lot of that surface area comes from the fact that, despite its name, your small intestine is
frickin' long: in a human, it can range anywhere from 4.5 to 10.5 meters. But that's not all! The whole inside is lined with
epithelial tissue that has tons of ridges and folds in it. Surface area to the max! And on those ridges and folds are these
little hair-like fibers of flesh called villi. Each villus has capillaries in it, so that it can absorb nutrients. And get this: each
villus, which is only like half a millimeter long is covered in teeeny tiny little microvilli, providing even more surface area!
In fact, apparently, the small intestine has a texture kind of like velvet, which is... [DISGUST]
Oh great, now I eat the milkshake? Fantastic. Okay. So another thing the small intestine does, with the help of its friend
the gallbladder, is break down fatty stuff, like this milkshake. Near the top of your small intestine is a little pipe
where bile salts, manufactured by the liver and stored by the gallbladder, are squirted out into the small intestine.
Bile works like dish detergent on a pan you just fried something in: it's an emulsifier. It takes hydrophobic fat molecules
and breaks them up into fatty acids and monoglycerides, which can be absorbed by all that epithelial tissue! I've never
had Chunky Monkey before Mmmmm! Nuts!
After your food passes through those yards and yards of small intestine, the chyme goes through another sphincter and
enters the cecum, the beginning of the large intestine. The large intestine's job is to remove most of the water
and bile salts from the chyme so you don't have constant diarrhea. So, you can thank it for that! It's called "large" because
it's wider than the small intestine, but it's not nearly as long: it's basically just a one- and-a-half meter victory lap around
outside of the small intestine and then it calls it good Also, I should mention, at the end of the cecum there is a little tube
where the the appendix comes in.

For a long time, we thought that the appendix was a worthless, vestigial structure that we used to need at some point
in our evolution but didn't need anymore. However, recent studies are finding that the purpose of the appendix in
modern humans is probably to act as a safe house for all of the good bacteria you need to help you digest your food. If
you get a virus or food poisoning or something and all your digestive systems say, "GET IT ALL OUT OF ME!"

The appendix has a little sample of your gut bacteria that it spits out to help recolonize you after your illness. So, I think
you're familiar with the final step in the digestive system. That's the pooping. Your food can spend as long as 3 days in
your digestive tract, and a lot of that time is spent in the large intestine, mostly reabsorbing the excess water from the
chime and prepping your poo for its great entrance into the world. When it's done, it passes through everybody's favorite
sphincters, the anal sphincters. There are two of them. And, you know...out in the world to live its own life.

The Excretory System

One of the coolest and most important things that our bodies do is maintain this thing called homeostasis, the regulation
of a stable internal environment, no matter where we are or what we're doing. After all, we put our bodies through a lot
every single day: We're always adding food and liquid and chemicals, and we're constantly changing temperature and our
levels of activity, but our bodies can roll with it. It's like, no big deal for them. All of our organ systems have some hand in
maintaining homeostasis. I mean, it's basically the thing that makes us not dead. But the excretory system, aka the urinary
system, which includes the kidneys, the ureters, the bladder, and the urethra, is the star quarterback of the homeostasis
team That's because your excretory system is responsible for maintaining the right levels of water
and dissolved substances in your body. This is called osmoregulation, and it's how our bodies get rid of the stuff we don't
need, like the byproducts of metabolizing food, while also making sure we don't get dehydrated.

It's the body's greatest balancing act, and your body is doing it right now, and all the time, as long as you're not dead. As
with other organ systems we've talked about, not all excretory systems in the animal kingdom are created equal. Different
animals excrete waste different ways based on their evolutionary history what environments they live in, and what their
hobbies and interests are. These factors all influence how an animal regulates water, and most metabolic waste needs to
be dissolved in water in order to be excreted.

The problem is, a main byproduct of metabolizing food is ammonia, which comes from breaking down proteins, and it's
pretty toxic. So, depending on how much water is available to an animal and how easy it is for the animal to lug a bunch of
water around inside it, animals convert this ammonia into either urea or uric acid. Mammals like us, as well as
amphibians, and some marine animals like sharks and sea turtles, convert ammonia into urea, a compound made from
combining ammonia and carbon dioxide, in their livers. The advantage of urea is its very low toxicity. It can hang out in
your circulatory systems for a while with no ill effects. But you have to have some extra water available to dissolve it and
get rid of it.

This isn't such a tall order, really, I mean peeing isn't a huge inconvenience, I mean, is it? It's not for me anyways. Well, it
would be, though, if you a bird or an insect or a lizard livings in the desert. Animals that have to be light enough to fly or
don't have a bunch of spare water hanging around, convert ammonia into uric acid, which can be excreted as a kind of
paste, so not a lot of water is needed. You've seen bird poop. If you haven't taken a close look, next time, do that.
Just look.
The white stuff in the bird droppings is actually the uric acid-y pee and the brown stuff is the poop. So, now that we've
established what is and what is not bird poop, let's get down to the brass tacks of how humans get all of this urea out of
our blood and into our toilets.

The excretory system starts with the kidneys, the organs that do all the heavy lifting, from maintaining those levels of
water and dissolved materials in our bodies to controlling our blood pressure. And even though they do an amazing job,
I'm not bad-mouthing your kidneys here, the way that they do it is frankly a little bit janky and inefficient. They start out
by filtering out a bunch of fluid and the stuff dissolved in the fluid out of your blood, and then they basically re-absorb
99% of it back before sending that 1% on its way in the form of urine. Seriously, 99% gets re-absorbed.

On an average day, your kidneys filter out about 180 liters of fluid from your blood, but only 1.5 liters of that ends up
getting peed out. So most of your excretory system isn't dedicated to excreting it's dedicated to re-absorbing. But the
system works, obviously, I'm still alive. So we can't argue with that. Now it is time to get into the nitty gritty details
of how your kidneys do all this, and it's pretty cool. But there's lots of weird words. So get ready. Your kidneys do all this
work using a network of tiny filtering structures called nephrons. Each one of your mango-sized kidneys has about a
million of them If you were, don't do this, but if you were to unravel all of your nephrons and put them end to end, they
would stretch over 80 kilometers. This is where all the crazy action happens, so to understand how they work, we're just
going to follow the flow, from your heart to the toilet.

Blood from the heart enters the kidneys through renal arteries, and just so you know, whenever you hear the word
"renal"
it means you're dealing with kidney stuff. As the blood enters, it's forced into a system of tiny capillaries until it enters a
tangle of porous capillaries called the glomerulus. This is the starting point for a single nephron. The pressure in the
glomerulus is high enough that it squeezes some of the fluid out of the blood, about 20% of it, and into a cup-like sac
called the Bowman's capsule. The stuff that's squeezed out is no longer blood, it is now called filtrate. It's made up of
water, urea, some smaller ions and molecules like sodium, glucose and amino acids. The bigger stuff in your blood, like
the red blood cells and the larger proteins, they don't get filtered. Now the filtrate is ready to be processed.

From the Bowman's capsule, it flows into a twisted tube called the proximal convoluted tubule, which means "the tube
near the beginning and that is all wind-y." WHY ARE WE SO BAD AT NAMING THINGS?! Anyways, this is the first of two
convoluted tubules in the nephron. And these, along with other tubules we're talking about, are where the
osmoregulation takes place. With all kinds of tricked out, specialized pumps and other kinds of active and passive
transport, they re-absorb water and dissolved materials to create whatever balance your body needs at the time.

In the proximal tubule, it's mainly organic solutes in the filtrate that are reabsorbed like glucose, and amino acids, and
other important stuff that you want to hang on to. But it also helps to re-capture some sodium, potassium and water
we're going to want later. From here, the filtrate enters the Loop of Henle, which is a long, hairpin-shaped tubule that
passes through the two main layers of the kidney. The outermost layer is the renal cortex, that's where the glomerulus,
bowman's capsule, and both convoluted tubules are, and the layer beneath that is the renal medulla, which is the center
of the kidney.

"Cortex," by the way, is Latin for tree bark, so whenever you see it in biology, you know that it's the outside of something.
"Medulla," on the other hand, meaning narrow or pith, so you know that it's the inside. Just to help you remember this
stuff. But, before we take a tour of this amazing loop I have to do a couple of things.

First, go pee. Because this is...you know. And second, a Biolo-graphy! So I'll be right back! The Loop of Henle was
discovered by 19th century German physician and anatomist Friedrich Gustav Jakob Henle. I'm pretty sure he was one of
those guys that you can't gross out since he spent most of his career dissecting kidneys, eyeballs, and brains. And also
seemed to be a huge fan of mucus and pus. He was by far the most important anatomist of his time. His three-volume
Handbook of Systematic Human Anatomy was recognized as the definitive anatomy textbook of its day and was famous
for its exquisite attention to detail and its intricate, even beautiful, illustrations. Not only did Henle discover the Loop of
Henle, arguably the linchpin of kidney function in mammals, he was an early adopter of the wildly unpopular
germ theory of disease. His student Robert Koch is considered one of the founders of microbiology, and the two worked
together to formulate the Henle-Koch Postulates, which today remain the four conditions that must be met to establish a
causal relationship between a microbe and a disease. Henle taught the world so much about the human body that there
are, right now, in you, no fewer than 9 features that bear his name. From the Henle's fissures between the muscle fibers
of your heart to the Crypts of Henle, which are microscopic pockets in the whites of your eyes.

Alright, so, review time. We've squeezed some filtrate out of the blood, and re-absorbed some of the important organic
molecules we want to keep. But most of the re-absorption action happens here, in the Loop of Henle, which does three
really important things. One, it extracts most of the water that we need from the filtrate as it travels down to the medulla.
Two, it pumps out the salts that we want to keep on the way back up to the cortex. And three, in the process of doing all
that, it makes the medulla hypertonic, or super salty relative to the filtrate. Creating a concentration gradient that will
allow the medulla to draw out even more water one last time from the filtrate, before the final journey to the toilet
begins. It's complicated and, again, kinda janky, but it's what allows us mammals to create urine that's as concentrated as
necessary, using only the amount of water that our bodies can spare at the time.

So first, filtrate starts going down the loop, and the thing to know here is that the membrane is highly permeable to
water,
not so much to salt or anything else, mainly water. Now, compared to the filtrate, the tissue of the medulla is already
pretty salty. And as the filtrate processes, the surrounding tissue becomes increasingly hypertonic the farther down you
go, the saltier it gets. So, applying everything we've learned about osmosis, you know that as the filtrate moves along, it
loses more and more water through the membrane. By the time the filtrate gets to the bottom of the Loop, it's highly
concentrated. Now the filtrate enters the ascending end of the Loop, and here it's basically the same but in reverse. The
membrane is NOT permeable to water, and instead it's lined with channels that transport ions like sodium, potassium and
chlorine. And because the filtrate is so concentrated now, it's actually hypertonic compared to the fluid outside in the
medulla. So as it ascends, huge amounts of salts start flowing out of the filtrate, which makes the renal medulla really,
really, really salty. This salty medulla also creates a concentration gradient between the medulla and the filtrate which
we're going to need in the final step of pee-making. But first! Once the filtrate is back up in the cortex and out of the loop,
it enters the second of our convoluted tubules, called the distal convoluted tubule, or "farther-away curly tube."

While the first tubule worked mostly on reabsorbing the organic compounds in the filtrate, here the focus is on regulating
levels of potassium, sodium, and calcium. This work is mainly done by pumps and hormones that regulate the
reabsorption process. By the time it's done, we've finally taken everything we want to keep out of the filtrate, so now it's
mainly just excess water, urea and other metabolic waste. This stuff all gets dumped into collecting ducts that channel it
back down to the center of the kidney, the medulla. And remember, the medulla is super-salty, right?

Now more hormones kick in that tell the collecting ducts how porous to make their membranes. If the membranes are
made very porous, more water is absorbed into the medulla, which makes the urine yes, we can start calling it urine now
even more concentrated. And here's a fun fact: If you've ever had one drink too many, you might've noticed that you start
to pee a lot, and your pee is clear. That's because alcohol interferes with these hormones especially one called anti-
diuretic hormone which tells the collecting ducts to be very porous so that you reabsorb most of the water. With those
hormones all confused and out of commission, you just starting peeing out all kinds of water, which also means you're
getting dehydrated, which means you're officially on a one-way trip to Hangover City. So, now you know why that
happens.
Now at this point, the urine leaves both kidneys and flows down to the urinary bladder by tubes called ureters. Once in
the bladder, the urine just sits around, waiting for us to decide when it's time to find a bathroom. And when that time
comes, a little sphincter muscle relaxes and releases the urine from the bladder into a tube called the urethra, which
empties out wherever you point it. So that's how your excretory system works! And that's basically how it works for most
mammals, although some modifications are made based on, again, where they live and what they do.

For instance, kangaroo rats, which are tiny and adorable and live in the desert, have the most concentrated urine
of any animal anywhere, because it can't spare the water. So it has a very, very long Loop of Henle that reabsorbs most of
the water from the filtrate.
On the other end of the spectrum, we have the beavers, who have very short Loops of Henle, because they're like, "Water
reabsorption, schmater reabschmorption. Do you see what I do all day?" And so now you know the true origins of pee.

The Reproductive System

The number one question on the mind of every organism on earth, if that organism happens to have a mind, is: how do I
make more of myself? It's bigger than all the other questions combined, including how am I going to feed myself? And
what's the meaning of life? Because, from a biological perspective, we know what the meaning of life is. Biology has
answered that question. It's reproduction.

Different organisms go about reproducing in different ways: You can make more of yourself by yourself, a strategy called
asexual reproduction, OR you can team up with somebody else and make a baby that's genetically different than both of
you through sexual reproduction. From liver flukes to pine trees, 99% of the eukaryotic organisms on earth use sex to
reproduce, at least some of the time, because by creating offspring with a slightly different genome, helps the new
generation stay one step ahead of pathogens or competitors. Or if you're the pathogen, it helps you stay ahead of that
pesky host that's always trying to kick you out. But still, sex is inconvenient and it's a lot of work:

First you have to find somebody to mate with, which means you have to get out of bed and brush your teeth and stuff.
Then, if you're an animal, you have to find somebody who's willing to mate with you, and then figure out whether he or
she
is going to provide higher or lower quality genes than yours. Thankfully, and unsurprisingly, animals' reproductive systems
have evolved to streamline all of those inconveniences to address one, and only one, aim: to get your sex cells where they
need to be. So, sex. How does it work? I thought you'd never ask.

Reproductive systems, like all the other systems we've discussed, take on an incredible diversity within the kingdom
Animalia. For instance, some female spiders mate with a bunch of different males and stash their sperm in different
storage units. When she's ready to fertilize her eggs, the female spider will choose which male spider she liked the best
and let his sperm out the storage unit to fertilize her eggs! Hyenas, meanwhile, have female-dominated social system, and
it's the alpha female who chooses who she mates with. And she has sex using an enlarged sensitive sex organ, a clitoris,
that looks exactly like a penis, called a pseudopenis. And a duck's penis can be a quarter of the length of its body, and
shaped like a corkscrew. Want to know why? Look it up! Actually don't! Google that with care! Just don't press play on the
video.
The point here is that while the delivery systems may be somewhat different from animal to animal, the fundamentals are
the same. In order to do the sex, an organism needs to find another of its species that has a different type of gamete, or
sex cell, than their own. Gametes, you'll recall, are haploid cells, meaning they have only one set of chromosomes, and
they're formed by the process of meiosis. And there are only two kinds of gametes: One is the ovum, or egg.
In plants it's called an ovule.
The egg is always a large cell that takes a lot of time and energy investment to make, and it's usually not very mobile. The
other type of gamete, sperm, are smaller, a lot more plentiful, easy to make, and always more mobile than eggs. Most
animals have either one or the other type of gamete, though hermaphroditic species, like garden snails and some
flowering plants, can produce both. In the magical moment that one of these sperm finds one of those eggs, the two fuse
together to create a single diploid cell that has all of the instructions to make a new seahorse or secretary bird or
whatever it is. But let me get your mind right about what we really mean when we talk about sex. Because we humans
have
external sex organs, called genitals, we tend to think of them as key indicators of who's male and who's female. But the
fact is, genitals are only one byproduct of a much, much more important and fundamental distinction:

From a biological perspective, the only thing that makes sexes different is that females produce big, not very mobile
gametes, and the males make smaller, much more mobile gametes. Across the spectrum of all things that reproduce
sexually, that's pretty much the only consistent difference between boys and girls. Therefore, all reproductive systems
and reproductive behavior are designed entirely around the production, storage, and delivery of these gametes.
For instance, because sperm are really mobile, males within a species are generally the more mobile ones who go out to
find a mate. This is even true for plants: female gametes of a flowering plant generally stay in one place while the pollen,
which ends up producing the sperm, gets picked up by a pollinator, or sometimes just sprays out every which way in the
wind, hoping to bump into the right kind of ovule.

In animals, we see all kinds of crazy behaviors where mating is concerned. And of course not every animal goes about
courtship in the same way, but one thing is pretty consistent: Females tend to be pickier about the quality of their mates,
because while a male animal could conceivably fertilize thousands of eggs every year, a female has only a limited number
of eggs, and she's spent a lot of energy developing them, so she wants them to be fertilized with high quality genes. Plus,
in cases where both parents stay together after fertilization, she also wants those genes to be attached to a high-quality
provider.
This often results in males having to do a lot of showing off in order to get a lady's attention. Males of a species are
generally louder, larger, brighter, more combative than the females. Basically, they're putting on a big show so the
females can size up how awesome that guy's genes are. But for all those differences, during the development of the
embryo, there are actually very few physical differences between males and females, at least at first. You and I, we didn't
start out being a male or a female. While you were hanging out in our mom's uteri, you didn't have a sex at all until about
two months. Before that, we had all the pieces to become either male or female, but our genes hadn't gotten together to
determine whether or not our gonads, the glands that make the gametes, were going to become ovaries or testes.

In mammals, that decision is made by the sex-determining chromosome: If an offspring has two of the same kind of sex-
determining chromosome, called XX, it will be female. And if it has two different chromosomes, XY, it will be male.
The same is true for some other animals like fruit flies, and even some plants like Gingko trees. However, the opposite is
true for birds: boy birds have XX and girl birds have XY. Go figure.

In mammals, the default setting for sex is always female. Absent a signal from the Y chromosome, ovaries form and begin
working on developing female structures. If there is a Y, the ovaries instead form into testes, and parts that would be
female turn into male structures, for instance, the clitoris I mentioned, which is sensitive and has spongy tissue in it,
actually becomes part of a penis. But it's worth pointing out that by this time, some features are already in place before
the sex is determined.

Nipples, for instance, form before this point, so that's why men have them, even though they don't do anything. Now,
once the sex is determined, the ovaries and testes pump out estrogen and testosterone. Meanwhile, the brain is growing
and creating receptors, organized differently in males and females, that will later determine how both estrogen and
testosterone are used in the body. Soon after a baby girl is born, she'll have half-formed versions of all the eggs she's ever
going to have for her whole life, then at puberty, once a month, one of those eggs will finish forming and be released.
But for baby boys, the sperm-making does not begin until around puberty.

Most of the time when a young animal starts getting close to sexual maturity, secondary sex characteristics crop up: In
humans more body hair appears, boys all of a sudden develop facial hair, while both sexes get more pubic hair.
Also muscle and fat get redistributed around the body, the most obvious example being breasts. In other animals,
secondary sex characteristics include things like manes on male lions, a big old funky rack of feathers on male peacocks,
antlers on male deer. Males really have the market cornered on fancy, showy secondary sex characteristics. So, by the
time an animal has reached sexual maturity, the males and females of a species often look pretty dissimilar. Not just of
each other, but of their previous non-sexually-developed forms.
Basically showing the world that their different reproductive structures that they were born with are now in full gear, and
they've got some really different jobs to do, based on what sex they are. So let's go over how this all works with human
people. And of course, ladies first. As you know, the gonads of a female embryo turn into two ovaries, one on either side
of the uterus, with its oviducts, or fallopian tubes, reaching out toward them. The ovaries are where those precious eggs
are kept. Maybe the biggest difference between women's and men's reproductive set-up is that women have a menstrual
cycle, typically a four-week process in which one egg matures in an ovary and is released to be drawn into the fallopian
tubes, a process called ovulation. If, while the egg makes its way down the fallopian tube to the uterus, a sperm finds it
and fertilizes it, there's a chance that the fertilized egg will implant on the endometrium, a tissue layer inside the wall of
the uterus and a baby will grow. However, it's estimated that up to 70% of fertilized eggs don't take hold in the
endometrium.
This could be because women's bodies have sort of a built-in genetic testing: If something's suspected to be wrong with
the growing embryo, the lining of her uterus that she's built up over the past month will shed, and the woman will
menstruate as usual. This material leaves the female reproductive system through the narrow lower end of the uterus,
the cervix, and then out into the muscle-lined tract of the vagina, and those are of course the same structures through
which
a newborn baby passes and through which the sperm enter. While a woman's body is busy all month developing the next
egg, getting it ready for fertilization, and shedding her uterine lining if it's not fertilized, males are undergoing a
completely different process that calls on a lot of other highly specialized reproductive structures. We start of course,
with the testes, which are made up largely of a bunch of coiled tubes called seminiferous tubules, which are where the
sperm form.

Unlike a woman's ovaries, the testes are outside of the body, because in order to make sperm, they have to be kept
at a specific temperature, usually about 2 degrees Celsius cooler than inside of the body cavity. For that reason, the testes
are kept in a pouch called a scrotum that's in charge of keeping the testes at the perfect sperm-making temperature.
After being produced in the testes, human sperm spend about 3 weeks coiled in tubes in the scrotum called the
epididymis, and that's where they mature and grow flagella, the little whip-like tails sperm are so famous for and which
make them able to move around and swim. Now the sperm stay here until they're ready to leave the body, so before we,
or they, can go any further, we have to set the stage for that. As you know, in humans and some other animals, the penis
usually sits around, not doing much except for letting urine out of the bladder from time to time. But every so often a
male realizes that he's totally going to get the chance to mate! At this point, spongy tissue in the penis fills with blood, and
BAM, erection!

Some animals like raccoons, whales and walruses actually have a literal bone in their penis to help the erection along.
But either way, the point is to allow the penis to enter the vagina, which scientists call coitus, and deposit the sperm he's
put so much into making. These sperm travel in a special fluid, semen, whose ingredients aren't combined until they're
ready to be released by a series of muscular contractions that cause emission, more commonly known as ejaculation. At
this point, the contractions carry the mature sperm from the epididymis, through two muscular ducts called the vas
deferens, which carry them up from the testes, up and over the bladder, and down past the seminal vesicles.

Here, with contributions from the nearby prostate gland, they pick up a bunch of fluid that contains mucus, coagulating
enzyme, ascorbic acid and sugars that the sperm are going to need for their trip. Now the semen is complete, and it
travels down the short ejaculatory ducts to the urethra to be released at the end of the penis, where if the timing is right,
one among the hundreds of millions of sperm in that emission can find and fertilize an egg. That, my friends, is how we all
get our start.

Central Nervous System

James was healthy professional, a father of two. He had lots of friends, loved telling jokes, and played softball on Sundays.
Then one day, at the age of 45, he suffered a stroke. He bounced back fairly quickly, with one major exception: He was no
longer able to speak. The stroke damaged a specific area in the left hemisphere of his brain called Broca’s area, and left
him with what’s known as Broca’s aphasia.

Broca’s area is partly responsible for the ability to produce and process language, and Broca’s aphasia often leaves its
sufferers with some ability to understand speech, but an inability to produce intelligible words. James could understand
his wife when she asked if he wanted cereal for breakfast, but he could only respond by repeating the word “too” --
although he could still intonate as though he were speaking a whole sentence. Then, after some time and therapy,
something rather unexpected happened -- James regained some ability to communicate through singing.

Broca’s aphasia can sometimes be treated by teaching patients to sing, because singing uses a different region of the
brain -- one that’s on the right side and that’s analogous to Broca’s area on the left. So after some practice, James could
sing words, and he eventually relearned how to talk by teaching the right side of his brain to speak rather than sing.
Whether it’s a stroke affecting your speech, a tumor destroying your memory, a concussion affecting your aggression, or
that fateful iron rod that shot straight through Phineas Gage’s skull -- a lot of what we know about how the brain works
has come through studying injuries to it. And what we’ve learned so far is that, even though it looks like a 1.4-kilogram
lump of
gray, congealed oatmeal, the brain is made up of super-specific areas that have super-specific functions. You might
actually say the same thing about your brain that’s sometimes said about politics: Everything is local.

You’ll remember that our nervous system is divided into two main networks that work in harmony -- the central nervous
system, consisting of your amazing brain and spinal cord, and the peripheral nervous system, made up of the nerves
coming out of that central nervous system. The central nervous system’s main game is integrating the sensory
information that the peripheral system collects from all over the body, and responding to it by coordinating both
conscious and unconscious activity. The sun is bright, so I’ll shade my eyes; I’m hungry, so I’m calling the pizza man; the
phone is ringing, maybe I’ll answer it. All these sensations, thoughts, and directions process through this two-part system.
It’s the brain, of course, that sorts out all that sensory information and gives orders. It also carries out your most complex
functions, like thinking, and feeling, and remembering.

Meanwhile, your spinal cord conducts two-way signals between your brain and the rest of your body, while also governing
basic muscle reflexes and patterns that don’t need your brain’s blessing to work -- this is how a chicken can still run
around even if the poor thing has been decapitated. Both your spinal cord and brain are made of fragile, jelly-like nervous
tissue that is extremely susceptible to injury. So all that goo is well-protected by the bones of your vertebrae and cranium,
as well as membrane layers, or meninges, before being bathed in a cushy waterbed of clear cerebrospinal fluid.

This fluid actually allows your brain to float somewhat in your skull, reducing its weight and letting it slosh around while
you and your head are free to move. But even with all that extra protection, your brain is still vulnerable. And one thing
James’s story taught us is that its vulnerabilities can be incredibly specific, because your brain is divided into specialized
regions that may, or may not, interact with each other to produce a given action. We can better understand this division
of labor by looking at how the brain first develops into its main component parts. Inside a developing embryo, the central
nervous system starts off as a humble little neural tube. Soon the caudal, or lower, end of the tube stretches out, forming
the spinal cord, while the cranial end begins to expand, divide, and enlarge into three primary brain vesicles,
or interconnected chambers. This is kind of your proto-brain. We call these chambers the prosencephalon, the
mesencephalon, and the rhombencephalon -- or forebrain, midbrain, and hindbrain. By an embryo’s fifth week of
development, these main three chambers start morphing into five secondary vesicles that essentially form the roots of
what will become your grown-up brain structures.

The prosencephalon divides into two sections -- the telencephalon and the diencephalon. The rhombencephalon forms
into another pair, called the metencephalon and the myelencephalon. And in between, the mesencephalon, thanks to
evolution, remains undivided. The real action starts as these five secondary vesicles start developing into the major adult
brain regions that you might be more familiar with -- the brainstem, the cerebellum, the diencephalon, also known as the
interbrain, and finally the cerebral hemispheres. But, in order to go from a simple tube into that classic, wrinkly icon we
think of as the “brain,” each of these five vesicles grows in different ways. Basically, some develop a lot more than others.
The least dramatic changes occur in the three most caudal or lower sections: the mesencephalon, the metencephalon,
and the myelencephalon. They go on to form the cerebellum, which mostly helps coordinate muscular activity, and the
brainstem, which plays a vital role in relaying information between the body and the higher regions of the brain.
The brainstem actually has three main components -- and I know this is getting to be a lot of vocabulary here -- you have
the midbrain, the pons, and the medulla oblongata. Together they regulate many of the most basic, vital involuntary
functions, like keeping your heart on pace, keeping your lungs working, and controlling things like sleep, and appetite, and
pain sensitivity, and awareness. But of the three brainstem parts, it’s your midbrain that carries out the higher-level
functions. Like, when your eyes track a fast moving object, or when you look behind you after hearing some sudden loud
sound, it’s the midbrain that receives and processes that sensory information and sends out the reflexive motor signals,
so you react without thinking.

The midbrain also passes that data to regions like the cerebral cortex, which do the actual conscious thinking about the
stimuli, like “What is that thing whizzing across the sky?” or “WHAT JUST EXPLODED BEHIND ME?!” So with the brainstem
and cerebellum covering your basic life and motor functions, you start to see somewhat more complex tasks being carried
out in the next major brain structure, the diencephalon. This is where you find the thalamus, hypothalamus, epithalamus,
and the mammillary bodies, which regulate things like homeostasis, alertness, and reproductive activity. Here we also find
part of the limbic system, which is a center for strong emotions, like fear. This area is sometimes called the “reptilian
brain” because we share it with some of our less philosophical animal brethren like lizards and fish. I’m not putting these
guys down, but by our standards, they don’t think so much as focus on the more instinctual pursuits that are ruled by the
caudal regions of the brain -- eat, drink, sleep, mate, stay safe.
All those things are awesome. But it wasn’t until the appearance of birds and mammals that some animals’ brains came to
be dominated by the last of the five vesicles, the telencephalon. During your brain’s growth, the telencephalon undergoes
the biggest changes of all, as it develops into the most brainy part of your brain -- the two classic, walnut-looking
hemispheres we collectively call the cerebrum, that cover the rest of your brain like a mushroom cap on its stalk. That’s
the cerebrum -- not to be confused with Cerebro, which is Professor X’s telepathy-enhancing device -- and it is the largest
region of the brain and performs the highest functions. It’s made up of the wrinkled, outer layer of “gray matter” called
the cerebral cortex, and the inner squishy layer of “white matter” beneath it. And it’s the cerebrum that rules our
voluntary movements and our most advanced tricks, like thinking, and learning, and regulating and recognizing emotions,
and experiencing consciousness in general.

You’ll remember that higher processing requires lots of synapses, which require lots of nervous tissue. So as the cerebrum
grew through evolutionary time, it got more massive but our skull didn’t exactly keep up. So in order to squeeze all that
material into your skull, the brain forms little creases, called gyri, and larger grooves, or sulci, giving it more folds than
than an origami pineapple. And although a big fissure separates the left and right hemispheres, the two halves
communicate, through a series of myelinated axon fibers called the corpus callosum. And each hemisphere has other,
smaller fissures that divide it into lobes -- each with a different set of major functions.

The frontal lobe, for example, governs muscle control and cognitive functions like planning for the future, concentration,
and preventing socially unacceptable behaviors. In most people, this area doesn’t finish developing until after the teenage
years, which tells you a lot about the teenage years. Since Broca’s area lives in this lobe in the left hemisphere, it also is
important in language comprehension and speech. If you’re enjoying a beautiful sunset, you can thank your occipital lobe
at the back of your head for processing those bright visual cues. And the next time you step on a lego, you can curse your
parietal lobe, which processes the sensations of touch, pain, and pressure. Meanwhile the temporal lobe helps sort out
auditory information, including language. It contains Wernicke’s area -- another important region of the brain associated
with the production of written and spoken language. This part of the limbic system includes your short-term memory
keeper, the hippocampus, and the emotional amygdala, which controls sexual and social behavior. So, if you damage
the wrong part of your temporal lobe, you may never again be able to remember what you ate for lunch… or you might
suddenly become a total jerk who kicks kittens and cuts in line. We could do a whole course on the finer-grained functions
and consequences of malfunction in every bit of brain in your gourd, but, well, we can’t do that today. And you got to
remember that, when it comes to your body, no organ or system is an island. Your brain would be pretty useless if it
weren’t hooked up to the outside world. That’s where the peripheral nervous system comes in, which we’ll be spending
the next few lessons exploring.

The Peripheral Nervous System

When it comes to the nervous system, or just your body in general, let’s face it: your brain gets all the props. And it
deserves those props! It’s a complicated, and crucial, and sometimes crazy boss of an organ. But your brain would be
pretty useless without a support team that kept it connected to the outside world. Because frankly, like any leader, the
more isolated your brain gets, the weirder it gets. Put a person in a watery, pitch-black sensory deprivation tank, and
you’ll see the brain do some really weird stuff. Without a constant flood of external information, the brain starts to
confuse its own thoughts for actual experiences, leading you to hallucinate the taste of cheeseburgers, or the sound of a
choir singing, or the sight of pink stampeding elephants. It’s your peripheral nervous system that keeps things real, by
putting your brain in touch with the physical environment around you, and allowing it to respond. This network snakes
through just about every part of your body, providing the central nervous system with information ranging from the
temperature, to the touch of a hand on your shoulder, to a twisted ankle.

The peripheral nervous system’s sensory nerve receptors spy on the world for the central nervous system, and each type
responds to different kinds of stimuli. Thermoreceptors respond to changes in temperature. photoreceptors react to light,
chemoreceptors pay attention to chemicals, and mechanoreceptors respond to pressure, touch, and vibration. And then
we’ve got specialized nerve receptors called nociceptors that, unlike those other receptors, fire only to indicate pain,
which is the main thing I want to talk about today. Because, as unpleasant as a stick in the eye or tack in the foot may be,
pain is actually a great example of where everything we’ve talked about over the last few weeks all comes together, as we
trace a pain signal through your nervous system, from the first cuss to the Hello Kitty band aid. By the end of this episode
of Crash Course Anatomy & Physiology you’ll never think of a stubbed toe, pounding headache, or burned tongue the
same way again.
Most people go to great lengths to avoid pain, but really, it’s an incredibly useful sensation, because it helps protect us
from ourselves, and from the outside world. If you’re feeling physical pain, it probably means that your body is under
stress, damaged, or in danger, and your nervous system is sending a cease and desist signal to stop twisting your arm like
that, or to back away from that bonfire, or please seek medical attention, like, RIGHT NOW.

So in that way, pain is actually good for you -- that’s why it exists. I’m not saying it’s pleasant, but if you’ve ever wished for
an X-Men-like power to be impervious to pain, I’ve gotta say, that is one foolish monkey’s paw of a wish. Just ask Ashlyn
Blocker. She’s got a genetic mutation that’s given her a total insensitivity to any kind of pain. And as a result, she’s absent-
mindedly dunked her hands in pots of boiling water, run around for days without noticing broken bones, and nearly
chewed off her own tongue. Luckily, such congenital conditions are very rare. The rest of us have a whole nervous
system dedicated to making sure our bodies react with a predictable chain of events at the first sign of damage.

Like say you just wake up and you’re extraordinarily hungry for some reason, so you run downstairs to grab some clam
chowder, but you didn’t put any shoes on and suddenly you’re like, “YOWW!” There’s a tack, fell out of the wall, and you
stepped right on it -- of course. Your foot immediately lifts off the ground, and then you’re assuring your dog that you’re
not yelling at her, you’re just yelling, and then you limp over to the couch, and sit down, and you pull up your foot, and
remove that spiny devil from your flesh. You want to talk physiology? So what exactly just happened in your body? Well,
the first step was a change in your environment -- that is, a stimulus that activated some of your sensory receptors.

In this case, it was a change from the probably completely ignored feeling of bare skin on a smooth floor to a distinct
feeling of discomfort -- the sharp metal tack piercing your skin. Your peripheral nervous system’s mechano- and
nociceptors provided that base sensation, or awareness that something had changed. Then it went to your central
nervous system -- first to the spinal cord that caused the immediate reflexive action of pulling up your foot, and then your
brain eventually interpreted that awareness into the perception of pain, and decided to pull the tack out and probably say
an expletive or two. Pain itself is a pretty subjective feeling, but the fact is, we all have the same pain threshold.
That is, the point where a stimulus is intense enough to trigger action potentials in those nociceptors is the same for
everybody. But, you and I might have different tolerances for discomfort. In general, most doctors think of pain as the
perception of pain -- whatever any given brain says pain is. So, you’ve got the stimulating event -- foot meets tack -- and
then the reception of that signal, as the nociceptors in your foot sense that stimulus, and then the transmission of
that signal through your nerves to your spinal cord and eventually up to the brain.

Now remember back how every neuron in your body has a membrane that keeps positive and negative charges separated
across its boundaries, like a battery sitting around waiting for something to happen? Well that tack in your flesh is that
something. And it snaps those nociceptors to attention. Some neurons have mechanically-gated receptors that respond to
a stretch in their membranes -- in this case, that happens when the tack punches through them. Meanwhile, other
neurons have ligand-gated receptors that open when the damaged skin tissue releases chemicals like histamine or
potassium ions. These channels allow sodium ions to flood into the neuron, causing a graded potential, if that hits the
right threshold, it activates the electrical event that sends the signal all the way up the axon and gets one neuron talking
to another -- the action potential.
When that action potential races down the length of its axon to the terminal, the message hits the synapse that then
flings it over that synaptic gap to another neuron that’s in your spinal cord. Remember, signals travel between neurons
either by electrical or chemical synapses. The electrical ones send an electrical impulse, while the chemical ones -- the
ones I’m talking about now -- first convert that signal from electrical to chemical, by activating neurotransmitters to
bridge the synaptic gap, before the receiving neuron converts that chemical signal back into an electrical one. In this case,
news of the tack-attack is carried by specific neurotransmitters whose sole job is to pass along pain messages.

Now, so far, your body’s response to the stimulus has been handled by the sensory, or afferent, division of your peripheral
nervous system. This is the part that’s involved expressly in collecting data and sending it to the central nervous system.
But at this point, the responsibility changes hands. The torch is passed. Because the pain signal has just triggered an
action potential in a neuron in the spinal cord, which is part of the central nervous system, and there it reaches an
integration center. From here, the response is taken over by the motor, or efferent division. Once the integration center
interprets the signal, it transmits the message to motor neurons, which send an action potential back down your leg,
where it reaches an effector. And an effector is just any structure that receives and reacts to a motor neuron’s signal, like
a muscle contracting or a gland secreting a hormone. From here, the motor neurons complete the whole foot-lifting
response until the rest of your nervous system gets engaged in the complicated tasks of figuring out what the problem is,
and fixing it. Those are the five steps that your highly specific neural pathways go through to produce what’s known as a
reflex arc. A lot of your body’s control systems boil down to reflexes just like this – immediate reactions that can either be
innate or learned, but don’t need much conscious processing in the brain. Lifting your foot when you step on a tack is an
innate, or intrinsic, reflex action -- a super fast motor response to a startling stimulus.

These reflexes are so invested in your self-preservation that you actually can’t think about them before you respond.
All this processing happens in the spinal cord, so that the control of muscles can be initiated before the pain is actually
perceived by the brain. Learned, or acquired reflexes on the other hand, come from experience. Like how you learn
to dodge obstacles while riding a bike or driving a car. That process is also largely automatic, but you learn those reflexes
by spending time behind the wheel, or behind the handlebars. And reflex arcs stimulate some muscles, while inhibiting
others. For example, the tack in your right foot ended up activating the motor neurons in your right hip flexors and
hamstring, causing that knee to bend and your foot to lift up. But it also told the quad muscles in your left leg to extend
and stand tall, allowing you to shift your body’s weight off the tack. Of course not all reflexes come from pain, as you’ve
probably experienced when a doctor tapped your knee and your foot kicked. Your muscles and tendons are very sensitive
to being stretched too far, or too fast, because that kind of movement can cause injury. So for this we have receptors
called muscle and tendon spindles that specifically sense stretching. If triggered by an over-stretch, they generate a reflex
arc that contracts the muscle to keep it from stretching further. So, when does the brain actually get involved in all this?
Well, when your spinal cord sent impulses down the motor neurons, it also sent signals up your spinal cord toward the
brain.

News of the tack arrived first at your thalamus, the information switchboard that then split the message and sent it to the
somatosensory cortex -- which identifies and localizes the pain, like: “sharp, and foot”; as well as the limbic system, which
registers emotional suffering -- like, “why tack? Why me?!” And it also went to the frontal cortex, which
made sense of it all, assigning meaning to the pain -- like, “oh, I see this tack fell from the Crash Course poster on the wall
here.”
So basically, although your body has been reacting all along, it’s not until those pain signals hit the brain that you have the
conscious thoughts of both “dang, that hurt,” and “oh, that hurt because I stepped on a specific pointy thing.“ And this is
where I want to point out that we here at Crash Course cannot be held responsible for any injuries sustained in the
process of owning a Crash Course poster. Enjoy them at your own risk.

The Autonomic Nervous System

Your autonomic system is the branch of your peripheral nervous system that regulates the functions of your internal
organs, like your heart and stomach, and also controls your smooth and cardiac muscles, and your glands. All things that
you you do not consciously control, so, yes, you could say it has a lot of power over you. But the funny -- or maybe
confusing -- thing about this system is that its effects on your organs and muscles and glands are by no means consistent.
Not at all.
At any given moment, whether you happen to be totally relaxed or completely flipping your wig, your autonomic system is
constantly making involuntary, fine-tuned adjustments to your body, based on what signals your central nervous system is
picking up. This could mean changing your body temperature, sending extra blood to a particular area, slowing your heart
beat, or tweaking your stomach secretions. Its effects change, depending on the situation you’re in -- and also which part
of your autonomic system is in charge at that moment. Because this weird little corner of your nervous system that keeps
you alive is actually run by two competing interests. Two divisions that serve the same organs, but they create opposite
effects in them, battling it out back and forth, to either excite your body’s functions or subdue them.

One of them is dedicated to amping you up and preparing you for activity -- that’s your sympathetic nervous system. And
the other one talks you down and effectively undoes what its foil did. And that is the parasympathetic nervous system.
Together, they’re what make your body experience stress, fear, relaxation, and defiance. Courage and cowardice. Panic
and peace. If there’s an epic novel going on in your body right now it is probably being written by these two. Let’s talk
about names for a minute.

One of the two divisions of your autonomic nervous system is called the sympathetic system. That sounds kinda nice,
doesn’t it? It’s like understanding, and calming, and telling you that it’s not so bad after all. WRONG! Contrary to its
comforting name, the sympathetic system is what sounds your internal alarm bells. It’s the hardware behind the famous
“fight or flight” response. It is synonymous with stress. If the sympathetic is for “fight or flight,” the parasympathetic is for
“resting and digesting” -- it’s responsible for maintaining your body and conserving energy for later.

I recognize that this is confusing. But, when you explore the anatomy of these two systems, like we’re gonna do today,
they start to make a little more sense. Because, even though their basic components are essentially the same, their
physical structures turn out to be different in a few really important ways. And those differences can help explain why
they act like the foils that they are, and why sometimes you feel more like Sherlock than Watson, or the other way
around.

First big difference: the nerves of these two divisions originate at different sites in your body. Your sympathetic fibers are
thoracolumbar -- meaning that they originate from between your thoracic vertebrae where your ribs attach, and the
lumbar vertebrae just inferior to your ribs. Early anatomists saw how a network of nerves radiating from the middle of the
spine like this could quickly coordinate the functions of many major organs at once. So it was called the sympathetic
system, from the Greek words for “feeling together.” But the nerve fibers of your parasympathetic system begin both
above and below where the sympathetic ones do. They’re craniosacral, meaning they sprout from the base of your brain
and also from your sacral spinal cord, just superior to the tailbone. And because the roots of these nerve fibers basically
frame the starting points of the sympathetic nerves, they were called parasympathetic -- literally “beside the
sympathetic.”

Another difference between these two foils? Their ganglia. Unlike your sensory or motor neurons, where a single axon
can reach all the way from your spinal cord to whatever muscle or touch receptor it works with, both parts of your
autonomic system require two neurons in order to work. And those two neurons meet in ganglia -- clusters of neuron cell
bodies that house millions of synapses. But where these ganglia appear relate to their function, and which division of the
autonomic system they're serving.
Sympathetic ganglia are found closer to the spinal cord, because in those fight-or-flight moments of high excitement or
activity, they need to be able to send a single message far and wide, like the Bat Signal. This way, excitatory signals
traveling into a ganglion near the spine -- ganglion being the singular of ganlia -- can trigger action potentials in a whole
bunch of other neurons that lead to many different effectors, like the heart, and lungs, and stomach, and adrenal glands.
By contrast, most parasympathetic ganglia are found way out from the spine -- near, or even inside of their effector
organs. Because this system is responsible for taking care of particular functions only when you have the time and energy
to do it -- like digesting food or excreting waste -- it uses more specific, strategic signals. It’s more like Commissioner
Gordon calling Batman on the batphone, one on one just to talk about how things are going and you know whether
Alfred’s doing OK after his meningitis. It’s a private conversation -- not everybody needs to be involved. Anyway, because
the ganglia of these two divisions appear in different places in your body, it also makes sense that their neurons
themselves have slightly different forms, namely the length of their axons. Now, ganglia can be kind of complex -- it
actually comes from the Greek word for “a knot in a string” -- so when dealing with neurons around these structures, we
look at the fibers before they run into the ganglion, as well as after they come out of it. Understandably enough, the axon
lengths of the neurons before the ganglion are called the preganglionic fibers, and the ones coming out are
postganglionic.

The key here is that, in the sympathetic system, the preganglionic fibers are much shorter than the postganglionic ones.
Which makes sense, when you think about it, because sympathetic ganglia are really close to the spinal cord, and the
axons don’t have, or need, very far to go from the central nervous system. But they do have a lot of distance to cover, on
the other side of the ganglion, in order to reach their effectors. So naturally the fibers leading out of the ganglia are a lot
longer. And, foils being what they are, the reverse is of course true for the parasympathetic system. Since
parasympathetic ganglia are so close to, or even inside of, their effector organs, the preganglionic fibers are a lot longer.
They extend from the cranium and sacrum where they start, out to the lungs or liver or bladder -- wherever their effector
is -- where they reach their ganglion.

From there, the postganglionic fibers are super short -- just long enough to communicate with their effector. So, once
again -- it’s anatomy and physiology -- the structure of each of these systems is related to its function. The sympathetic
nervous system is set up in such a way that even a small stress signal sent down one path could trigger a response in
many effectors at once. Which is one reason why your reaction to a sudden, stressful event can feel so all-encompassing.
By the same token, the resting and digesting that’s overseen by the parasympathetic system doesn’t require urgent, all-
hands-on deck communication. If you need to process a burrito or take a nap or maybe a trip to the bathroom, it can
communicate with the organs involved, one on one. But still none of this tells us how these systems do what they do --
how these nerves communicate with your organs, and muscles, and glands in times of either stress, or relaxation. We’ll
start that next week, with a white-knuckle ride through your sympathetic nervous system.
So, between now and then, rest up.

Sympathetic Nervous System

So you’re sound asleep, when your smoke alarm goes off. Before you even know what’s going on, you start to feel it.
Those smoke alarms are loud -- for a good reason. Your heart starts to race, your breathing picks up, you become sweaty
all over your body. You are stressed. And I’m not talking about the my-iPhone-just-died kind of stress. I’m talking about
the I’m-afraid-I-might-die kind of stress. Even though it’s often seen as a dirty word, stress, like pain, isn’t all bad -- it’s
actually very useful if you’re, y’know, trying to get out of a burning building. Your sympathetic nervous system is the part
of your nervous system that responds to stress, and it does its job exceedingly well by focusing on what your body needs
to do right now. Like, when you’re facing a life-or-death ordeal, you don’t need to be digesting that cashew cluster in your
intestines, or producing reproductive cells, or fighting off an infection. That’s all stuff that you can deal with later, when
you’re out of harm’s way.

So your sympathetic nervous system sweeps these suddenly trivial functions aside to blast all of your energy to your brain
and heart and muscles to deal with the threat at hand. So, this is where I tell you that you’re lucky to have a sympathetic
nervous system. And that it keeps you alive. And that you would probably die in X Period of Time if you didn’t have one.
All of which is true. But here’s the thing: the problem is, nowadays our bodies’ stress responses are triggered all the time,
practically every day, even when we are not in mortal danger. I mean, worrying about paying your wireless bill or being
late for an important meeting -- those things are terrible, but they will not kill you. But, good luck explaining that to your
nervous system.

Because your physiological responses to non-immediate stresses are largely the same as when you’re fighting for survival.
So, if stress is, like, ruining your life, that’s why. And that’s part of the reason that should get to know how it works.
Because by learning about your sympathetic nervous system, you come to understand one of the key players in the
physiology of stress. You may recall from our tour of the anatomy of your autonomic nervous system, that in both your
sympathetic and parasympathetic divisions, almost every signal has to cross two synapses. Each neuron travels from its
root in the spinal cord to a ganglion, where it synapses – and yes, that is a verb as well -- with another nerve fiber. And
that one, in turn, leads to an effector organ, where it synapses again to create whatever response was signaled – like
sending more blood to your skeletal muscles, or making your heart pump faster. But you gotta wonder -- or at least I gotta
wonder: how do these neurons and effectors actually communicate with each other? And how do all of those signals
result in the high-octane sensations that we know as “stress”? By and large, the stress response includes two kinds of
chemicals, both of which I’m sure you've heard of.

The first, of course, are neurotransmitters. These are made and released from neurons themselves, and like we talked
about in our lesson about synapses, they are what neurons use to communicate with each other -- or their effector
organs -- across a synapse. The other chemicals involved in stress are hormones, which are secreted by your glands. There
are at least 50 different hormones at work in your body right now, and they do everything from regulating your sleep
cycles to making your body retain water so you’re not dying of dehydration all over the place. I’m telling you all of this
now, up front, because hormones and neurotransmitters are 100% necessary for understanding how your sympathetic
division ultimately works. BUT! When you trace a single sympathetic signal, from the initial stimulus to the final response,
those chemicals can be kind of hard to keep track of. That’s because the very same substance can have different effects --
actually, sometimes, totally opposite effects -- depending on where it’s received in your body.

And to make things even more fun, even though neurotransmitters are part of your nervous system, and hormones are
products of your endocrine system, a compound can be considered either a neurotransmitter or a hormone -- even
though it hasn’t changed one iota – depending on where it happens to be operating in your body. So all of this can make
understanding your stress responses pretty confusing! You might even say … stressful! All right, we’re going in. The smoke
alarm wakes you up. You smell smoke. It is time to move muscles. Fast. Your brain sends action potentials down your
spinal cord and preganglionic neuronal axons. Those signals flow all the way to their ganglia. When the signals reach the
synapses inside the ganglia, the nerve fibers then release a neurotransmitter -- called acetylcholine, known to its friends
as ACh. If you haven’t heard of acetylcholine yet, you’re gonna wanna remember that name.

In addition to working in sympathetic ganglia like this, it’s also what the rest of your peripheral nervous system and lots of
your central nervous system uses to communicate. So when it comes to nervous communication, ACh is really the coin of
the realm. The premium currency. So, that acetylcholine crosses the synapse and, if there’s enough of it, it can stimulate
action potentials in several neurons on the other end -- in the postganglionic fibers. That’s all it does, but it’s important.
It’s basically a signal booster. Those postganglionic neurons then carry the action potential to the effector organs – in this
case, let’s say your leg muscles, which are going to need an influx of blood if they’re going to hustle you out of that house.
And at the end of that second, postganglionic neuron, the fiber releases a different neurotransmitter. This one’s called
norepinephrine. And it is always norepinephrine that’s released from postganglionic fibers in the sympathetic nervous
system. It’s what crosses that final synapse and creates a response in the effector, like opening up blood vessels that lead
to the leg muscles.

So, the preganglionic fiber releases ACh, and the postganglionic releases norepinephrine. Boom. Congrats. Your life is on
its way to being saved. But, your body has more than one mechanism for responding to things, especially things like a
burning house. There’s another alternative for getting the message out. I mentioned those hormones, remember? In
addition to nerve fibers that lead to ganglia and then your effectors, there’s also a set leaving the spinal cord that goes
directly to your adrenal glands.

Like all preganglionic fibers, these release acetylcholine, too. But here, the signal doesn’t end up in another neuron that
triggers blood vessels to open or whatever. Instead, it triggers your adrenal medulla to release a flood of epinephrine and
norepinephrine – hormones that rush through your bloodstream toward your heart, lungs, and other organs. Now, hold
up! Did you notice what I just said? Yeah, I said the adrenal glands release norepinephrine as a hormone. Whereas in that
first scenario I said that norepinephrine was a neurotransmitter that sent the final signal to control blood flow to the leg
muscle. Now, how can I say both of those things? Because they’re both true.

Norepinephrine is BOTH a neurotransmitter and a hormone, and which one it is depends on how it’s being used. If it’s
being released from a neuron and travelling across a synapse, we refer to a messenger chemical -- no matter what it is --
as a neurotransmitter. If it’s being secreted by a gland into the bloodstream for more widespread distribution, it’s a
hormone. Even if it’s the same chemical. And to an effector, hormonal norepinephrine is just as good as neurotransmitter
norepinephrine. But as scientists, we describe them differently, because they’re functioning differently.

Now, the ways in which a neurotransmitter-slash- hormone like norepinephrine works, is a good example of another
confusing aspect of your sympathetic nervous system. Because it works by both stimulating and inhibiting the same
systems in your body at the same time! So, in our house-burning scenario, the norepinephrine your system releases
causes an increase of blood flow in some parts of your body -- like your leg muscles -- while restricting blood flow in other
places where it’s not urgently needed -- like your guts. How can the same chemical cause opposite responses? Well, it all
depends on the particular kind of receptors that an effector has for receiving that chemical.

In the case of norepinephrine, its effector is smooth muscle -- the muscle that controls all of your involuntary functions of
hollow organs, like the stomach, and bladder, and also your blood vessels. On the smooth muscle cells controlling some
blood vessels, there are receptors called alpha receptors -- when norepinephrine, or epinephrine, bind to those receptors,
they make those smooth muscle cells contract, thereby restricting blood flow. But on smooth muscle cells that control
other blood vessels, there are lots of beta receptors for epinephrine and norepinephrine, and when they are activated,
they make the muscles relax, letting more blood flow through. So it makes sense that the smooth muscle around your
blood vessels, which feed your skeletal muscles -- which you’ll need to get out of that smoky house -- are covered in beta
receptors. Because you want those blood vessels to relax, and provide plenty of oxygen to the muscles in your arms and
legs. And since running away is more important than digesting your dinner, the blood vessels leading to your stomach and
intestines have lots of alpha receptors, which reduce blood flow to those areas, because that burrito can wait until you’re
out of the house.

So, there’s a lot going on in your sympathetic responses. And much of it can seem complicated, or even contradictory.
But the thing is, all of these functions work together to create a full-body response, which is exactly what you need in an
emergency. After all, it wouldn’t do you much good to speed up your heart without sending that blood to your muscles,
where it’s needed. It’s up to those neurotransmitters and hormones, and the receptors on the corresponding effectors, to
make sure that everyone is on the same page. So, the system works well. Really well. Sometimes, too well. Remember
when I said at the beginning, how your body doesn’t know life-threatening stress from life-annoying stress? Since your
body’s reaction tends to be a full-body response either way, it can become pretty taxing over time. I mean, we’re talking
about throwing parts of your body into overdrive, while depriving others of blood and oxygen. That’s not something you
want happening every morning. So the irony here? The real kick in the head? It’s that non-life-threatening stressors can
actually end up endangering your life in the long run, because your body’s stress response is so effective. The frequent
activation of your sympathetic nervous system, and the triggering the other part of your stress response -- the part that’s
driven by hormones -- can have nasty consequences, like high blood pressure, digestive problems, and even the
suppression of your immune system. So what your body needs to do is figure out how to relax. Rest and digest. Feed and
breed. That is where your sympathetic system’s more mellow half-brother, the parasympathetic system comes in. And
yeah, that’s what we’re gonna be talking about next time.

Parasympathetic Nervous System

Consider your heart for a moment. For the average person -- at rest, like you probably are, sitting there watching me – the
heart beats at around 60 beats per minute. Once a second. Nice and easy. But if you were to somehow disconnect your
heart from your autonomic nervous system, things, as you might imagine, would change. But, your heart would not stop.
Actually, it would be the opposite. It would speed up. It would start beating at around 100 beats per minute -- and that’s
just at rest. With your heart beating two-thirds faster than normal, before you even broke a sweat, your cardiac muscle
would experience a lot of extra wear and tear. The surrounding blood vessels would be under enormous pressure. And
your body would suddenly require – and waste -- a lot of energy.
Basically, you’d be out of balance.

Part of what keeps your heart under control is your parasympathetic nervous system. It’s often described as the calming
side of your autonomic system -- a kind of antidote to the effects of stress created by the sympathetic system. But it’s
really much more than that. Unlike your sympathetic division, which lets you deal with the crisis of the now, the
parasympathetic system allows your body to handle … everything else. It not only calms you down after being stressed
out, it’s what allows you to digest food, to reproduce, to excrete waste, to fight off infections. Basically, it lets you do the
business of living. But our bodies can only do that when they are in balance, somewhere between excitement and
inhibition, both aroused enough and calm enough to keep things working.

So the parasympathetic system is why our hearts don’t pump so hard that they explode, sure. But it also explains a lot of
other stuff about our bodies. Oh, but just one thing? Learning about the parasympathetic system is going to involve a lot
of memorizing. Hope that doesn’t stress you out. You will recall that our sympathetic & parasympathetic systems not only
have different functions -- more or less engaging the same organs to opposite effects -- they also have different
structures. Their ganglia, for example, are located in different places: The sympathetic ganglia are located near the spinal
cord, while on the parasympathetic side, they’re close to the effectors. And likewise, the use of neurotransmitters in the
two systems is similar, but not quite the same.

In both systems, neurons release acetylcholine, or ACh, in their preganglionic synapses. But in your parasympathetic
system, the postganglionic neurons release ACh at their synapses with the effector organs, too... ...as opposed to in the
sympathetic system, where effectors get a dose of norepinephrine instead. But the biggest anatomical difference
between these two systems has to do with the physical networks that they form as they reach throughout your body.
While the sympathetic nerves all spring from the thoracolumbar area of your spinal cord, right around your midsection,
the nerves of the parasympathetic division are craniosacral. And with the exception of a couple of sacral nerves near the
tailbone that run to the bladder and genitals and rectum, most of these nerves never go through the spinal cord. Instead,
they run right from the brain almost all the way to their effectors.
There are 12 of these cranial nerves, and they vary in terms of what kinds of neurons they contain. I mean, we’re talking
about the autonomic system here, but they are not all autonomic motor fibers. Some of your cranial nerves also carry
motor fibers that control voluntary functions, like moving your eyeballs around. And others carry only sensory fibers,
which relay data to and from your sensory organs. And, you know, just to keep things interesting, some of your cranial
nerves carry both motor and sensory neurons. So, which ones are where? And what exactly does each one of these 12
nerves do? As anatomists, we have to keep track of the human wiring-diagram that are cranial nerves, because you don’t
want to end up like some sidekick in a ‘90s action movie who has to defuse a bomb all by himself. SHOULD I CUT THE RED
WIRE OR THE BLACK WIRE? Honestly, though, if you find yourself inside of somebody’s brain stem you probably shouldn’t
be cutting anything.

Since all 12 of these cranial nerves are important, you’re gonna have to come up with some kind of mnemonics to help
you keep track of both their names and their functions. You’ll need to know what each one is called, whether it’s a
sensory nerve, a motor nerve, or both. And the map that we follow of the cranial nerves is based on a ventral view of the
brain -- looking at its underside, with the anterior portion at the top, and posterior on the bottom.

First, let’s tackle the names. Starting at the top, the first cranial nerve you encounter is the olfactory nerve, which
takes scent information gathered by the nose and sends it to the brain. Followed by the optic nerve, which does the same,
but with visual data. Then there’s oculomotor, which controls four of the six muscles that control the movements of your
eyes. The next nerve, near the center of the brain’s ventral side, is the trochlear nerve, which controls a single muscle in
the eye, and it lets you do this. Just below that is the trigeminal nerve, the largest of the cranial nerves, which branches
into three main strands -- hence the ‘tri’ -- and innervates the face and jaw muscles. After that there’s the abducens,
which stimulates the muscles that let your eyes do this – from side to side, followed by the facial nerve, which operates
the muscles that make most facial expressions possible. Then there’s the auditory nerve. You can probably guess what
that’s for. You might notice that, up until the auditory nerve, the cranial nerves mostly control organs in the front of the
cranium -- mainly the eyes and facial muscles. But as you work your way down, the nerves tend to innervate the lower
and more posterior portions of the head. Like the glossopharyngeal nerve, which leads to your tongue and your pharynx.
That’s followed by your vagus nerve -- you should definitely remember that one – and then the spinal accessory nerve --
which has to do with moving your head and shoulders, and not whether your belt matches your shoes.

Lastly there’s the hypoglossal, the nerve that allows you to swallow and talk, among other things that you do with your
mouth and tongue. That was a lot of information and probably new words, so how are you gonna remember it all? Well,
by finding a way to remember the first letter of each name, in order. Which is: O-O-O .. T-T … A-F-A … G-V-S-H. That
doesn’t spell anything useful at all. There is a mnemonic that you’ll probably hear in school that goes like this: On old
Olympus’ towering top, a Fin and German viewed some hops. That’s pretty weird sounding -- not terribly easy to
remember. I mean, Olympus? Fin? Hops? There’s gotta be something more relevant to us 21st century science lovers.
Like, the Lord of the Rings fans out there might prefer something along the lines of: Onward old orcs! Toward the
Argonath for a Great Villain! Slay Hobbits! I’m just trying to help. Whatever device you use to remember the names of the
cranial nerves, you also have to keep track of their functions -- that is, whether they’re sensory, motor, or both. So, again
from top to bottom, a lot of teachers use this sequence of S’s, M’s and B’s to remember: Some say marry money, but my
brother says big brains matter more. That one’s not so bad. But I don’t know, maybe you’ll have better luck with
something like this: Sorry, Sherlock -- Mean Moriarty Beat Me, But Some Bobbies Busted Moriarty Masterfully! You are, of
course, invited to think up your own. And feel free to share them in the comments, hopefully there will be some good
ones down there -- anything would be better than Fins and Hops But if you’re going to commit one cranial nerve type to
memory, it should be 10, the vagus nerve. This long and extensive nerve stretches from near the brainstem down to most
of your visceral organs, including your heart, lungs, and stomach.

The vagus nerves work as a two-way street, ferrying incoming sensory information from the peripheral system to the
brain, and transmitting outgoing motor instructions from the brain to the rest of the body. So it’s a “B” nerve, because it
has “both” sensory and motor functions. And usually you don’t notice this nerve at work, because its functions are mostly
automatic. Say you’ve had a really stressful day, so your sympathetic system is charged up. You come home, crash on the
couch, mow down a half a pizza. Your stomach sends signals to your brain through the sensory nerve axons in your vagus
nerve, telling you that your belly is full of starch and protein and fat. Your brain sees that your stomach is churning away,
which is a usual parasympathetic activity, so it sends signals back down through the vagus nerve, triggering other
parasympathetic responses -- like slowing down your heart rate, putting some glucose back into storage, and reducing all
that norepinephrine that your sympathetic system was pumping out all day. Soon, you start feeling more relaxed. Which
is just one reason why, for some people, eating is a way of reducing stress and anxiety. In fact, it can feel so good that
even though your stomach is full, you might continue eating.

So, like I mentioned before, it can be easy to think of the two divisions of your autonomic system as opposites or even
rivals, but that’s a little off the mark. Looking at your body as a whole, you should picture them as two sides of a scale –
sometimes it’s balanced in the middle, and sometimes it leans to the left or right, depending on what's happening. That
balance is the essence of homeostasis, and as you’ll recall, homeostasis is the key to life. Here’s something else that’s
important for life: sex. It mostly falls within the parasympathetic domain of “necessary but not an emergency.” But in
order to effectively do it, you need help from both systems. First, the parasympathetic system has to make sure you’re
calm enough to even think about sex, and then funnel extra blood away from your muscles and down to your genitals –
which is why too much stress and anxiety can lead to sexual dysfunction. But you also need a burst of that sympathetic
system to excite you, and keep you excited. So like two sides of the scale, the balance depends on having the right
amount of both. The rate of action potentials travelling through each division is known as your “sympathetic tone” and
your “parasympathetic tone.” And, most of the time, our parasympathetic tone is actually dominant, keeping down the
caged animal that is your sympathetic response. That’s why you need your parasympathetic system to keep your heart
from racing like a rabbit’s. And why, most of the time, our bodies can do the eating, and sex-having, and all of the other
fun tasks that make up the business of living.

The Heart (part 1)- under pressure

Your heart, that throbbing, beating muscle, is probably the most iconic organ in your body. No other organ gets its own
holiday, or as much radio play. And you’re not likely to get a love note decorated with a kidney or a spleen, or even a
brain, which is really what rules the emotions. Don’t get me wrong, the heart does some great things -- namely, it powers
the entire circulatory system, transporting nutrients, oxygen, waste, heat, hormones, and immune cells throughout the
body, over and over. But in the end, the heart does not make you love. It doesn’t break apart if you get dumped by your
boo. And it’s not a lonely hunter. The truth is, the heart is really just a pump -- a big, wet, muscley brute of a pump. And it
doesn’t care about poetry or chocolate, or why you’re crying. The heart only has one concern: maintaining pressure. If
you’ve ever squeezed the trigger on a squirt gun or opened up a shaken can of soda, you’ve seen how fluids flow from
areas of high pressure -- like inside the gun or the can -- to areas of low pressure, like outside.

The heart’s entire purpose is to maintain that same kind of pressure gradient, by generating high hydrostatic pressure to
pump blood out of the heart, while also creating low pressure to bring it back in. That gradient of force is what we mean
when we talk about blood pressure. It’s basically a measure of the amount of strain your arteries feel as your heart moves
your blood around -- more than five liters of it -- at about 60 beats per minute. That’s about 100,000 beats a day, 35
million a year, 2 to 3 billion heart beats in a lifetime, the basic physiology of which you can easily feel, just by taking your
own pulse. I don’t have a watch Now, that might not inspire a lot of poetry, but it turns out, it’s still a a pretty good story.
Let us begin with a little anatomy. Unless you happen to be of the Grinch persuasion, the average adult human heart is
about the size of two fists clasped together -- one of the few bits of trivia you often hear about human anatomy that is
actually true.

The heart is hollow, vaguely cone-shaped, and only weighs about 250 to 350 grams – a pretty modest size for your body’s
greatest workhorse. And although Americans tend to put their right hand over their left breast while pledging allegiance,
the heart is actually situated pretty much in the center of your chest, snuggled in the mediastinum cavity between your
lungs. It sits at an angle, though, with one end pointing inferiorly toward the left hip, and the other toward the right
shoulder. So most of its mass rests just a little bit left of the midsternal line. The heart is nestled in a double-walled sac
called the pericardium. The tough outer layer, or fibrous pericardium, is made of dense connective tissue and helps
protect the heart while anchoring it to some of the surrounding structures, so it doesn’t like bounce all over the place
while beating. Meanwhile, the inner serous pericardium consists of an inner visceral layer, or epicardium -- which is
actually part of the heart wall -- and an outer parietal layer. These two layers are separated by a thick film of fluid that
acts like a natural lubricant, providing a slippery environment for the heart to move around in so it doesn’t create friction
as it beats. The wall of the heart itself is made of yet more layers, three of them: that epicardium on the outside; the
myocardium in the middle, which is mainly composed of cardiac muscle tissue that does all the work of contracting; and
the innermost endocardium, a thin white layer of squamous epithelial tissue.

Deeper inside, the heart has a whole lot of moving pieces that I’m not going to pick apart here, because the really big
thing to understand is how the general system of chambers, and valves, veins, and arteries all work together to circulate
blood around your body. Of course fluid likes to move from areas of high pressure to areas of low pressure, and
the heart creates those pressures. Form once again following function.

Your heart is divided laterally into two sides by a thin inner partition called the septum. This division creates four
chambers -- two superior atria, which are the low pressure areas, and two inferior ventricles that produce the high
pressures. Each chamber has a corresponding valve, which acts like -- like a bouncer at a club at closing time -- like he’ll let
you out, but not back in. When a valve opens, blood flows in one direction into the next chamber. And when it closes,
that’s it -- no blood can just flow back into the chamber it just left. So if you put your ear against someone’s chest -- and
yeah, ask for permission first -- you’ll hear a “lub-DUB, lub-DUB”. What you’re really hearing there are the person’s heart
valves opening and closing. It’s a relatively simple, but quite elegant set up, really.

Functionally, those atria are the receiving chambers for the blood coming back to the heart after circulating through the
body. The ventricles, meanwhile, are the discharging chambers that push the blood back out of the heart. As a result, the
atria are pretty thin-walled, because the blood flows back into the heart under low pressure, and all those atria have to do
is push it down into the relaxed ventricles, which doesn’t take a whole lot of effort. The ventricles are beastly by
comparison. They’re the true pumps of the heart, and they need big strong walls to shoot blood back out of the heart
with every contraction. And the whole thing is connected to the rest of your circulatory system by way of arteries and
veins. We’ll go into a whole lot more detail about these later, but the thing to remember first, if you don’t already
remember it, is that arteries carry blood away from the heart, and veins carry it back toward the heart. To differentiate
the two, anatomy diagrams typically depict arteries in red, while veins are drawn in blue, which, incidentally, is part of
what has led to the common misconception that your blood is actually blue at some point. But, it isn’t. It is always red. It’s
just a brighter red when there’s oxygen in it.

So let’s look at how this all comes together, starting with a big burst of blood flowing out of your heart. The right ventricle
pumps blood through the pulmonary semilunar valve into the pulmonary trunk, which is just a big vessel that splits to
form the left and right pulmonary arteries. From there -- and this is the only time in your body where deoxygenated blood
goes through an artery -- the blood goes straight through the pulmonary artery into the lungs, where it can pick up
oxygen. It finds its way into very small, thin-walled capillaries, which allow materials to move in and out of the blood
stream. In the case of the lungs, oxygen moves in, and carbon dioxide moves out. The blood then circles back to the heart
by way of four pulmonary veins, where it keeps moving to the area of lowest pressure -- because that is what fluids do --
and in this case that’s inside the relaxed left atrium. Then the atrium contracts, which increases the pressure, so the blood
passes down through the mitral valve into the left ventricle.

So the thing that just happened here, where a wave of blood was pumped from the right ventricle to the lungs and then
followed the lowest pressure back to the left atrium? There is a name for that, it is the pulmonary circulation loop. It’s
how your blood unloads its burden of carbon dioxide into the lungs, and trades it in for a batch of fresh oxygen. It’s short,
it’s simple -- at least in the way I have time to describe it -- and it’s just delightfully effective. Of all of the substances you
need to continue existing, oxygen is the most urgent – the one without which you will die in minutes instead of hours, or
days, or weeks. But it’s pretty useless unless the oxygen can actually reach your cells. And that hasn’t happened yet. For
that, your newly oxygenated blood needs to travel through the rest of your organ systems and share the wealth. And that
fantastic journey -- known as the systemic loop -- begins in the left ventricle, when it flexes to increase pressure. Now the
blood would like to flow into the nice low pressure left atrium where it just came from, but the mitral valve slams shut,
forcing it through the aortic semilunar valve into your body’s largest artery -- nearly as big around as a garden hose -- the
aorta, which sends it to the rest of your body. And after all your various greedy muscles, and neurons, and organs, and the
heart itself have had their oxygen feast at the capillary-bed buffet, that now-oxygen-poor blood loops back to the
heart, entering through the big superior and inferior vena cava veins, straight into the right atrium. And when the right
atrium contracts, the blood passes through the tricuspid valve, into the relaxed right ventricle, and right back to where we
started.

This whole double-loop cycle plays out like a giant figure eight -- heart to lung to heart to body to heart again -- and runs
off that constant high-pressure, low-pressure gradient exchange regulated by the heart valves. So the first “lub” that you
hear in that lub-DUB is made by the mitral and tricuspid valves closing. And they do that because your ventricles contract
to build up pressure and pump blood out of the heart. This high pressure caused by ventricular contraction is called
systole. Now, the “DUB” sound -- and, just to be clear, I am not talking about dubstep sounds -- that’s the aortic and
pulmonary semilunar valves closing at the start of diastole. That’s when the ventricles relax, to receive the next volume of
blood from the atria. When those valves close, the high-pressure blood that’s leaving the heart tries to rush back in, but
runs into the valves. So you know when you get your blood pressure measured, and the nurse gives you two numbers,
like, 120 over 80? The first number is your systolic blood pressure -- essentially the peak pressure, produced by the
contracting ventricles that push blood out to all of your tissues.

The second reading is your diastolic blood pressure, which is the pressure in your arteries when the ventricles are relaxed.
These two numbers give your nurse a sense of how your arteries and ventricles are doing, when they’re experiencing both
high pressure -- the systolic -- and low pressure -- the diastolic. So if your systolic blood pressure is too low, that could
mean that, say, the volume of your blood is too low -- like, maybe because you’ve lost a lot of blood, or you’re
dehydrated. And if your diastolic is too high, that could mean that your blood pressure is high, even when it’s supposed to
be lower. Considering how much we’ve talked about the importance of homeostasis, it should come as no surprise that
blood pressure that’s too high or too low, or anything that affects your blood’s ability to move oxygen around can be
dangerous. Prolonged high blood pressure can damage arterial walls, mess with your circulation and ultimately endanger
your heart, your lungs, brain, kidneys, and nearly every part of you. So I guess you could say the best way to break a heart
is to mess with its pressure.
But good luck trying to write a song about that.
The Heart (part2)

You’ve seen this before in a TV show or a movie. A patient is on the table in the emergency room, bleeding from a stab
wound or having ODed on drugs, when suddenly, an alarm goes off and the beep beep beep on the heart monitor
flatlines. It’s cardiac arrest, and you can tell just by that sound that it’s terrible. Doctors swarm around, barking orders,
and McDreamy or Dr. Grey -- or whoever the latest foxy doctor is on the TV show that you’re addicted to -- grabs the
paddles, pushes them down on the patient’s chest, and is all CLEAR! Then the patient’s chest jumps up, and everyone
stares at the monitor waiting for those steady beeps to reappear. That is a pretty classic scene. When it comes to medical
crises you see in popular culture, it’s probably only rivaled by the pool-side CPR scene, or someone shaking an
unconscious body, pounding on their chest and saying “Don’t you die on me!” Which hopefully you know is no way to
keep someone alive.

These are classic script tropes, but they contribute to some misconceptions about how defibrillators, CPR, and the
electricity of the heart work. Because the truth is, CPR can help prolong heart function during cardiac arrest, but it usually
can’t save a life without help from a defibrillator. And when McDreamy -- may he rest in peace -- finally bursts out those
paddles, that high-voltage shock isn’t turning the heart back on -- it’s actually stopping it. Confused? Well, I’m here to get
your head to understand your heart. We’ll get back to hot TV doctors in a coupla minutes, I promise. But in order to
understand what’s actually going on during cardiac arrest, we have to understand some basics about your heart cells.

We’ve learned a lot about skeletal muscle tissue -- how it’s striated, and contracts using the actin-myosin sliding filament
dance you’ve heard so much about. Your cardiac muscle is also striated, and uses sliding filaments to contract, but the
similarities end there. For one, their cells look pretty different. Skeletal muscle tissue has long, multinucleate cells, while
cardiac cells are squat, branched out, and interconnected, each one with one or two central nuclei. The cells are
separated by a loose matrix of connective tissue called the endomysium, which is chock full of capillaries, to serve up a
constant supply of oxygen.

Cardiac cells are also loaded with energy-generating mitochondria. In fact, mitochondria take up as much as 25 to 35
percent of each cell, making it resistant to fatigue, which is partly why your heart can beat nearly 3 billion times in a
lifetime. Not surprisingly, the differences between skeletal and cardiac muscle tissues are key to understanding their
functions. Skeletal muscle fibers are both structurally and functionally separate from each other, meaning that some cells
can work while others don’t -- that’s why you can grasp a delicate flower with the same hand that you can use crush a
soda can. Cardiac cells, on the other hand, are both physically and electrically connected, all of the time. It takes precise
coordination to create the high and low pressures required to pump your blood after all, and cardiac cells need to be
linked in order to have that perfect timing. And there’s one more thing you need to know about your heart’s cells:
Some of them can generate their own electricity. How in the name of Raymond de Vieussens can that be? Well, rewind
your brain to when we explored the electrical marvel that is the action potential, and how it triggers both neurons and
muscle cells. That process started by depolarizing the cell -- that is, pushing the cell’s membrane potential from negative
toward positive, past a threshold that triggered voltage-gated ion channels to open.

Most cells in your body only depolarize after being triggered by an external stimulus, or by a neighboring cell, in a long
chain reaction of action potentials that’s set off by the nervous system. But that is not the case for a special group of cells
found only in your heart – ones that can trigger their own depolarization. These are your pacemaker cells. Pacemaker cells
are what keep your heart beating at the correct rhythm, and ensure that each cardiac muscle cell contracts in
coordination with the others, because you don’t want your brain to have to send a series of action potentials every time
you need your heart to beat. Your brain has got other stuff to do.
So pacemaker cells are, in a way, your heart’s very own brain, generating the initial spark that sends a current through
your heart’s internal wiring system, known as the intrinsic cardiac conduction system. This system transmits electricity
along a precisely-timed pathway that ends with atrial and ventricular contractions -- also known as heart beats.
And it begins with pacemaker cells generating their own action potentials.
In most cells, the action potential starts with the resting potential, which the cell maintains by pumping sodium ions out
and potassium ions in, Right? Then, when some stimulus causes the sodium channels open up, the sodium ions flood back
in, which raises the membrane potential until it reaches its threshold. Pacemaker cells operate the same way, except for
that initial stimulus. They don’t need it. Their membranes are dotted with leaky sodium and potassium channels that
don’t require any external triggers. Instead, as their channels let sodium ions trickle in, they cause the membrane
potential to slowly and inevitably drift toward its threshold. Since the leaking happens at a steady rate, the cells fire off
action potentials like clockwork. And the leakier the membrane gets, the faster it keeps triggering action potentials. The
pacemaker cells at the start of the conduction system have the leakiest membranes, and therefore the fastest inherent
rhythms, so they control the rate of the entire heart. And those fast, leaky cells are found in the sinoatrial node, or the SA
node, up in the right atria. They essentially turn the whole SA node into your natural pacemaker. After those pacemaker
cells make themselves fire, they spread their electrical impulses to cardiac muscle cells throughout the atria.
The impulses leap across synapse-like connections between the cells called gap junctions, and continue down the
conduction system until they reach the atrioventricular node, or AV node, located just above the tricuspid valve. Now,
when the signal hits the AV node, it actually gets delayed for like, a tenth of a second -- so the atria can finish contracting
before the ventricles contract. Without that delay, all the chambers would squeeze at once, and the blood would just
splash around and not go anywhere. So instead, the atria contract and blood drops down into the ventricles, and then a
moment later, the signal moves on and triggers the ventricles to squeeze, making the blood flow out of the heart. And
there are two tricks to a good ventricular contraction. One, the ventricles are so large that the signal has to be distributed
evenly to ensure a coordinated contraction. And two, the ventricles need to squeeze like their squeezing a tube of
toothpaste – from the bottom up -- to accelerate the blood through the big arteries at the top of the heart.

So from the AV node, the signal travels straight down to the inferior end of the heart and gets distributed to both sides.
The path the electrical impulse takes to the bottom of the heart is called the atrioventricular bundle, also known by the
more rad name, the bundle of His, where it branches out to the left and right ventricles. Finally, the signal disperses out
into Purkinje fibers, which trigger depolarization in all surrounding cells, causing the ventricles to contract from the
bottom up like toothpaste tubes, at which point the whole cycle starts all over again. And everything I just described to
you -- from when the SA node fires to when the last of the ventricular cells contract -- takes about 220 milliseconds. So
that is how your heart beats. But I know what you want -- you want to get back to talking about TV shows and McDreamy
and his paddles. It’s totally understandable. OK, so picture all your individual heart cells as a bunch of musicians in an
orchestra. They all sound really great together, but then the conductor suddenly needs to go to the bathroom. And it
sounds OK at first, but then the tuba gets a little weird, and the triangle is half a beat off, and soon everyone is playing a
different note at a different time, jamming to their own personal rhythm. In the heart, we call this out-of-sync behavior
fibrillation, and it can be caused by all sorts of problems, especially ones that affect the pacemaker cells in the SA node. In
an orchestra, this just sounds really terrible. In a heart, there’s no coordinated contraction, no lub-dub, no blood moving
through the body. Which means you will soon be dead.
But then the conductor comes back from the bathroom break, taps her wand, and everybody stops. It’s silent for a second
before the wand comes up, and then they all start playing again, this time in unison. If your heart in fibrillation is an out-
of-sync orchestra, then a defibrillator is that conductor. It stops the chaotic noise by overriding all the individuals, and hits
a sort of reset button so everyone can start again on the same page. The paddles send so much electricity through the
heart that they trigger action potentials in all of the cells at once. Then, the cells repolarize, and start leaking again, and
then the most leaky cells, in the pacemaker SA node, reach their threshold and fire first, re-setting the rhythm that keeps
everyone in harmony so your heart functions properly. And that is how hot doctors and their paddles actually stop hearts
to save lives. Now the thing about CPR -- or cardiopulmonary resuscitation -- is that it can’t correct fibrillation. What those
chest compressions can do is force a fibrillating heart to keep circulating oxygenated blood until help arrives. But if a
person is in cardiac arrest, just breathing into their mouth and compressing their chest won’t deliver the electricity
needed to give the pacemaking cells a chance to reset.

Blood Vessels Part 1 (Form and Function)

No doubt about it, your heart is a champion. It electrifies itself, it maintains your blood pressure, it keeps your blood
moving, and it’s got like a nice shape you can put some chocolates inside of and give to people you like. But the
circulatory system is much, much more than just that pump. Because the heart also needs a network to actually send all
that blood through, right? Cue the blood vessels. Although it’s easy to think of them as a glorified plumbing system for
your body, that’s not a very good analogy. These aren’t just passive tubes made merely to carry liquid around, like the
pipes behind your walls at your home.

Blood vessels are actually active, dynamic organs, capable of contracting and expanding as they deliver oxygen and
nutrients to cells throughout the body; carry away waste products; and do their part in maintaining that all important
blood pressure. You already know about the three major types of blood vessels: the arteries that carry blood away from
the heart, the veins that bring it back, and the little capillaries that act as the transfer station between the two. But you’ve
also got arterioles -- which are like mini-arteries that branch out into those capillaries -- and venules, the smallest vein
components that suck blood back out of the capillaries and merge into the larger veins that head home to the heart.
And it’s quite an incredible journey, really. If all your blood vessels could be strung together in a single line, they’d stretch
out for 100,000 kilometers! That’s like...if you...and then...carry the two -- that’s like two and a half times around the
Earth. And together this extensive network forms a closed system that begins and ends in the heart. That means that all of
the five or so liters of blood in your body are contained within it at all times, unless you’re bleeding, which I hope you’re
not.
If you prick a finger and watch a drop of blood pop up, you know that you’ve nicked a blood vessel, and that blood is
leaking out of its closed system. Likewise, if you slam your shin against a corner of the coffee table on your way to the
bathroom, and an hour later you see a big nasty bruise forming, then you know you’ve damaged your blood vessels again,
because bruising is internal bleeding, usually into loose connective tissue. And, if you’re embarrassed about the piercing
shriek that you let out when you bumped your leg, and you start to blush. Well, that’s your blood vessels, too, expanding
just to say hello.
Blood vessels are another great example of how anatomy and physiology go together like peanut butter and jelly. How
they look and what they do go hand in hand. Most of your blood vessels share a similar structure consisting of three layers
of tissue surrounding the open space, or lumen, that actually holds the blood. Anatomists call these layers “tunics,” and
the innermost section is the tunica intima -- which should be pretty easy to remember because, you know, it has like
intimate contact with the lumen. It’s like your circulatory underpants. The cool thing about this layer is that it contains the
endothelium, which you may remember is made up of simple squamous epithelium tissue and is continuous with the
lining of the heart. These cells form a slick surface that helps the blood move without friction. Surrounding the tunica
intima is the middle layer, the tunica media, made of smooth muscle cells and sheets of the protein elastin. That smooth
muscle tissue is regulated in part by the nerve fibers of the autonomic nervous system, which can decrease the diameter
of the lumen by contracting this middle layer during vasoconstriction, or expand it by relaxing during vasodilation. That
right there should tell you that the tunica media plays a key role in blood flow and blood pressure, because the smaller
the diameter of the blood vessel, the harder it is for blood to move through it -- kinda like trying to drink milk through a
cocktail straw versus a soda straw.

And finally, the outermost layer of your blood vessels is the tunica externa. It’s like an overcoat, if that coat were made
mostly of loosely woven collagen-fiber. Actually, if your coat happens to be made of leather, it is made of collagen. And
like a coat, this outer layer is what protects and reinforces the whole blood vessel. Now the ratio of the thicknesses of
three layers varies between blood vessels of different types -- because, guess what?!
Yes, form follows function! Let’s take a look. Say you’re gearing up for a big tournament of thumb wrestling, or what has
been called the “miniature golf of martial arts.” How does blood move through your systemic circulatory loop, to get from
your heart to your champion right thumb-flexing muscle, the flexor pollicis brevis, and back again?
Well, you will remember from our lessons on the heart that blood leaves the left ventricle through the aorta -- the biggest
and toughest artery in your body, roughly the diameter of a garden hose. The aorta and its major branches are elastic
arteries -- they contain more elastin than any other blood vessel type, so they can absorb the large pressure fluctuations
as blood leaves the heart. What’s more, that elasticity actually dampens that pressure so that big surges don’t reach the
smaller vessels, where they could cause damage. This is really where that whole pipe analogy falls apart. These arteries
are really more like a balloons -- they’re pressure reservoirs, able to expand and recoil with every heartbeat. If they were
rigid like pipes, they’d eventually leak or burst after being battered by so many waves of pressure.

So that blood leaves your aorta, and since it’s headed to your thumb, it travels along the elastic subclavian artery, which
gives way to a series of muscular arteries – in this case, the brachial artery in your upper arm, and the radial artery in the
lower arm. Muscular arteries distribute blood to specific body parts, and account for most of your named arteries. They’re
less elastic and more muscular. These arteries invest in additional smooth muscle tissue, and proportionally, have the
thickest tunica media of any blood vessel. This allows them to contract or relax through vasoconstriction and vasodilation,
which we’ve talked about a lot in terms of the nervous system’s stress response. These arteries keep tapering down until
they turn into the nearly microscopic arterioles that feed into the smallest of your blood vessels, your tiny, extremely thin-
walled capillaries which serve as a sort of exchange or bridge between your arterial and venous systems. They may be
little, but your capillaries are where the big, important exchange of materials actually happens.

Capillary walls are made of just a single layer of epithelial tissue, which form only the tunic intima, so they’re able to
deliver the oxygen and other nutrients in your blood to their cellular destinations through diffusion. The capillaries are
also where those cells can dump their carbon dioxide and other waste back into the blood and send it away, through the
veins to the lungs and kidneys. But I will come back to that in a second. Unlike arteries and veins, capillaries don’t operate
on their own, but rather form interweaving groups called capillary beds. Besides exchanging nutrients and gases, your
capillary beds also help regulate blood pressure, and play a role in thermoregulation. Say you’re in the room where
you’re, like, practicing thumb calisthenics -- which probably isn’t a thing -- but the room is a little chilly, so the blood
feeding your dermis loses a lot of heat to that cold air.
Well, smooth muscle forms tiny sphincters -- yeah, you’ve got sphincters everywhere -- around the vessels that lead to
each of your capillary beds. When they tighten up, they force blood to bypass some of those capillaries, which means less
blood is exposed to the cold, and you lose less heat. If it’s really cold, the smooth muscles around your larger arterioles
and muscular arteries -- like that radial artery in your lower arm -- will also squeeze, slowing blood flow to your whole
hand. Which is no way to win at thumb wrestling.
But it’s one reason why your fingers get all stiff and numb in the cold -- they’re not getting as much warm blood, because
your blood vessels are trying to conserve heat. Conversely, if your thumb is working really hard and producing heat from
all that exertion, those capillary sphincters relax and open wide, flooding the capillary bed with blood to help disperse
heat -- which is part of the reason that you might get red-faced when you’re hot or exercising hard. So anyway, now your
thumb muscles have just feasted on a batch of oxygen and glucose served up on a fresh bed of capillary, and they’re
ready to take out the trash.

The cells send their CO2 and other junk out to the venal end of the capillary exchange where the capillaries unite into
venules, and then merge into veins that head back to the heart. Remember that the pressure in these vessels has to be
dropping, since fluids always flow from higher to lower pressure. But since the pressure is so low in your veins -- it’s like
one 12th of the pressure in your arteries -- there isn’t much pressure gradient left to push the blood back to your heart.
So veins require some extra adaptations to keep the blood moving in the right direction. That’s why some of them
especially veins in the arms and legs that have to work against gravity -- have venous valves that help keep the blood from
flowing backward. If those valves leak, or a vein experiences too much pressure, the backflow of blood can stretch and
twist the vein, leaving you with varicose veins, or if this happens in another part of the body, hemorrhoids.

But, anyway, we’ve gotten pretty far from your thumb at this point. We’ve got a loop to finish here! From the capillaries
and venules in your thumb, that low-pressure blood flows from the radial vein to the brachial vein to the subclavian vein,
where it dumps into the superior vena cava and settles for a second in the right atria, before dropping into the right
ventricle. From there it’s sent to the lungs, where it gets oxygenated, and then comes back into into the left atria, before
sliding down into the left ventricle, where it builds up pressure again, and spurts back out into your aorta.

It takes about a minute for all the blood in your body to complete that circuit, which means, even if you’re mostly at rest,
your hardworking circulatory system moves about 7,500 liters of blood through your heart every day. Just in the time that
you’ve been sitting there listening to me, probably about 52 liters has coursed through. So yes. Much like the Internet,
your blood vessels are more than just “a series of tubes.” During the time that you’ve been circulating all that blood, you
learned about the basic three-layer structure of your blood vessels; how those structures differ slightly in different types
of vessels; and you followed the flow of blood from your heart to capillaries in your right thumb, and all the way back to
your heart again.

Blood Vessels Part 2

Why is everybody so worried about high blood pressure? I mean, of all of the millions of things that could go fatally wrong
in your body at any given moment, it doesn’t seem like the biggest threat you face is that your blood might be pumping a
little too hard. A heartbeat that’s too strong? Isn’t that like an awesome song that’s played too loud? Or nachos with too
much cheese?! Well, no. The fact is: You’re more likely to die from diseases related to your cardiovascular system than
anything else. And probably you would like to know why. Well, we know that blood pressure is the circulatory system’s
way of getting your five liters of blood flowing throughout your body, so that your tissues can get the oxygen and
nutrients they need. But chronic high blood pressure, or hypertension, can cause serious damage to both the heart
that creates the high pressure, and to the blood vessels that have to withstand that extra pressure.

Over time, the increased force of blood against the arterial walls can make them stiffen, leak, or rupture, while the heart
itself may simply wear out from all the extra work it’s doing to keep blood moving. Luckily, your nervous and endocrine
systems have some tricks to try and balance blood pressure when it gets too high or too low, in an effort to create
homeostatic balance. But, even with these defenses in place, if your blood pressure stays out of balance long enough,
things will start going really, really wrong. One in three adults in the U.S. has high blood pressure -- that’s two in six, that’s
three in nine -- and often there are often no symptoms until it’s too late. Just like your favorite song turned up to 11, or an
unlimited supply of nacho cheese, even your own life-giving blood can become…too much of a good thing.
Part of what might seem confusing about high blood pressure is the seemingly random list of things that can contribute to
it, some of which are easier to control than others. There’s emotional stress. Physical exertion. Dehydration. Too much
salt on your mashed potatoes. Or too much butter on your bacon. Why are we buttering bacon?

But all these things can in some way affect how efficiently your blood flows. But in order to understand why your heart
hates french fries, you first have to understand how blood flow works. When we talk about blood flow, what we’re talking
about is the volume of blood flowing through any given vessel, or through the circulatory system as a whole, per minute.
This is also called cardiac output, and it’s determined by the blood volume pumped during one beat, and the number of
beats per minute.

Now, if you’ve ever used a garden hose, or fixed a leaky sink, or had any other hands-on experience with fluids, then you
know that flow can change in response to a number of factors, especially those that affect resistance. Resistance is just
anything that hinders flow or creates friction. In the case of your blood, resistance can be the result of increased viscosity
– the thicker your blood is, the harder it is to move -- or it could be because of increased vessel length, since longer
vessels are more resistant to flow in general. But those factors tend to be pretty constant in your body over time. Instead,
for most people, the biggest factor that affects resistance has to do with vessel diameter.
Changes in diameter can be temporary, like during vasoconstriction or vasodilation – when the diameter increases or
decreases, allowing more or less blood through. But an excess of low-density lipoprotein, or LDL, the so-called “bad”
cholesterol, in the blood can build up to form a fatty plaque on the inside of your arteries, permanently increasing the
resistance, and hindering blood flow. So your blood pressure, blood flow, and resistance are all tied together. In fact,
they’re so closely and predictably tied together that we can even express their relationship mathematically.

Among the truths contained in this equation is the fact that blood flow increases as the difference in blood pressure
between two points increases. Remember: The ventricles of the heart create very high pressure, while the atria, or the
receiving chambers of the heart, have very little pressure, at least while the body’s at rest. So, the bigger the gap is
between the high pressure in your ventricles and aorta, and the low pressure in your vena cava and atria, the faster a liter
of blood will flow through your system. If you have to back up a few times to review that, nobody’s gonna judge you.
Now, at the same time, blood flow decreases as resistance increases. So, with a little algebraic rearranging, you can see
that blood pressure equals blood flow, or cardiac output, times resistance. In theory, this means that any change in
resistance or cardiac output would also change blood pressure. But when any one of these variables actually does change,
your body tries its best to compensate for it, in its eternal quest to maintain homeostasis. And it accomplishes this in a
few different ways, mainly using neurons, hormones, and the kidneys. One major short-term fix to wonky blood pressure
comes from your brain, which targets both cardiac output and resistance, by altering the distribution of the blood flow
around the body, or by changing the diameter of certain blood vessels.

This comes in handy when you’re walking up fourteen flights of stairs in your apartment building with three bags of
groceries. The vessels feeding your digestive organs constrict, which increases resistance there, so more blood goes to the
skeletal muscles in your legs. Most neural responses like this use baroreceptors, special nerve endings found in the carotid
arteries, the aorta, and other larger arteries in the neck. When blood pressure stretches arterial walls, that opens
mechanically-gated sodium channels in these little receptors. The higher the blood pressure is, the more frequently they
send action potentials to the midbrain informing it just how much pressure the artery is feeling. When the brain learns
what’s happening, it can do any number of things to correct the situation -- like dilate some arterioles to reduce
resistance, or reduce the heart rate to lower cardiac output -- and these things work pretty well for a while. But
baroreceptors are not effective for long-term pressure changes, in part because they end up adapting, essentially
reprogramming themselves to read a high blood pressure as the new normal. Now, other short-term effects on your
blood pressure come from your hormones.

We’ve already talked about one classic example: when your body actually needs a little high blood pressure -- like when it
has to get ready to fight or flee -- the adrenal medulla starts flooding your blood with epinephrine and norepinephrine.
These hormones raise both the heart rate and the blood volume, and therefore cardiac output, while also constricting
vessels in less essential regions, increasing the overall resistance, and therefore pressure. And again, just like with neural
controls, these hormonal controls work by changing vessel resistance and cardiac output. But the way to get more long-
term control of blood pressure is to alter the blood volume, and for that, you need the kidneys.

Your kidneys cook up hormones like renin and angiotensin which help regulate levels of sodium and fluids in your body,
and also help expand and constrict blood vessels. And when blood pressure gets too high, your kidneys will try to reduce
the volume of blood, by getting rid of any extra water. Basically, they make you pee. Blood volume, by the way, is partly
why sodium is the root of all evil in blood pressure treatments: The excess sodium used in processed foods and salty
snacks causes your body to retain water, which creates higher blood volume and leads to higher blood pressure, which,
you know, is a bummer. So, those are some of your body’s automatic solutions for high blood pressure. But how it
compensates for sustained high blood pressure will almost certainly lead to trouble -- either for the heart that has to work
harder to overcome the resistance, or for the vessels that have to take that extra pressure.

An increase in either blood flow or resistance leaves the heart struggling to do its job, so it might actually build more
muscle around that left ventricle to help generate the force needed to move the blood. Now, a more muscley heart may
not sound like a bad thing, but trust me, it is. Why? Well, because more muscle needs more oxygen, and your body just
can’t create new blood vessels to feed that super-sized ventricle. So that big muscle is left starving. It’s literally a hungry
heart.

Plus, if you have cholesterol plaques in your arteries, then you probably have them in your coronary arteries, which carry
oxygen and nutrients to the heart muscle itself. That increases the resistance, and therefore decreases blood flow to the
heart muscle. So you have this bigger, hungrier heart muscle, but its nutrient supply is diminished, and eventually those
heart muscle cells can slowly die.When that happens, it’s known as heart failure. Or, if you completely block one of those
coronary arteries with plaque, or maybe a blood clot, a whole bunch of heart cells quickly starve to death. And that can
lead to a myocardial infarction, which is what we also call a heart attack. And then we’ve got the problems on the blood
vessel end, under sustained hypertension, you may see your otherwise elastic arteries go from being flexible balloons to
stiff, hardened pipes, in what’s called arteriosclerosis. Or… the high pressure may make them weak and bulgy in spots,
until they’re stretched too thin and leak or burst. And that is an aneurysm. And if the weak spot ruptures in a smaller
arteriole blood vessel -- say, one that leads to an organ like a kidney or eye, it can lead to organ damage or failure.

So although having a huge, strong flow of blood coursing through your body might sound like some kind of comic book-
caliber superpower, it is not. It’s just too much of a good thing. That’s why you just learned, first and foremost that you
should not butter your bacon – and also what blood flow and resistance are, and how they relate, in direct proportion, to
blood pressure. We also looked at your body’s short term responses to high blood pressure -- including neural, hormonal,
and kidney response -- and wrapped up by describing all of the ways that chronic high blood pressure can kill you.
Blood part 1 (true blood)

Don’t take this the wrong way, but you’re pretty replaceable. When it comes to your body, science has figured out how to
hack, synthesize, or replace a surprising amount of its parts and processes. We have implants to keep heart beats steady,
and steel rods to mimic bones. We’ve got drugs that can replace hormones, and antibiotics to cover for your immune
system, and pretty soon you’ll be able to just 3D print a new ear if you need one. Really! But one thing we absolutely
cannot manufacture -- despite what True Blood would have you believe -- is blood. And yet blood is a thing that we all
need. And sometimes, because of injury or illness, we need extra blood. In fact, every two seconds, someone in the U.S.
needs a blood transfusion. This could be a victim of a car accident, someone undergoing surgery, or a cancer patient who
needs new blood to maintain their health during chemotherapy. And because we can’t grow it on trees, or make it in a
lab, or even it store it for all that long, the blood that people need -- nearly 16 million pints a year in the U.S. -- has to
come from people who have donated it. So let’s talk blood, shall we?

The meal of choice for vampires and female mosquitoes, blood is red, sticky, salty, and kind of metallic tasting. It is indeed
thicker than water, and super viscous -- which is why Hitchcock used chocolate syrup as a stand-in in a certain classic
shower scene. For most purposes, blood comes in eight different types, and it accounts for about 8% of your body weight.
You might remember from our episodes on tissues that blood is a type of connective tissue, which means it’s made of
living cells suspended in a nonliving matrix, which in this case is the fluid ground substance called plasma. And of course
one of blood’s main missions is to transport and distribute oxygen, nutrients, waste products, and hormones around the
body. But it also helps regulate and maintain body temperature, pH levels, and the volume of fluids in your body. Plus it
protects you from infection and from the loss of blood itself. Perhaps second only to your brain, your blood is the one
component of your body that we haven’t figured out how to reproduce, synthesize, or imitate. It’s a part of you that is
literally irreplaceable. It’s Saturday and you feel like doing a good deed, so you head over to your local Red Cross for a
blood drive. You get your finger pricked and then somebody directs you toward a lounge chair, swabs your
inner elbow with alcohol, and then comes at you with a hollow needle. Once the bag is full -- they usually take about a
pint -- you get unhooked and grab a cookie and a juice to replace the blood sugar you lost. And the whole process takes
around 20 minutes.

But for your blood, the day is just beginning. Soon it will be taken to a lab, where it’ll be tested for infectious diseases and
separated into different parts before heading out to hospitals. So, hold up: What exactly do I mean by different parts?
Well, the blood that flows from your arm into that bag is whole blood, a mixture of cells and cell fragments called formed
elements, along with water, and lots of dissolved molecules. A patient who needs a transfusion may only need some of
those things and not others, so the parts are separated. Once your blood makes it to a lab, technicians put it in a
centrifuge, which spins it around fast enough to send the heavier components to the bottom of the tubes, and bring the
less dense elements to the top. In the centrifuge, three distinct layers emerge. Down at the bottom you’ve got a heavy
red layer of erythrocytes, or red blood cells that carry oxygen and carbon dioxide. They make up about 45 percent of your
total blood volume. Then you’ve got this thin little whitish layer in the middle. Those are your warriors, the leukocytes or
white blood cells, that defend your body from toxins and foreign microbes. And there are also the cell fragments, called
platelets, which help with blood clotting and make up less than one percent of your blood.

Finally, up at the top you see the yellowish plasma, which accounts for about 55%of your blood volume. Plasma is actually
90 percent water, but the other 10 percent is chock full of 100 different solutes, including proteins, electrolytes, gases,
hormones, and waste products. The most of abundant of these solutes are electrolytes -- which you may have heard of as
the secret ingredient in sports drinks. But they’re really just positively-charged cations -- like calcium, sodium, and
potassium -- and negatively-charged anions, like phosphate, sulfate, and bicarbonate. Together these ions help regulate
your blood’s chemistry, maintaining its pH levels and proper osmotic pressure, and allowing other tissues to do their jobs,
like making muscles contract and sending action potentials. But when measured by weight, the bulk of the solutes in your
blood are really the plasma proteins.

Most of these proteins -- like albumin, and alpha and beta globulins -- are made by the liver, and do things like balance the
osmotic pressure between the blood and surrounding tissues, and transport lipids and ions. Others run defense for you,
like the gamma globulin antibodies that are released by plasma cells during an immune response, or fibrinogen proteins,
which are vital to forming blood clots and stopping bleeding. All right, bleeding. I want to talk about that. Because, for the
very reason that I mentioned at the beginning -- that we can’t replace your blood with some synthetic wonder-fluid -- the
LAST THING that your circulatory system wants is for you to fritter away your blood, in some sidewalk scrape or kitchen
accident. So, it has a whole system in place to prevent you from losing too much of it, through a process known as
hemostasis.
So imagine you’re slicing a nice garlic-cheese bagel one morning, and you lacerate the distal phalanx of your pollex -- in
other words, you cut the tip of your thumb. And now you’re bleeding all over your breakfast. At the very first sign of a
rupture, the blood vessel actually constricts itself, to slow the flow of blood through it. Then little cell fragments called
platelets gather at the site of the injury, creating a plug that dams the breech and keeps the blood from leaking further.
Now these free-floating platelets don’t clump together during regular circulation – that would be terrible -- but when the
endothelial cells lining a blood vessel wall tear, the underlying collagen fibers are suddenly exposed. And they chemically
react with the platelets, turning them all sticky and glue-like at the scene of the injury. But that platelet plug still isn’t as
strong as it could be -- it needs reinforcement to complete the clotting process. This reinforcement comes in the form of
fibrin threads, protein strands that join together to make a sort of mesh that traps the platelets and blood cells.

Eventually, the threads actually pull the opposite sides of the wound together, to close the vessel wall, so the endothelial
cells can be replaced. Over a few days, the blood vessel heals, and the blood clot dissolves. Or at least, that’s how it is
supposed to happen. People who suffer from disorders related to hemostasis may have trouble with unwanted clotting,
or the inability to form clots. In the family of disorders known as hemophilia, a patient can usually complete the first two
steps of hemostasis just fine, but they can’t make an effective fibrin clot. So it’s not that they bleed more than anyone
else, it’s just that they bleed longer. Which, I guess means that they bleed more.

As a result, they may need frequent blood transfusions throughout their lifetime. Which brings me right back around to
that Saturday morning blood drive. Another thing you’re going to need to know before you give blood is what type you
have -- do you have A, B, AB, or O? These different types all do the job equally well, they just sort of have a different flavor
related to your immune system.

All the cells in your body have a plasma membrane with specialized glycoprotein markers on them that act like name tags
or labels, sort of like “This cell is Property of Hank.” These markers are your antigens. And your body’s immune system is
totally fine with your particular antigens, but if it detects antigens from someone else’s cells -- including viruses or
bacteria – then it’ll send out antibodies to bind to those markers, often to tag them for destruction by the immune
system. Your red blood cells have specialized antigens on them, called agglutinogens, that activate antibodies that work
by binding invading cells to each other, which causes coagulation, or the clumping of blood. Which agglutinogens you
have on your erythrocytes defines your blood type.

But they’re classified in two different ways. In the most important blood classification -- the kind people are most familiar
with -- there are only two kind of agglutinogens, simply A and B. And your blood can either have one, or both, or neither
of these molecules. So the name of your blood type refers to what kind you have or don’t have: A-type has A antigens, B-
type has B, AB has both, and O has neither. So, why do you need to know what type you are before you give or receive
blood? Well, like I mentioned: If you have either of these antigens, your body will be fine with it, because it doesn’t
produce any antibodies that label it for attack. So if you don’t have a particular antigen on your blood cells -- say the type
B – then you do have antibodies that are going to label those B antigens for attack, should they enter your space. So AB-
type folks are called universal recipients, because they have both antigens, and therefore no antibodies for either. So they
can accept A, or B, or AB, or O blood.

Meanwhile, O-type doesn’t have A or B antigens, so those folks have antibodies for both. That means that they
can only accept other O blood. And yet that lack of antigens means that Type O blood can mix with other types of blood
without getting attacked, which is why it’s known as the universal donor. But just to complicate things a little bit more,
you’ve got a whole other set of antigens with totally different protocol. These are your Rhesus, or Rh antigens, named
after the species of monkey they were first identified in. Much like A and B, you either have the Rh antigens, in which case
you’re Rh positive, or your don’t, and are Rh negative.

Most of the population is Rh positive, so they don’t have the anti-Rh antibodies, which means they can accept either
positive or negative blood. But negative types should stick to just the Rh negative blood. And since the presence of A-B
antigens is controlled by different genes than the Rh ones, we end up with eight different blood types -- four separate
groups, each with two variations. And now, hopefully, you understand why it’s so hard to replace blood, and why True
Blood is...not true. I’ve not actually ever seen that show.

Blood part2

I feel like I haven’t spent nearly enough time lately talking to you about all the stupid and dangerous things that you can
do to your own body, so let’s talk about doping. You probably have heard of this thanks to Lance Armstrong, who secretly
messed with his own blood so that he could illicitly win the Tour de France seven times in a row. You might be dimly
familiar with the fact that doping isn’t like shooting steroids, but it is still cheating, even though, like, why is it cheating?
And how does it work? And is it even possible to make your blood better at being blood? In other words, how can some
people treat -- or mistreat -- their own blood like it’s some sort of drug? Short answer: Because your blood is incredibly
powerful stuff. And its power rests largely in your erythrocytes, or red blood cells. They’re the most abundant cell type in
your blood, accounting for nearly 45 percent of its volume. Every time you take a breath, they pick up oxygen in your
lungs and distribute it through your body, and then grab carbon dioxide, and bring it back to the lungs where it can be
exhaled.

The main mission of erythrocytes is to keep your body fed with oxygen, so your muscles can do their thing, and your brain
can continue to think and feel and boss around your various parts. But you don’t want to mess around with your red
blood cells because erythrocytes are weird characters. They go places that other cells won’t. They purge themselves of
their most precious inner belongings, preferring instead to live as hollow shells. Because of the crushing demands of their
job, they don’t live very long. And just like with your blood pressure, too much of these good things can turn bad quickly.
So the erythrocyte must be respected. It is not for doping. Or for dopes. Despite their prominent role in some
international sports scandals, your red blood cells are fairly simple and unassuming little cells. They’ve got a distinct
biconcave shape -- which just means that they’re concave on both sides -- making them look kinda like a breath mint… a
tiny, bloody breath mint. And while they have a plasma membrane, they don’t have a nucleus and don’t have most of the
parts other cells do. So they’re basically just glorified, protein-filled phospholipid-bilayer sacks. But they’re still another
great example of that harmony between form and function.

For one thing, that biconcave shape gives them a large surface area that’s ideal for gas exchange. It also makes them
flexible, able to change shape as they squeeze through tiny capillaries with diameters smaller than the cell itself. Of
course, all that squeezing and twisting is hard on a cell’s membrane, and that, combined with their general lack of
organelles to help repair the membrane, means these cells don’t live very long, surviving on average only 120 days. But
they sure work hard while they’re alive. And their work is mostly in gathering and transporting oxygen. They’re able to do
this because, if you don’t count their water content, red blood cells are 97 percent hemoglobin -- a molecule that easily
binds to, and releases, oxygen. It’s like an oxygen sponge.

Every hemoglobin molecule is really made of eight different component molecules – four are a red pigment called heme,
and four are a protein called, you guessed it, globin. Each globin is a globular polypeptide chain -- hence its name -- and
proteins, you’ll probably remember, like to bind to stuff. So each globin has its own personal ring-shaped heme molecule,
and in the center of that heme is an iron atom, kinda like a cherry on top of a protein-and-pigment sundae. It’s that iron in
the center of the heme that makes our blood red. Incidentally, not all animals have red blood, because not all animals use
hemoglobin to move oxygen. For example, most mollusks like squids and snails have blue blood, because it contains
hemocyanin, a copper-rich protein-pigment that turns blue when exposed to oxygen. But, iron is what we’re stuck with,
and I have to say it’s great at its job, because each iron can bind with one whole oxygen molecule. And that oxygen really
adds up.

Since you have four iron atoms in every molecule of hemoglobin and every red blood cell contains something like 250
million hemoglobin molecules that means each one of your tiny, floppy red-breath-mints can grab about a BILLION
molecules of oxygen. Exactly how they transfer oxygen and carbon dioxide from your tissue cells is something that we’ll
get into when we talk about the respiratory system. But if you’re wondering why all this hemoglobin can’t simply skip the
red blood cell rigamarole and just run around naked in your blood, it’s because free-range hemoglobin would actually
thicken the blood, making it so viscous that it would impede blood flow. This also happens to play a part in blood doping,
which -- stick with me -- I will explain in a bit. So what does the brief but glorious four-month life of an erythrocyte look
like? Well, remember when I said a red blood cell doesn’t have a nucleus? And maybe you thought, hey wait a second,
how can it even be a cell without a nucleus, or DNA?
First of all, nice catch. But actually, erythrocytes do start off with a nucleus and DNA, they just get rid of them, because
their entire purpose for existing is to schlep around hemoglobin and oxygen, and they want the extra room. The whole
process of forming blood cells, called hematopoiesis, happens in your red bone marrow, which is mostly made of reticular
connective tissue that’s snuggled up to special capillaries called blood sinusoids.

In short, the process begins with a hemocytoblast -- or a specialized stem cell -- which soon differentiates into an early
erythroblast. Then, it starts making a whole bunch of ribosomes, the organelles that manufacture proteins. And in this
case, the ribosomes start cooking up tons of hemoglobin, as the cell transforms into a late-stage erythroblast. When it’s
got enough hemoglobin, it suddenly jettisons most of its organelles, which causes the cell walls to collapse a little, giving it
its biconcave bloody breath-mint shape. Now you’re left with a reticulocyte, which is pretty much just an early erythrocyte
that still has a little group of ribosomes left, called a reticulum. So far, this whole journey so far has taken about fifteen
days. When the reticulocyte is finally bursting with hemoglobin, then it leaves the marrow and enters the bloodstream,
and a couple of days later, when the last ribosomes have degraded, you’ve officially got yourself a mature red blood cell.
And that cell travels around your body, doing its job for a few months before it gets old or damaged and needs to be
replaced. Now, maintaining the balance between production and destruction of these cells is crucial. Too many will make
the blood too viscous and difficult to pump, and too few leads to oxygen deprivation, or hypoxia.

The process of maintaining the right levels of red blood cells is regulated by a special hormone called erythropoietin, or
EPO. It’s produced mostly in the kidneys, but also in the liver, and is constantly circulating in the blood. If you’re anemic,
or hiking at a high altitude, or hemorrhaging blood, or experiencing anything else that creates a drop in your blood oxygen
levels, certain cells in your kidney will notice, and take action. And they can do that, because they traffick in a signaling
molecule called hypoxia-inducible factor, which monitors your blood’s levels of oxygen. The way that this works is pretty
cool. These special kidney cells need oxygen in order to break down that signaling molecule, so if oxygen levels in the
blood are low, they can’t turn the signal off. This means that the signal keeps going, which triggers the release of more
and more EPO, which stimulates your red bone marrow to pump out more red blood cells to carry around more oxygen.

As oxygen levels in your blood increase, the signal is degraded, and EPO production slows. And EPO is a key player in
blood doping too, but we’re gonna have to wait a minute before we get there, because I first want to get back to the fate
of your hard-working erythrocytes. So if you’re generating about two million new blood cells every second, you also have
to dispose of about the same number of dead ones to maintain the balance, right? When these cells get old, they turn
rigid, and their hemoglobins starts to fall apart. As they get stiffer, they can end up getting stuck in capillaries in your brain
or heart, which would not be good.
Luckily, you have certain channels to corral these dying cells, especially around the spleen, which some anatomists call
“the red blood cell graveyard.” So these tired old cells get trapped and then basically ambushed by big macrophage white
blood cells in the spleen, liver, and bone marrow, which break them down and recycle their various components. The
globin proteins are broken down into their basic components -- amino acids -- which go back into the blood to be used by
other cells for making more proteins. Iron from the heme group is separated and either bound to proteins and stored in
the liver, or put right back into a new hemoglobin molecule. And the heme gets turned into bilirubin, a yellowish pigment
that goes to the liver where it’s added to the bile that it secretes into the intestine and eventually leaves the body in your
poop.

Now that you know how your erythrocytes function naturally, it’s easier to see how they can be messed with -- often with
bad results. You can dope your blood in a few different ways, but the most common technique is to inject natural or
synthetic EPO hormone to boost your red blood cell production. It’s also possible to draw and store some of your own
blood, and then transfuse it back into your body after your body has recovered from the blood loss, effectively raising
your volume of red blood cells.

The logic, if you can call it that, is that more red blood cells equals more oxygen being carried to your muscles, and
therefore better physical performance. Now, the extra oxygen can’t change your actual muscle strength, but the added
aerobic capacity does reduce muscle fatigue and enhance endurance, by allowing your muscles to work harder for longer.
And it can provide enough of an extra edge to win a race, like, say, the Tour de France. Seven times. But not only is it
banned in athletic competitions, blood doping is also dangerous. Because, remember, a red blood cell count that’s too
high thickens the blood, and that actually makes it harder for the heart to pump blood around the body. In addition
to defeating the purpose of enhancing the blood’s effectiveness, this can lead to blood clots, and strokes, and heart
failure. So, no thank you. Plus, cheating sucks.

Respiratory System part 1

Let me introduce you to one of the bravest pioneers in the history of life on planet Earth. An organism that blazed the trail
for every single vertebrate that lives on land today -- and many that don’t. It’s one of your most important ancestors.
Meet…well, it doesn’t have a name. And we don’t know exactly what it looked like, either. But we do know that about 380
million years ago, this fishy-looking thing with big, fleshy fins achieved one of the animal kingdom’s greatest milestones:
breathing air. Sounds simple enough, but believe me it wasn’t. Because, for billions of years before this fishy ancestor
came around, basically all of life evolved in water. From the very beginning, the earliest, simplest forms of life -- like
bacteria – extracted oxygen they needed right from the water, through their membranes. And they did it through simple
diffusion -- when a material automatically flows from where it is concentrated, to where it is less concentrated, so it
balances out.

Diffusion works really well, and it requires zero effort, but it wasn’t gonna cut it in the big leagues. Anything larger than a
small worm is simply too big and needs too much oxygen for diffusion to work. So in order to get bigger, early life forms
needed a circulatory system that could move bulk amounts of oxygen around faster inside their bodies, and a respiratory
system to bring more oxygen in contact with their wet membranes. So their respiratory surfaces moved from their outer
surfaces to the insides of their bodies. First, there were gills. But gills, of course, still only work inside of water. And a little
over 380 million years ago, this was starting to lose some of its charm. Earth was getting warmer, the seas were getting
shallower, and much of the planet’s surface water had lower concentrations of oxygen than it used to. Finally, a humble
little lobe-finned fish got fed up, swam up to the water’s surface, and started breathing air. It could do this because it had
evolved a fancy new interface to move gases between the air and its cell membranes. I’m talkin’ about lungs. Wet lungs.
With an efficient new way to take in nearly limitless amounts of oxygen from air, animals were eventually able to get
bigger and more diverse over the ages, and now all of us lung-having vertebrates share that common ancestor.

For lots of animals, including humans, those lungs come with a bunch of other equipment, like protective ribs, a stiff
trachea, and in mammals a strong diaphragm. And together, they form your respiratory system. Which happens to be
best friends and business partners with your circulatory system. It’s only by working together and using both the bulk flow
and simple diffusion of oxygen that they can make possible the process of cellular respiration. In other words: life itself.

So, a lot of improvements have been made to it over the eons, but the respiratory system that you are using right now is
your inheritance from that ancient, ambitious fish – leader of one of the most important anatomical revolutions of the
past half-billion years. Pretend for a minute that you can’t breathe. Like, you just don’t have lungs anymore. You are some
bizarre evolutionary oddity -- a huge, human-shaped organism that doesn’t have a respiratory system. Instead, you get all
of your oxygen the way that your oldest, smallest evolutionary ancestors did -- by simple diffusion. Or at least, you try to
get your oxygen that way. How would it work? Well, poorly.
And that’s partly because one of the keys to efficient diffusion of any material is distance. If you want a molecule to
diffuse across a space quickly, you want it to be as close to its destination as possible, with the fewest obstacles in the
way. But, for a single molecule of oxygen to diffuse from the air through, say, your scalp and then go to a neuron deep
inside your brain, it would have to move through your skin, and then your skull, and then your connective tissue and all
sorts of things. It would eventually get there, like maybe a month later, but at that point, the cell that needed the oxygen
in the first place would have, you know, suffocated to death. Basically, obtaining oxygen through diffusion alone is like
wanting to go to a party at your friend’s place across town, and then walking 20 miles to get there. You could do it, but it
would take forever, and by the time you arrived, you’d be all haggard and the party would be over. So, diffusion alone
isn’t enough to get the job done. We do use it, but only when a whole bunch of the materials we need are right up against
the tissues that can absorb them. So you know what else we need? Bulk flow. Bulk flow is like public transportation -- it
moves large numbers of molecules, quickly. Rather than walk the whole way across town, you can hop on a bus with a
bunch of other people, and get there in twenty minutes.

Every time you take a deep breath, you’re bringing a hundred quintillion oxygen molecules into your lungs all at once --
they’re on a bulk-flow bus ride. And once those oxygen molecules filter down into the cells in your lungs, they’re suddenly
very close to the blood they’re trying to reach. All they have to do is diffuse across four layers of cell membranes to get
from the lung cell into the blood. It’s like just hopping off the bus, and then walking half a block to your friend’s
apartment. That’s why your respiratory system is the way it is: It’s set up to take full advantage of both bulk flow and
simple diffusion. The bulk flow part of things is handled by some of your system’s biggest and most obvious moving parts.
Starting with your lungs, which basically operate like a pump, or a bellows. They don’t have any contractible muscle tissue,
because they need to be able to expand, so they require outside help in order to move.

Enter the diaphragm -- a big, thin set of muscles that separates your thorax from your abdomen. When your lungs empty,
your diaphragm relaxes and looks kinda like an arc pushing up to squish your lungs. You also have the weight of your rib
cage, pushing on your lungs from the top and sides, and together these forces decrease the volume of your lungs. When
you breathe in, your diaphragm contracts, pulling itself flat, and your external intercostal muscles between your ribs
contract. They lift the ribs up and out, causing the chest cavity to expand. This makes the pressure inside the lungs lower
than the air outside your body, and – since fluids like gases move from areas of high pressure to low pressure -- the lungs
fill up with outside air. Then the diaphragm relaxes again, and the weight of the ribs settles in, and the pressure inside the
lungs becomes higher than the outside air, and the air rushes out. And that, my friends, is breathing 101.

Now, your respiratory system contains a lot of parts besides your lungs -- some prominently displayed on your face,
others hidden deep within your chest. And functionally, all of these organs fall into one of two physiological zones.
The upper parts that funnel the air in, make up what’s known as the conducting zone, and it starts with this thing.
Your nose is supported by bone and cartilage, and the bristly hairs and mucus inside it that help filter out dust and other
particles. But it, along with your sinuses, performs another important function: It warms and moistens incoming air, so it
doesn’t dry out those sensitive lung cells that must remain wet. Remember, moisture is key. We evolved from organisms
that lived in water. So, just like with our aquatic bacterial ancestors, we need water for oxygen to dissolve into, before
it can diffuse across the phospholipid bilayer membrane of our cells. Now, if you’ve ever choked on a poorly timed sip of
water, you’ve noticed that you breathe through the same tube that you also move foods and liquids through. This is yet
another leftover from those first fish lungs, which evolved as a branch off the esophagus. Looking back, it was not ideal.
But we are stuck with it. So, the stuff that you swallow soon encounters the epiglottis -- a little trap door of tissue
-- which covers the larynx, and directs bites of sandwich and sips of cola toward your esophagus and keeps them out of
your lungs. And you’ll notice that the esophagus, which heads to your stomach, is nice and flexible, while your trachea, or
windpipe, is rigid and has prominent rings.
That’s because your trachea is basically built like a vacuum hose -- since the lungs create negative pressure with every
breath, the trachea needs those rings to keep it open. If it were soft and floppy, it would collapse every time the pressure
dropped, and you wouldn’t be able to breathe. From there, the trachea splits in two, forming the right and left main
bronchi. You can imagine these inner lung parts as sort of an upside-down tree.

Now we are in the lung tissue, and have entered what we call the respiratory zone. This is where the actual gas exchange
occurs, and everything you find here has a form to suit that function. So the smaller branches of the upside-down tree are
bronchioles, which taper down into progressively narrower tubes, until they empty into the alveolar ducts and then dead
end into tiny alveolar sacs, where the bulk of the gas exchange finally occurs. Because that’s because each sac contains a
cluster of alveoli, these tiny cavities lined with super thin, wet membranes made of simple squamous epithelium tissue.
It’s here that oxygen molecules dissolve in the wet mucous, diffuse across the epithelial cells, and then cross the single
layer of endothelial cells lining the capillaries to enter the bloodstream. And of course it’s also where carbon dioxide
diffuses out of the blood, and then follows the same route back up to the nose and mouth, where it’s exhaled. So it’s your
alveoli where diffusion meets bulk flow. Because while you’re picking up oxygen and dispensing with CO2 one molecule at
a time, you're doing it in enormous quantities at any given second. Both of your lungs contain about 700 million alveoli,
which together provide an amazing 75 square meters of moist membrane surface area. So, the principles that make
respiration possible are relatively simple -- diffusion and bulk flow. And so are the mechanisms in your body that use
them. It just took us about 400 million years to figure out how to make it all work.
Respiratory System part2

Picture this: You’re getting ready to give a big presentation in front of, like, a lot of important people. You’re practicing in
front of your mirror, and then just for a second you forget how to speak. Suddenly, you feel that familiar sting of anxiety,
like an icy hand on the back of your neck. You look at yourself in that mirror and you start imagining some of the worst,
worst-case scenarios. Like, what if you totally lose your train of thought up there? What if you barf? What if everybody
gets up and leaves? Now you’re really nervous. I’m getting freaked out just talking about it. So you start taking quick,
shallow breaths, and you’re feeling light-headed, and seeing stars, and now you, my friend, are hyperventilating. When
we talk about respiration, we tend to focus on oxygen -- and who could blame us?

It’s easy to forget the equally important role that carbon dioxide plays in maintaining homeostasis. Your internal balance
between oxygen and carbon dioxide factors heavily into all sorts of stuff -- especially in your blood, where it can affect
your blood’s pressure, its pH level, even its temperature. And now -- at, like, T-minus 5 minutes to your presentation -- all
of those things are out of whack, because you’re exhaling more CO2 than you should. You’re just about to faint, when a
friend suddenly hands you a paper bag to breathe into. And you’ve never been so grateful for a lunch bag in your life,
because, somehow, it does the trick. Within seconds, you’re back to normal. The drop in CO2 that occurs in your blood
when you hyperventilate is called hypocapnia, and it signals a breakdown in one of the most complex and important
functions that your respiratory system performs. That is: the exchange of gases inside your blood cells, where the stuff
your body doesn’t want is swapped out for what it desperately needs.

This exchange -- between carbon dioxide and oxygen -- is regulated by a whole series of biological signals that your blood
cells use to communicate with your tissues, about what they have, what they want, and what they need to get rid of. It’s
almost like a code, one that’s written into your blood’s chemistry, in the folding of its proteins -- even in its temperature
and acidity. It’s what allows you to perform strenuous physical tasks, like climbing a mountain. It’s also what lets you
reboot your whole respiratory system, with nothing more than a paper bag. I’ll admit it: when we’ve talked about the
chemistry of your blood so far, we’ve tended to keep things pretty simple. Like, hemoglobin contains four protein chains,
each of which contains an iron atom; since iron binds readily with oxygen, that’s how hemoglobin transports oxygen
around your body. But the fact is, hemoglobin’s affinity for oxygen isn’t always the same. In some places, we want our
hemoglobin to have a high affinity for oxygen, so it can easily grab it out of the air. And in others, we want it to have a low
affinity for oxygen oxygen, so it can dump those molecules to feed our cells.

So how does your hemoglobin know when to collect its precious cargo and when to let it go? Well, a lot of it has to do
with a principle of chemistry known as partial pressure. One of the things that fluids always do is move from areas of high
pressure to low pressure. And molecules also diffuse from areas of high concentration to areas of low concentration. But
when we talk about gases in a mixture, we need to combine the ideas of pressure and concentration. See, air is a mixture
of molecules. And when you’re studying the respiratory system, you often need to focus on the oxygen, which makes up
about 21% of it. But that doesn’t tell us how many oxygen molecules there are. For that, we need to know the overall air
pressure, since more molecules in a certain volume means more pressure. So, partial pressure gives us a way of
understanding how much oxygen there is, based on the pressure that it’s creating.

Example: The pressure of air at sea level is about 760 millimeters of mercury. But since only about 21 percent of that air is
oxygen, oxygen’s part of that pressure -- or partial pressure of oxygen -- is 21% of 760, or about 160 millimeters of
mercury. Now, that’s just outside, at sea level. When that air mixes with the air deep in your lungs -- including a lot of air
that you haven’t exhaled yet -- the partial pressure of oxygen drops to about 104 millimeters of mercury.
And in the blood that’s entering your lungs -- after most of its oxygen has been stripped away by your hungry muscles and
neurons -- the oxygen partial pressure is only about 40 millimeters. This big differences in pressure make it easy for
oxygen molecules to travel from the outside air into your blood plasma, because, as a rule dissolved gases always diffuse
down their partial pressure gradients. This is why it’s so much harder to breathe at higher altitudes. When you climb a
mountain, the concentration of oxygen stays at about 21%. But the pressure gets lower, which means the partial pressure
of oxygen also decreases to about 45 millimeters of mercury at the top of Mt. Everest. So the partial pressure of oxygen at
the top of the highest peak in the world, is almost the same as the de-oxygenated blood that’s entering your lungs. So
basically there is no partial pressure gradient, which makes it really hard to get oxygen into your blood. But, let’s get back
to the red blood cells.

Remember that the globin in your hemoglobin is a protein -- and when proteins bind to stuff, they tend to change shape.
And that shape-change can make the protein more or less likely to bind to other stuff. When an empty hemoglobin runs
into an oxygen molecule, things are a little awkward. It’s like a first date -- bonding isn’t so easy. But once they finally bind,
hemoglobin suddenly changes shape, which makes it easier for other oxygen molecules to attach, like friends gathering
around the lunch table. That affinity for joining in -- or cooperativity, as it’s known -- continues until all four binding sites
are taken, and the molecule is fully saturated.

Now your hemoglobin is known as oxyhemoglobin, or HbO2. It is not...not why the cable network is called that. That’s the
“Home Box Office.” Anyway. By the time the blood leaves the lungs, each hemoglobin is fully saturated, the oxygen
partial pressure in your plasma is about 100 millimeters, and now it is ready to be delivered to where it is needed most.
Active tissues, like the brain, heart, and muscles, are always hungry for oxygen. They burn through it quickly, lowering the
oxygen partial pressure around them to about 40 millimeters. So when the blood arrives on the scene, oxygen moves
down the gradient from the plasma to the tissues, to feed those hungry cells. That makes the oxygen partial pressure in
your plasma drop, so your hemoglobin starts to give up more of its oxygen to the plasma. BUT! Partial pressures are only
part of what’s prodding your hemoglobin to give up the goods. All of that metabolic activity going on in your tissues is also
producing other triggers, in the form of waste products -- specifically heat and CO2. Both of those things activate the
release of more oxygen, by lowering hemoglobin’s affinity for it.

Say you’re climbing that mountain again, and your thighs are feeling the burn. Red blood cells saturated with oxygen are
going to the muscle tissue in your quads, where the hemoglobin can dump a bunch of O2, because of the lower partial
pressures of oxygen in your muscles. But a hard-working quad will also heat up the surrounding tissues, and that rise in
temperature changes the shape of hemoglobin -- and it does it in such a way that lowers its affinity for O2. So when those
red blood cells hit that warm active tissue, they release even more oxygen -- like 20 percent more -- beyond what partial
pressures would trigger.

Carbon dioxide triggers the release of oxygen, too, because it also binds to the hemoglobin, changing its shape again,
lowering its affinity for oxygen still more. And as oxygen jumps ship, the hemoglobin can pick up more CO2. Finally, JUST
IN CASE the hemoglobin isn’t getting the message at this point, there’s one more trigger that your respiratory system has
up its sleeve. The spike in CO2 that’s released by your active muscle tissues actually makes your blood more acidic. Since
your blood is mostly water, when CO2 dissolves in it, it forms carbonic acid, which breaks down into bicarbonate and
hydrogen ions. Those ions bind to the hemoglobin, changing its shape yet again, further lowering its affinity for oxygen. So
now, at last, your tissues have the oxygen they need, and your red blood cells are stuck with all this CO2 that they need to
get rid of.

Your red blood cells ride the vein-train back to the lungs, where they encounter a new wave of freshly inhaled oxygen.
And when that O2 binds to the hemoglobin -- which, again, is hard at first -- it eventually changes its shape back to the
way it was when we started, which decreases its affinity for CO2. So the hemoglobin drops its carbon dioxide, which
moves down its partial pressure gradient into the air of your lungs, so you can exhale it, and the whole thing can start all
over again. Now if that isn’t enough to make you hyperventilate, I’m not sure what is. But this brings us back to that
unfortunate episode you had before your big presentation. This whole complex code of chemical signals that I just
described? Well, it assumes that what your cells and tissues are telling each other is actually true. But as we all know,
sometimes our bodies don’t mean what they say. Thanks, body. Like, when you’re freaking out about your presentation,
your sympathetic nervous system makes your heart race and your breathing increase, to prepare you to fight or flee. The
problem is: there’s nothing to actually fight or flee from, so your muscles aren’t actually doing anything, so they’re not
using all the extra oxygen you’re breathing in. And they also aren't producing the extra CO2 that you're suddenly exhaling
all over the place.

So when you start to exhale CO2 faster than your cells release it, its concentration in your blood drops. And with less
carbonic acid around, your blood’s pH starts to rise. And you know what else? While low blood pH does things like change
the shape of your hemoglobin to deliver oxygen, high pH causes vasoconstriction. Normally, this is supposed to divert
blood from the parts you’re not using during times of stress, like your digestive organs, to the parts that you are using.
But when you hyperventilate, this constriction happens everywhere, which means less blood is delivered to your brain,
which makes you light-headed.

Luckily, that trick with the breathing into the paper bag -- it really does work. It works because it lets you breathe back in
all of the CO2 you just breathed out. So the partial pressure of carbon dioxide in the bag is higher, which forces that CO2
into your blood, which lowers its pH, and you get back to homeostasis. And of course, homeostasis is the key to life...and
you know, also to a successful presentation. If you were able to remain calm today, you learned how your blood cells
exchange oxygen and CO2 to maintain homeostasis. We talked about partial pressure gradients, and how they, along with
changes in blood temperature, acidity, and CO2 concentrations, change how hemoglobin binds to gases in your blood.
And you learned how the thing with the bag works.

Digestive System part 1

We all have our reasons for eating nachos at 3 in the afternoon. I happen to have my own. And don’t ask -- it’s personal.
But more generally, we all eat any kind of food to accomplish two simple things: to obtain the energy we need to stay
alive and to get the raw materials required for building all of our tissues and stuff. That’s because, when it comes down to
it, both you and the food you eat contain those two same things: Both you and food are made of “stuff” -- by which I
mean, matter, made of certain kinds of atoms -- and both you and food have energy stored in the bonds between those
atoms. So all living things need to take in stuff and energy, and convert it into slightly different stuff and energy. And you
can get some of the things you need pretty easily. Like, in order to get oxygen for respiration, to unleash the chemical
energy in your food, you just have to inhale. But you can’t just breathe in the stuff you need to build DNA, or actin, or a
phospholipid bilayer. So, how does your body really acquire “stuff”?
That’s where the nachos come in. This cheesy, crunchy dish is made of all different kinds of biological matter -- like
carbohydrates and fat and protein -- and it contains a certain, probably shocking, amount of calories, which is how we
measure energy stored in the chemical bonds in food. So if I take, like, a 100-calorie bite of nachos -- which probably with
this much cheese wouldn’t even be a very big bite -- I can convert the chemical energy stored in those carbohydrates and
proteins and fats to feed my muscle and heart cells and maybe, like, walk a mile -- an activity that happens to use about
100 calories. But I can’t just swallow the nachos and watch the lump of them travel straight to my heart or leg muscles.
In order to actually use this food, I have to convert the biological matter into something my body can work with on the
cellular level, which as you know, is pretty darn tiny. And, the work of converting the stuff in food, into the stuff that’s in
my body, is done by my digestive system.

Human digestion occurs in six main steps -- some of which you are intimately familiar with. Others less so. But every step
of the way, your body is working to reduce all the different kinds of molecules in food into their tiniest and most basic
forms. The first step? Is, uh, probably everybody’s favorite. When it comes to what your digestive system ultimately does,
just think of it as a sort of disassembly line. You could have an order of nachos with The Works -- I’m talking beef and
onions and sour cream and slices of jalapeño -- and your digestive system will deconstruct it, both mechanically and
chemically, one step at a time. It’s gotta do this because your cells work best with materials that are in their most basic
form. Your digestive system reduces food to that level in two main ways: by physically smashing it to smithereens, and by
bathing them, as much as it can, in enzymes. Enzymes are proteins that living things use as catalysts, to speed up chemical
reactions.

When used in digestion, enzymes break down the large molecules in your food into the building blocks that your cells can
actually absorb. Those large molecules are called biological molecules -- also known as macromolecules -- and everything
that you eat, I hope, is at least partially made of them. And there are four main kinds: you got the lipids, the
carbohydrates, the proteins, and the nucleic acids. Each possesses its own density of chemical potential energy, or caloric
value, like for example, 1 gram of carbohydrate contains about 4 calories, while a gram of fat contains about 9 calories.
But many of these biological molecules are polymers -- or sequences of smaller molecules -- and your cells aren’t really
equipped to take them up whole. What your body trafficks in are those polymers’ individual components – called
monomers – and there are four main kinds of those, too: fatty acids, sugars, amino acids, and nucleotides.

The simple idea behind the whole digestive system is to break down the polymers of macromolecules in your food, into
the smaller monomers that your cells can use to build their own polymers, while also getting the energy they need. And,
what your body needs to build at any given moment is always changing. Maybe you need new fat stores so you can have
energy to run a marathon, or new actin and myosin to build bigger muscles, or more DNA so you can replace the skin cells
you scraped off your knee when you fell, or more enzymes so you can digest more food to get more building materials.

To meet your body’s constant, and constantly shifting demands, your digestive system requires a lot of organs that
perform a lot of specific tasks to break down and absorb the right nutrient at the right time. Now, I’m quite sure that
you’re familiar with the key players here -- they’re the hollow organs that form the continuous tube that is your
alimentary canal, aka the gastrointestinal tract, which runs from your mouth to your anus. It’s worth pointing out that
these organs are hollow, because you are basically hollow, too. Your digestive tract is really just one unbroken, insulated
tunnel of outside that just happens to run through your body, and is open at both ends. You’re a donut. So the layer of
stratified squamous and columnar epithelial cells that line your tract is actually a barrier between the outside world and
your inside world -- but it’s a barrier that allows for the selective movement of materials between them. It’s these hollow
organs that do the actual moving, digesting, and absorbing of food, and they include your mouth, pharynx, esophagus,
stomach, and small and large intestines.

In your mouth, in your esophagus, and at the other end of things, at your anus, you have stratified squamous epithelial
tissue, just like your epidermis, to help resist the abrasive action of like, chewing, like corn chips, maybe. From your
stomach on down, though, the inner GI tract is lined with simple columnar epithelial cells, which secrete all sorts of stuff,
and which absorb and process various nutrients. Most of those columnar cells secrete mucus, which lubricates everything,
and protects your cells from being digested by your own digestive enzymes. So, the innermost epithelial layer of the tube
is known as the mucosal layer, and it contains some connective tissue as well, which supplies it with blood. Surrounding
the mucosal layer is the submucosal layer, made of loose areolar connective tissue, which helps provide the elasticity that
the tube needs when you eat a whole pizza in one sitting, and it contains more blood vessels.

And outside that, you have the muscularis externa layer, which as you might guess, is where you find the muscles
responsible for moving food through your tube. Beyond these layers, the GI tract gets tons of support from the accessory
digestive organs, like your teeth, and your tongue, your gallbladder, salivary glands, liver, and pancreas. They’re kind of
like a pit crew, and they mostly help by secreting various enzymes that help take apart food as it comes down the tube.
Together, these two groups on the digestive disassembly line work in six steps to destroy your food and release and
recycle its nutrients. First, of course, you’ve got to introduce the food to your digestive system. What you know as eating,
or ingestion, is basically just creating a bulk flow of nutrients from the outside world into your tissues.
This is where the work of disassembly begins: In your face-hole, which scientists call your mouth. Now, we’re going to get
to the details of what happens here another time, but remember that food disassembly is both mechanical and chemical:
So your teeth pulverize the bite of nacho or whatever, while your salivary glands begin that food’s hours-long enzyme
bath. But the food, at this point, is not nearly “micro” enough to be of any use to your cells, so you have to move that
mush further down your tube. This stage is called propulsion, and its initial mechanism is swallowing -- which, as you
know, is a voluntary action -- but then it’s quickly turned over to the involuntary process of peristalsis. In peristalsis, the
smooth muscles of the walls of your digestive organs take turns contracting and relaxing to squeeze food through the
lumen, or cavity, of your alimentary tract. Waves of peristalsis continue through the esophagus, stomach, and intestines,
and they’re so strong that even if you were hanging upside down while eating your lunch and drinking your tea, the food
would still soldier on, fighting gravity, and eventually make it to its final destination. Don’t do that, though. There’s other
reasons why you shouldn’t be upside down.

Anyway, all of this shipping and handling mechanically breaks down the food even more, and even after it goes through
the stomach and its gastric acid, the mechanical work still continues once it reaches your small intestine, as more smooth
muscle segments push the food back and forth to keep crumbling it up. The goal of all this pulverization is to increase the
surface area of that bite of food by breaking it down into increasingly tiny pieces, to prepare it to encounter more
enzymes in step four: chemical digestion. Really, the actual process of digestion only occurs when the main action
becomes more chemical than mechanical. And here, the accessory digestive organs -- namely, the liver, pancreas and
gallbladder – secrete enzymes into the alimentary canal, where they ambush the mush and break it down into its
most basic chemical building blocks.

Like I said before, our cells prefer to do business in the really basic currency of monomers, like amino acids, fatty acids,
and simple sugars. And digestion allows for the absorption of those nutrients as they pass from the small intestine into
the blood, by both active and passive transport. Once those nutrients are absorbed by your cells, you can finally use the
energy inside of them or use them to build new tissues. The absorption of the nutrients is the goal of the entire process.
But, of course, it is not the end of it. Once your body has sucked out all the nutrients it wants, indigestible substances like
fiber are escorted out of your body. Yeah, I’m talking about pooping, or defecation. And that is the end of the digestive
line -- unless you are a capybara, or one of the other animals who make sure that they get the most out of their lunch, by
giving the whole process another round and practicing coprophagia, aka eating their own poop.

Now, you should notice here that some of the processes of digestion occur in just one place, and are the job of a single
organ -- like hopefully you’re only ingesting through your mouth and eliminating from the large intestine. But most of
these six steps require cooperation among multiple organs. For example, both mechanical and chemical digestion start in
the mouth, and continue through the stomach and small intestines. And some chemical breakdown continues in the
large intestine, thanks to our little bacterial farm there. Over the next couple of weeks we’re going to take you and your
nachos on a stroll through your digestive system and see who’s doing what, where, how, and why. But for now, I’ve got
some nachos to finish, so I gotta go. And eating those nachos, as you learned today, will provide me with energy and raw
materials, by first ingesting something nutritious, propelling it through my alimentary canal where it will be mechanically
broken down, and chemically digested by enzymes until my cells can absorb their monomers and use them to make
whatever they need. And eventually, there will be pooping.

Digestive System part 2

In the summer of 1822, a French-Canadian fur-trapper named Alexis St. Martin was going about his business near Lake
Michigan, when he was shot by a hunter, right in the stomach. The wound was severe, and everyone expected St. Martin
to die that night. But…he didn’t. A local army doctor named William Beaumont kept him alive. In fact, Beaumont
performed so many surgeries on the injury over the next several months, that he decided, somewhat questionably, to just
keep St. Martin’s stomach wound open. St. Martin was left with a hole, or fistula, in his abdominal wall, which allowed
anyone to see right into his stomach. Now, it’s probably hard to work as a fur trapper with a hole in your guts, but
Beaumont saw -- or possibly created -- an opportunity. He hired St. Martin -- technically as a handyman, but really as a
guinea pig. Over several years and some 238 experiments, Beaumont recorded what St. Martin ate, and what his stomach
did to his meals. Sometimes they just skipped the eating part all together and just shoved some food, tied to a string, right
into the guy’s gut-hole. Beaumont took samples of gastric juices and had them analyzed by chemists – something no one
else had done before -- and he also noticed that St. Martin’s digestion slowed at certain times, like when he was sick or
stressed. I mean, like, beyond the stress of having a gaping hole in your abdomen. Through his somewhat questionable
research, Beaumont discovered some major secrets of the digestive system, like that the stomach’s extremely strong
acids and muscular contractions break down food, and that some foods are more digestible or less digestible than others,
and that the brain can affect the stomach. Beaumont’s findings -- as well as his methods of clinical observation --
revolutionized the field of physiology. And St. Martin? Don’t worry about him. He lived to be 83 years old, in great health.
And a hole in his guts... Now, I sincerely hope that you can’t actually see what’s going on in your stomach, but lemme tell
you, the story there is epic.
In your digestive system’s mission to disassemble food into its tiniest, most basic molecular forms, the stretch that runs
from your mouth to your stomach unleashes all of the mechanical and chemical powers at its disposal. It physically roughs
up food; douses it in protein-loving, acid-triggered enzymes; reduces it all into a creamy paste -- and as a bonus, because
it likes you, it also kills a whole host of harmful invaders that, for whatever reason, found their way through your face and
into your tube. But your stomach’s not the end of the line for your food. Unless…it is. I mean, most of the time, everything
from your mouth to your stomach prepares food to be absorbed by your tissues. But sometimes...food finds its way back
up. Yeah, in case the story of Alexis St. Martin didn’t make you wanna do this already, now I’m talking directly about
vomiting.

Let’s begin with the beginning: with your mouth, aka your oral, or buccal, cavity. Now we don’t usually think of it this way,
but that is where digestion starts – the mechanical and chemical breakdown of food through chewing and enzyme-action.
The inside of your mouth is lined with a tough, thick layer of stratified squamous epithelium that can stand up to lots of
friction, like getting scraped by tortilla chips and, like, grilled cheese sandwiches that maybe were cooked a little too much
on the top. Your anterior hard palate and the flexible posterior soft palate form the roof of your mouth. The hard palate
provides, like, a hard surface for the tongue to mash food against, while the soft palate forms a movable fold of flesh that
reflexively closes off the nasopharynx when you swallow, so food gets directed down your esophagus and not up into your
nasal cavity. We all know what teeth are for, and you have roughly 32 of them in your basic types that help you masticate,
or chew your food. The tongue lives on the floor of your mouth, and is basically just a big muscle that grips and constantly
repositions your food as you chew.

The resulting ball of mush actually has its own special name -- it’s a bolus – and the tongue rolls it back to the pharynx, in
preparation for swallowing. But that’s just the physical action that goes on in your mouth. Just as much destruction is
taking place through chemistry. The bolus is broken down with the help of three major pairs of salivary glands that churn
out an average of 1.5 liters of slightly acidic saliva every day. More than four soda cans worth of spit. Per day. And all that
saliva delivers enzymes like salivary amylase, a digestive enzyme that breaks down starches into glucose monomers.

Now, once the food enters the pharynx, it’s propelled by peristalsis into the esophagus, which, except for the little
sphincter at the end that keeps food moving in the right direction, is really just a glorified laundry chute lined with smooth
muscle. The only time you probably even remember that you have an esophagus is when something’s stuck in there, or if
you’re feeling intense heartburn, or if you just puked. But, moving on. Assuming you have not puked yet, then the bolus
moves on to Dr. Beaumont’s ticket to fame: The stomach. The stomach is the stretchiest part of your digestive tube,
capable of holding 2 to 4 liters of material at any given time. TWO TO FOUR LITERS! That’s a lot of nachos. Mixed with spit.
But of course it’s much more than just a storage tank -- it’s lined with the same four main layers found through most of
the GI tract -- the mucosa, submucosa, muscularis externa, and serosa -- but it’s got a few special modifications.

For one thing, the muscularis includes an additional layer of smooth muscle that gives it extra strength, allowing the
stomach not just to hold materials, but to actively smush them around. And the inner mucosa is made up almost entirely
of mucous cells, which produce a protective coat that keeps the stomach tissues from getting digested along with your
lunch. This inner lining is dotted with millions of tiny, deep gastric pits which lead down to tubular gastric glands. These
glands, in turn, contain various types of secretory cells that brew up some of the most potent chemicals in your body. For
example, your stomach has parietal cells that release hydrochloric acid -- a substance more acidic than battery acid --
which lays waste to most of the bacteria, viruses, and other stuff that could make you sick. It also helps denature, or
change the shape of, proteins to make it easier for enzymes to digest them.
And maybe more importantly, when the hydrochloric acid is combined with pepsinogen, an inactive enzyme that’s
secreted by another kind of stomach cell called chief cells, the mixture creates the protein-digesting enzyme pepsin.
Together, this super-powered acid and protein-hungry enzyme can annihilate nearly anything they encounter. This was
apparently something that Beaumont observed first-hand, by dropping hunks of meat into a cup filled with St. Martin’s
personal gastric fluids. He watched the gobbits of food dissolve over time, which is partly how he discovered the
stomach’s role in digestion was as much chemical as mechanical. But with so much mind-blowingly powerful stuff at your
stomach’s disposal, somebody down there has to be in charge -- so your gastric glands also contain enteroendocrine cells.
These cells release regulatory hormones, like serotonin and histamine, which act locally to trigger other cells, to, say,
release more acid, or contract muscle tissue. And when the time comes to tamp the action down, they secrete other
hormones like somatostatin, to inhibit secretions. And then there are G-cells, which produce the most important
hormone for stimulating gastric activity: gastrin. Most signals that increase stomach activity get the job done by increasing
the secretion of gastrin, which then stimulates the release of other gastric fluids, as well as stomach-muscle activity. Now,
if the smell of baking cookies has ever made your mouth water and your belly grumble, then it might not surprise you to
learn that these stomach secretions are ruled by neural mechanisms as well hormonal ones. In fact, stomach regulation
occurs in three phases, based on where the food is sensed -- the brain, the stomach, and the small intestine.

The cephalic phase is the one ruled by your brain, and it kicks in when you first see, smell, taste, or even think about food.
That sensory input gets relayed to the hypothalamus, which stimulates the medulla oblongata, which then taps the
parasympathetic fibers in the vagus nerve. From there, the signals are sent to the stomach with the word that, “Hey, we
think that maybe cookies are on the way, so you might want to prepare yourself.” Now this is a conditioned reflex, so it
only works if you want to eat the food in question. If I happen to be super-full, or not feeling well, or somebody puts a pile
of squid eyeballs in front of me, the cephalic phase isn’t gonna happen. And no offense if squid eyeballs are totally your
thing, they’re just not my thing. But say I eat the plate of squid eyeballs anyway because, you know, I’m trying to be polite.
Well, even without the cephalic warm up, when that food hits my stomach, local mechanisms, both neural and hormonal
jump start the gastric phase. For the next few hours, as my stomach grows distended from the food, it activates stretch
receptors that again stimulate my medulla and get my vagus nerves to tell my stomach to turn up the juice. At the same
time, the secretion of gastrin is activated by other signals, like the rise in alkalinity caused by the stomach acid getting
neutralized as it does its job.

Conversely, as stomach acidity increases, it inhibits the release of gastrin. Now, the third phase of gastric regulation -- the
intestinal phase -- speeds or slows the rate in which your stomach empties, so that the small intestine doesn’t get too
overloaded with too much acid -- or with the creamy paste that your stomach turns your food into, known as chyme.
Now remember, not a lot of absorption actually occurs in the stomach. The stomach is more like a decontamination tank.
Sure, it pummels your food down to a paste, but it’s also where your body tries to obliterate any nasties that could make
you sick. As long as food is still in there, your body has a chance to kind of size it up, and feel it out, and it reserves the
right to eject anything that it feels is potentially dangerous. Lots of factors can trigger the stomach’s urge to purge, or
vomit, but the most common are simply ingesting too much food, or eating some kind of irritant or toxin, like those
produced by bad bacteria, too much alcohol, certain drugs, or unappealing foods.

Of course if you’ve ever puked in a moment of trauma or stress, you know how emotions and anxiety can also trigger your
stomach to launch its lunch. That’s the brain influencing the cephalic phase of gastric regulation again, by sending extra
fight or flight signals to the stomach. Beaumont noticed this mind-stomach connection whenever St. Martin’s digestion
was affected by illness or stress -- something you’d think he’d have felt every time that doctor came at him with some
meat on a string. If you were able to keep down your lunch down today, you learned how mechanical and chemical
digestion start in the mouth and continue in the stomach, where food is pummeled by acids and enzymes and turned into
chyme. We also looked at the stomach’s cephalic, gastric, and intestinal phases of digestive regulation.

Digestive System part 3

You know, we’ve been talking about a lot of serious stuff here lately. Heart failure. Respiratory gas exchange. People with
holes in their stomachs. Nachos. Some might say I’ve even been flaunting my ability to eat, digest, and enjoy a plate
of chips and melted cheese. And I wouldn't blame them if they did, because sadly, nachos aren’t for everyone. In fact, I
can safely say nachos are really only a good idea for about a third of humans. For the rest, what may start as a party in
your mouth will surely end in gastric distress. Such is the fate of the lactose intolerant. Lactose is basically milk-sugar that
can only be digested with the help of a special intestinal enzyme -- lactase -- which many adults do not produce enough
of. In fact, way back in the day, none of us did, until about 7500 years ago, when a particularly handy genetic mutation
popped up in central Europe. This so-called lactase persistence trait probably spread as Neolithic groups trekked north
and west through Europe. Today nearly 90 percent of adult Britons and Scandinavians can chug all the milk they want,
whereas down toward the Mediterranean, probably less than 40 percent have lactase persistence, and fewer than ten
percent in Africa and Asia.

Now technically, a lactose intolerant person can still consume dairy at their own risk, but since their own bodies can’t
break down lactose, the job is left to the three-pound bacteria farm living in their large intestines -- bacteria that try their
hardest to make something of those milk sugars, the results of which are gas, and bloating, and diarrhea. So, it turns out,
nachos aren’t just a good way to talk about how the digestive system works, they’re also a good way to talk about when it
doesn’t. Remember how the stomach is great at obliterating matter, but not so hot when it comes to actually chemically
digesting stuff, or really absorbing much of anything? You might say the stomach lacks subtlety. But luckily, it’s got friends
in low places, and the small intestine is more than happy to pick up the slack and provide a cozy environment where your
food is at long last disassembled and absorbed by your cells.

Now there’s a lot of mechanical action and peristalsis going on here, but there’s also a ton of chemical digesting too. And
while homebrewed intestinal juices help digest the chyme that your stomach turns food into, the real power actually
comes from the outside helpers -- the liver, gallbladder, and pancreas. Now, the small intestine is called “small” not
because it’s short but because it’s about half the diameter of the large intestine -- the thing is actually like 6 or 7 meters
long. Not only that, but whole deal is lined with epithelial tissue that has more folds than an origami octopus. These folds
are lined with tiny hair-like villi and even tinier microvilli, which create a truly impressive surface area -- large enough that,
if it were unfolded, it would cover a tennis court. It’s this massive surface area, and the countless capillaries just beneath
it, that make the small intestine such a champion absorber of nutrients.

It shares the same four tissue layers seen throughout the GI tract, and has three main subdivisions: Straight outta the
stomach and snuggled around the pancreas, you’ve got the relatively short and mostly immovable duodenum, which is
where most of the chemical digestion occurs. The middle section is the jejunum, where most of the absorption takes
place. And finally at the end, running into the large intestine is the ileum, where important vitamins like A, B12, E, D, and K
are absorbed. But the duodenum is what you might call the business end of the small intestine. It receives chyme and
gastric juices from the stomach through the pyloric sphincter, but it also imports bile from the liver and gallbladder,
enzymes from the pancreas, and creates its own homegrown mix of enzymes. Some of the imported enzymes eventually
pass through your system on the wave of gooey chyme. But other enzymes are actually bound to cell membranes in the
intestinal mucosal layer, and they’re reusable.

Enzymes are proteins, and proteins are expensive. So these compounds -- known as brush border enzymes -- can just sit
around and process food as it passes by, without your body having to make new ones. And the lactase that so many of us
don’t have, is one of these. Now, the duodenum communicates with the stomach in the last phase of gastric regulation
that we talked about in the last episode -- the intestinal phase. This is where the duodenum lets the stomach know, with
hormones and nerve signals, when and how much chyme to release so it doesn’t get overwhelmed all at once. It’s also
where stuff like bicarbonate from the pancreas gets dumped, to help neutralize the stomach acid before it burns a hole in
your guts. And this brings me to your crucial accessory organs -- the things apart from the alimentary canal that never
come in contact with ingested material, but still play an essential role in digestion.

First up: the liver. The liver is a massive, fatty, four-lobed, and very important organ. It lives directly under your diaphragm
and -- fun fact -- it can actually fully regenerate itself after an injury or surgery, with as little 25 percent of its original
tissue. The liver serves tons of critical metabolic and regulatory roles that we don’t have time to get into right now, but its
main role in the digestive system is to make bile. Bile is the missing ingredient your body needs to attack fatty foods,
which is a tricky business. In part, that’s because fat isn’t water soluble, and since your insides are mostly water, fats will
clump together, becoming hard to digest. To keep fat from clumping, you need an emulsifier, so bile comes in to keep big,
hydrophobic fat molecules from sticking together, which allows lipid-hungry enzymes to move in and break them down
into fatty acids and monoglycerides that you can then digest and absorb. But while your liver creates the bile, it gets
stored and concentrated in the neighboring gallbladder, the thin, green sac cozied up to the liver. It gets the signal when
chyme slides into the duodenum, which activates the enteroendocrine cells to release a pair of hormones. Those
hormones in turn tell the gallbladder to contract and squirt bile through the cystic and bile ducts into the duodenum.
Another crucial accessory organ is the pancreas, a gland that looks like a fistful of cottage cheese stuffed in a plastic bag.

The pancreas also does lots of important things for your body, especially related to your endocrine system, but for our
purposes today, just know that it brews up a powerful enzyme cocktail that is also triggered by those same two
hormones. Pancreatic juice is kinda like the Neapolitan Ice Cream of bodily secretions -- it’s like everybody’s favorite
ingredients all put together, and when you mix them, the result is especially powerful. You’ve got trypsin and peptidase in
there, which break proteins down into amino acids, and you have lipases that turn triglycerides into fatty acids and
glycerol. Amylase, meanwhile, reduces carbs to glucose and fructose, and nuclease busts the nucleic acids that are in DNA
and RNA into nucleotides. Once all of those macromolecules have been dissembled into their monomers, the small
intestine’s epithelial cells can finally absorb and transport them through your capillaries and into the bloodstream, where
they can travel to pretty much to any cell in your body, and be used to build collagen, or store fat, or replace dying cells.
The true purpose behind all the eating that you do. So! Once the chyme has worked through your small intestine, it passes
through the ileocecal valve and hits the cecum, the first part of the large intestine, where, congratulations, your food is
now officially feces! The large intestine -- consisting of the colon, rectum, and anus -- is relatively short, at about one and
half meters, and it provides a nice little frame for the small intestine, here at the end of the alimentary canal.

By now, your body has sucked up almost all of the nutrients it can, and is basically just pushing indigestible goo around, so
the large intestine doesn’t have a lot of hard work to do. Its main functions are to absorb any remaining water so you
don’t have constant diarrhea, and to store the rest until it’s ready to exit the body. It also plays host to hundreds of
species and trillions of individual gut bacteria, which digest whatever chyme your body couldn’t, releasing essential B and
K vitamins, and some short fatty acids, which the large intestine can still absorb. In doing so, they also produce gases like
carbon dioxide and methane, sulfurous compounds called mercaptans, and hydrogen sulfide, which eventually...pass. OK,
I know what you’re thinking now, you’re like, “Hank, what’s up with the nachos?” “Surely you’re not just gonna to bring
up nachos at the beginning of the episode without explaining how they can turn on you?!” Well, I’ve never disappointed
you before, have I?

Those of us who can’t produce the enzyme lactase in our small intestine simply let milk and cheese pass through the
organ untouched, leaving the digestion to these bacteria in the large intestine. And those bacteria possess about 1000
different kinds of enzymes of their own, including lactase. But their digestion process produces a whole lotta extra gas,
which is why nachos may leave me feeling cheesy and satisfied, but leave you bloated, and crampy, and malodorous. But
enough farting around, let’s wrap up this fantastic journey. Fecal matter keeps moving through in a couple of different
ways. Slow, segmenting haustral contractions keep mixing and chopping it in the large intestine, occurring every 30
minutes or so and lasting about a minute. But most people also experience a few mass peristalsis movements a day -- big,
intense contractions that clear out a large swath of intestine at once, pushing feces into the rectum. These often occur
just after eating. Once in the rectum, your poop stimulates stretch receptors that tap the parasympathetic defecation
reflex, which signals the colon and rectum to contract, and the internal anal sphincter to relax.

This forces the poop into the anal canal, sending more messages to the brain that allow us to decide whether to
voluntarily open the external anal sphincter, or just hold it for a minute while we find a bathroom. And when that
moment arrives, what was once food says farewell to the alimentary canal that temporarily held it, and passes back into
the light of day. And that, my friends, is the end of your digestive system. Pretty cool, right? And so it’s all over, but that
doesn’t mean you should forget about what we learned today, which is that the small intestine performs most of your
chemical digestion in the duodenum, while accessory organs including the liver, gallbladder, and pancreas contribute
enzymes that all but finish the job. Then your large intestine, which is actually shorter than the small intestine, tries to
extract the last bit of nutrition, including the occasional attempt to turn nachos into energy, which for most humans, ends
in gassy failure.

MITOSIS

For any creature bigger than a single-celled organism, all of life stems from cells' ability to reproduce themselves, because
that's what allows organisms to develop, grow, heal and keep from dying, for as long as possible. This particular kind of
cell division is called mitosis, and it's responsible for a whole lot of your body's key functions. If you get a cut, your body
needs to make new cells. Mitosis. Have too much to drink and damage your liver? You gotta replace those cells. Mitosis.
Tumor growing in your spine? Unfortunately, again mitosis. While you go from a seven pound baby to a seventy pound
child it's not your cells that are increasing in mass; you're just getting more of them. Over and over and over again. That's
mitosis.

This process is so central to your life that it will take place in your body, over your lifetime, about 10 quadrillion times.
That's 10 thousand billion times! Like all split ups, it's not easy. It's going to maybe be a little bit messy, there's a lot of
drama, and it can take a surprisingly long amount of time. But trust me, after we're done with it we'll all be better off. So
you are made of trillions of cells just like giraffes and redwood trees. And remember that inside each cell there's a nucleus
that stores your DNA, which contains all of the instructions on how to build you. That DNA is organized into chromosomes
and as we've mentioned before, in your body cells, or somatic cells, you have 46 chromosomes grouped into 23 pairs, one
in each pair is from your mom, and the other one's from your dad. Cells with all 46 chromosomes are called diploid cells,
because they have 2 sets each. And that's what we're focusing on today. You also have haploid cells that have half as
many chromosomes (23). And those are your sex cells. They're produced in an equally fantastic process called meiosis,
which we'll be talking about in the next episode.

But for now, the main thing to remember about mitosis is that it allows one cell with 46 chromosomes to split into two
cells that are genetically identical, each with 46 chromosomes. All in order to keep the party of life going. Now, the
nucleus in your cell controls everything that goes on in the cell. It has all of the instructions necessary for making the cell
survive so you don't need to duplicate the whole cell. All you need to do is duplicate the DNA, get it wrapped up,
and then if you have two separate pockets of DNA, that's all you need to have two new cells.

Mitosis takes place in a series of discrete stages called prophase, metaphase, anaphase, and telophase. And you can just
say that over and over again, and let it sink into your head. And part of what's really amazing about this whole process is
that, while we know what these stages are, we don't always know the underlying mechanisms that make all of them
happen. And this is part of science. Science isn't all the stuff we know, it's how we're trying to figure all this stuff out.
Consider it job security if you want to be a biologist; there is a lot of stuff that future biologists have to still figure out,
and this is one of them. Alright, let's get our clone on. So, most of their lives, cells hang out in this limbo period called
interphase, which means they're in between episodes of mitosis, mostly growing and working and doing all the stuff
that makes them useful to us. During interphase, the long strings of DNA are loosely coiled and messy, like that dust
bunny of dog fur and laundry lint under your bed.

That mess of DNA is called chromatin. But as the mitosis process begins to gear up, lots of things start happening in the
cell to get ready for the big division. One of the more important things that happens is that this little set of protein
cylinders next to the nucleus, called the centrosome, duplicates itself. duplicates itself. We're going to have to move a lot
of stuff around in the nucleus and that's going to be regulated by these centrosomes. The other thing that happens is all
of the DNA begins to replicate itself too, giving the cell two copies of every strand of DNA. To brush up on how DNA
replicates itself like this, check out this episode and then come on back. Now the cell enters the first phase, or the
prophase, when that mess of chromatin condenses and coils up on itself to produce thick strands of DNA wrapped around
proteins - those my friends, are your chromosomes.

Instead of dust bunnies, the DNA is starting to look a little bit more like dreadlocks. And the duplicates that have been
made don't just float around freely; they stay attached to the original, and together they look like little X's; these are
called the chromatids and one copy is the left leg and arm of the X, and the other copy is the right leg and arm. Where
they meet in the middle is the centromere. Just so you know, these X's are also called chromosomes sometimes double
chromosomes, or double-stranded chromosomes. And when the chromatids separate, they're considered individual
chromosomes too. Now, while the chromosomes are forming, the nuclear envelope gets out of the way by completely
disintegrating. And the centrosomes then peel away from the nucleus, and start heading to opposite ends of the cell. As
they go, they leave behind a wide trail of protein ropes called microtubules running from one centrosome to the other.

You might recall from our anatomy of the animal cell that microtubules help provide a kind of structure to the cell;
and this is exactly what they're doing here. Now we reach the metaphase, which literally means "after phase" and it's the
longest phase of mitosis.; It can take up to 20 minutes. During the metaphase, the chromosomes attach to those ropey
microtubules right in the middle, at their centromeres. The chromosomes then begin to be moved around, and this seems
to be being done by molecules called motor proteins. And while we don't know too much about how these motors work,
we do know, for instance, that there are two of them on each side of the centromere. These are called Centromere-
associated protein E.

So, these motors proteins attach to the microtubule ropes and basically serve to spool up the tubules' slack. At the same
time, another protein, dynein, is pulling up the slack from the other ends of the ropes near the cell membranes. After
being pulled in this direction and that, the chromosomes line up, right down
the middle of the cell. And that brings us to the latest installment of Bio-lography. So how do those chromosomes line up
like that? We know that there are motor proteins involved but like, how? What are they doing? Well, remember when I
said earlier that there are a lot of things that we don't totally understand
about mitosis? I

It's sort of weird that we don't, because we can literally watch mitosis happening under microscopes, but chromosome
alignment is a good example of a small detail that has only very recently been figured out, and it was a revelation
about 130 years in the making.

Mitosis was first observed by a German biologist by the name of Walther Flemming, who in 1878 was studying the tissue
of salamander gills and fins when he saw cells' nuclei split in two and migrate away from each other to form two new
cells. He called this process mitosis, after the Greek word for thread, because of the messy jumble of chromatin, a term he
also coined, that he saw in the nuclei. But Flemming didn't pick up on the implications of this discovery for genetics, which
was still a young discipline. And over the next century, generations of scientists started piecing together the mitosis
puzzle, by determining the role of microtubules, say, or identifying motor proteins. Now, the most recent contribution to
this research was made by a postdoctoral student named Tomomi Kiyomitsu at MIT. He watched the same process that
Flemming watched, and figured out how at least one of the motor proteins helps snap the chromosomes into line. He was
studying a motor protein called dynein, which sits on the inside of the membrane. Think of the microtubles as tug-of-war
ropes, with the chromosomes as the flag in the middle. What Kiyomitsu discovered was that dynein plays tug of war with
itself. Dynein grabs onto one end of the microtubules and pulls the tubules and chromosomes toward one end of the cell.
When the ends of microtubules come too close to the cell membrane, they release a chemical signal that punts the
dynein to the other side of the cell. There, it grabs onto the other end of the microtubles and starts pulling, until SMACK it
gets punted back again. All of this ensures that the chromosomes will line up exactly in the middle, so that they will be
split evenly.

That discovery was published in February 2012, a couple of weeks before I sat in this chair, and 134 years after mitosis
was first observed. If you want to join the ranks of scientists who are answering the many questions left about mitosis,
and lots of other things about our lives maybe someday I'll do a Bio-lography about you.

Now so far we've gone through the interphase, when the centrosomes and DNA replicate themselves and get ready for
the split; the prophase, when the chromosomes form and the centrosomes start to spread apart; and metaphase, where
the chromosomes align in the middle of the cell. And now it's time to separate the chromosomes from their copies. This
time, motor proteins start pulling so hard on the ropes that the X-shaped chromosomes split back into their individual,
single chromosomes. Once they're detached from each other, they're dragged toward either end of the cell. The prefix
'ana' means 'back' that may help you remember the name of this phase, called anaphase.

After this, it's just a matter of using all of that genetic material to rebuild, so that the copied genetic material has all
the accouterments of home. In the last phase, telophase, each of the new cell's structures are reconstructed. First, the
nuclear membrane re-forms, and nucleoli form within them. And the chromosomes relax back into chromatin. Then a
little crease forms between the two new cells, which marks the beginning of the final split. That division between the two
new cells is called cleavage. All that's left is to make a clean break. This is done by cytokinesis literally "cell movement"
by which the two new nuclei move apart from each other, and the cells separate. We now have two new cells, each with
the full set of 46 chromosomes. These clones are called the daughter cells of the original cell, and like identical twins they
are genetic copies of each other and also of their parent. But, that's obviously not the case for you. Even if you are an
identical twin. While you kind of are a clone of your sibling you are not a clone of your parents. Instead, half of your DNA
in each of your cells is from your mom, and half is from your dad. To understand why that is, we have to understand how
eggs and sperm are formed. And that is meiosis.
MEIOSIS

The kind of reproduction that we're most familiar with of course is sexual reproduction, where sperm meets egg, they
share genetic information, and then that fertilized egg splits in half, and then those halves split in half, and so on and so
on and so on, to make a living thing with trillions of cells that all do specialized things. And if you're not suitably impressed
by the fact that we all come from one single cell and then we become THIS: then I don't- just- I don't know how to impress
you! But riddle me this my friends: If sexual reproduction begins with sex cells, the sperm and the egg, where do the
sperm and egg come from? So how do sex cells form so that they each have only half of the genetic information that the
resulting offspring will end up with? And for that matter, why aren't all of our sex cells the same? Like, why are my
brother John and I are different? Sure we both wear glasses and we both kind of look like a tall Dr. Who, but, you know,
we have different color hair and different noses and I'm way better at Assassin's Creed than he is. So why aren't we
identical? As far as we know, we both came from the same two people with the same two sets of DNA, right?
The answer to these questions, and a lot of other of life's mysteries, is meiosis.

In the last episode we talked about how most of your cells your body, or somatic, cells clone themselves through the
process of mitosis. Mitosis replicates a cell with a complete set of 46 chromosomes into two daughter cells that are each
identical to each other. But of course, even though the vast majority of your cells can clone themselves, you cannot clone
yourself, and for good reason. Actually, reasons. If mitosis were the only kind of cell division we were capable of, that
would mean:
A) you would be a clone of one of your parents, which would be, awkward to say the least, or possibly
B) half of your cells would be clones from your mom, and half would be clones from your dad. And you would look REALLY
weird.

But that's not how we roll; we do things a better way, where all of your body cells contain the same mix of DNA 46
chromosomes grouped into 23 pairs, one in each pair from your mom, and one from your dad. Those pairs of
chromosomes are pretty similar, but they're not identical. They contain versions of the same genes, or alleles, in the same
spot for any given trait. Since they're so similar, we call the pairs homologous chromosome pairs. Homologous is a word
that comes up a lot in genetics. It just means that two things have the same (homo) relation (logos) even if they are a little
bit different. However, there are some very special cells that you have that have only one half of that amount, 23
chromosomes. Those are sperm and egg cells. These are haploid cells they have half of a full set of chromosomes. And
they need each other to combine to make the complete 46. Creating those kinds of cells requires a process that's very
similar to mitosis but with a totally different outcome: meiosis. That's when a specialized diploid cell splits in half twice,
producing four separate cells, each of which is genetically distinct from the others.

Meiosis is a lot like mitosis, except, twice. It goes through the same stages as mitosis: prophase, metaphase, anaphase,
and telophase. But then it goes through another round of those stages again and they have the same names,
conveniently, except with a 'II' after them, they're like sequels. And just as with the Final Destination movies, the sequels
have pretty much the same plot, just some new actors. So the raw materials for this process are in your ovaries or your
testes, depending on... you know... you know what it depends on.

They're diploid cells called either primary oocytes or primary spermatocytes, depending on what kind of gamete they
make. Men produce sperm, you may have heard, and they produce it throughout their adult lives, whereas women are
born with a certain amount of eggs that they'll release over many years after puberty. Here you might want to watch the
previous episode about mitosis again because that's where we go into detail about each stage of the process. Once you're
done with that, we can start making some baby-makers! Now, just like with mitosis, there's a spell between rounds of cell
division where the cell is gearing up for the next big split. This is called interphase, when all the key players are replicating
themselves. The long strings of DNA in the nucleus begin to duplicate, leaving two copies of every strand. To jog your
memory about how DNA does this, we did a whole episode on it you can watch it and come on back.

A similar process takes place with the centrosomes, the set of protein cylinders next to the nucleus that will regulate how
all of the materials will be moved around, along these ropey proteins called microtubules. That brings us to the first round
of meiosis, prophase I. This is nearly the same as in mitosis as the centrosomes start heading to their corners of the cell,
unspooling the microtubules, and the DNA clumps up with some proteins into chromosomes. Each single chromosome is
linked to its duplicate copy to make an X-shaped double chromosome. Now keep this in mind: once attached, each single
chromosome is called a chromatid, one on each side of the X. Each double chromosome has 2 chromatids. Here, meiosis
prophase I includes two additional and VERY IMPORTANT steps: crossover and homologous recombination.

Remember that the point here is to end up with four sex cells that each have just one single chromosome from each of
the homologous pairs. But unlike in mitosis, where all the copies end up the same, here every copy is going to be different
from the rest. Each double chromosome lines up next to its homolog, so there's your mother's version lined up right next
to your father's version of the same chromosome. Now If you look, you'll see that these two double chromosomes, each
with two chromatids, add up to 4 chromatids. Now watch: One chromatid from each X gets tangled up with the other X.
That's crossover. And while they're tangled, they trade sections of DNA. That's the recombination. The sections that
they're trading are from the same location on each chromosome, so one is giving up its genetic code for, like, hair color or
body odor, and in return it's getting the other chromosome's genes for that trait.

Now this is important. What just happened here, creating new gene combinations on a single chromosome, is the whole
point of reproducing this way. Life might be a lot less stressful if we could just clone ourselves, but then we'd also clone all
our bad gene combinations, and we wouldn't be able to change and adapt to our environment. Remember that one of the
pillars of natural selection is variation. This is a major source of that variation. What's more, since all of the four
chromatids have swapped some DNA segments at random, that means that all four chromatids are now different. Later
on in the process, each chromatid will end up in a separate sex cell. This is why all eggs produced by the same woman
have a slightly different genetic code; same for sperm in men. And that's why my brother John and I look different, even
though we're both made from the same two sets of DNA. Because of the luck of the genetic draw that happens in
recombination, I got this mane of luscious hair, and John was stuck with his trash, brown puff. And don't forget about my
mad Assassin's Creed skills. But then of course, there is that one pair of chromosomes that doesn't always go through the
crossover or recombination: that's the 23rd pair. And those are your sex chromosomes. If you're female, you have two
matched, beautiful, fully capable chromosomes there, X chromosomes. Since they're the same, they do the whole
crossover and recombination thing. But, if you, like me, are a male, you get one of those X chromosomes and another
from your dad that's kinda ugly and short and runted and doesn't have a lot of genetic information on it. During prophase,
the X wants nothing to do with the little Y, because they're not homologous, so they just don't match up. And because the
X-Y pairs on these chromosomes will split later on into single chromatids, half of the four resulting sperm cells will be X
(leading to female offspring), and half will be Y (leading to male offspring). Now what comes next is another kind of
amazing feat of alignment. This is Metaphase I, and in mitosis you might recall that all of the chromosomes lined up in a
single row, powered by motor proteins, and were then pulled in half. But Not here.

In meiosis, each chromosome lines up next to its homologous pair-partner that it's already swapped a few genes with.
Now the homologous pairs get pulled apart and migrate to either end of the cell. That's Anaphase I. The final phase of the
first round, Telophase I, rolls out in pretty much the same way as mitosis. The nuclear membrane re-forms, and nucleoli
form within them. The chromosomes fray out back into chromatin. Then a crease forms between the two new cells,
called cleavage, and as the two new nuclei move apart from each other, the cells separate in a process called cytokinesis
literally again "cell movement."

That's the end of round 1. We now have two haploid cells, each with 23 double chromosomes that are new, unique
combinations of the original chromosome pairs. In these new cells, the chromosomes are still duplicated and connected
at the centromeres they still look like X's. But remember the aim is to end up with four cells, so it's time for those sequels.
Here the process is exactly the same as mitosis, except that the aim here isn't to duplicate the double chromosomes, but
instead to pull them apart into separate single-strand chromosomes. Because of this, there's no DNA replication involved
in Prophase II; instead the DNA just clumps up again into chromosomes, and the infrastructure for moving them, the
microtubules, are put back in place. In metaphase II, the chromosomes are moved into alignment into the middle of the
cell. And in anaphase II, the chromatids are pulled apart into separate, single chromosomes. The chromosomes uncoil into
chromatin, and the crease-forming cleavage and the final separation of cytokinesis then mark the end of telophase II.
From one original cell with 46 chromosomes, we now have four new cells with 23 single chromosomes each. If these are
sperm, all four resulting cells are the same size, but they each have slightly different genetic information.
And half will be for making girls, and half will be for making boys. But if this is the egg-making process, then it goes a little
bit differently here and the result is only one egg. To rewind a little, during telophase I, more of the inner goodness of a
cell (the cytoplasm, the organelles, heads into one of the cells that gets split off than to the other one. In telophase II,
when it's time to split again the same thing happens, with more stuff going into one of the cells than the other. This big
old fat, remaining cell becomes the egg, with more of the nutrients, cytoplasm, and organelles that it will take to make a
new embryo.

The other 3 cells that were produced, the little ones, are called polar bodies, and they're totally useless in people, though
they are useful in plants. In plants, those polar bodies actually also get fertilized too, and they become the endosperm,
that's the starchy-proteiny stuff that we grind into wheat or pop into popcorn. And it's basically the nutrients that feed
the plant embryo seed. And that's all there is to it. I know, you probably were really excited when I started talking about
reproduction but then I rambled on for a long time about haploid and diploid cells. But now you can say you know more
about the miracle of reproduction. It's not actually a miracle. It's science!
BIO/Biochemistry

Animal Cell

The thing that all of these other things have in common is that they're made out of the same basic building block: the
animal cell. Animals are made up of your run-of-the-mill eukaryotic cells. These are called eukaryotic because they have a
"true kernel," in the Greek. A "good nucleus". And that contains the DNA and calls the shots for the rest of the cell also
containing a bunch of organelles. A bunch of different kinds of organelles and they all have very specific functions. And all
this is surrounded by the cell membrane. Of course, plants have eukaryotic cells too, but theirs are set up a little bit
differently, of course they have organelles that allow them to make their own food which is super nice. We don't have
those. And also their cell membrane is actually a cell wall that's made of cellulose. It's rigid, which is why plants can't
dance. If you want to know all about plant cells, we did a whole video on it and you can click on it here if it's online yet. It
might not be. Though a lot of the stuff in this video is going to apply to all eukaryotic cells, which includes plants, fungi and
protists. Now, rigid cells walls are cool and all, but one of the reasons animals have been so successful is that their
flexible membrane, in addition to allowing them the ability to dance, gives animals the flexibility to create a bunch of
different cell types and organs types and tissue types that could never be possible in a plant. The cell walls that protect
plants and give them structure prevent them from evolving complicated nerve structures and muscle cells, that allow
animals to be such a powerful force for eating plants.

Animals can move around, find shelter and food, find things to mate with all that good stuff. In fact, the ability to move
oneself around using specialized muscle tissue has been 100% trademarked by kingdom Animalia. >>OFF CAMERA: Ah!
What about protozoans? Excellent point! What about protozoans? They don't have specialized muscle tissue. They move
around with cillia and flagella and that kind of thing. So, way back in 1665, British scientist Robert Hooke discovered cells
with his kinda crude, beta version microscope. He called them "cells" because hey looked like bare, spartan monks'
bedrooms with not much going on inside. Hooke was a smart guy and everything, but he could not have been more wrong
about what was going on inside of a cell. There is a whole lot going on inside of a eukaryotic cell. It's more like a city than
a monk's cell. In fact, let's go with that a cell is like a city. It has defined geographical limits, a ruling government, power
plants, roads, waste treatment plants, a police force, industry...all the things a booming metropolis needs to run
smoothly. But this city does not have one of those hippie governments where everybody votes on stuff and talks things
out at town hall meetings and crap like that. Nope. Think fascist Italy circa 1938. Think Kim Jong Il's- I mean, think Kim
Jong-Un's North Korea, and you might be getting a closer idea of how eukaryotic cells do their business. Let's start out
with city limits.

So, as you approach the city of Eukaryopolis there's a chance that you will notice something that a traditional city never
has, which is either cilia or flagella. Some eukaryotic cells have either one or the other of these structures--cilia being a
bunch of little tiny arms that wiggle around and flagella being one long whip-like tail. Some cells have neither. Sperm
cells, for instance, have flagella, and our lungs and throat cells have cilia that push mucus up and out of our lungs. Cilia
and flagella are made of long protein fibers called microtubules, and they both have the same basic structure: 9 pairs of
microtubules forming a ring around 2 central microtubules. This is often called the 9+2 structure. Anyway, just so you
know—when you're approaching city, watch out for the cilia and flagella! If you make it past the cilia, you'll encounter
what's called a cell membrane, which is kind of squishy, not rigid, plant cell wall, which totally encloses the city and all its
contents. It's also in charge of monitoring what comes in and out of the cell--kinda like the fascist border police. The cell
membrane has selective permeability, meaning that it can choose what molecules come in and out of the cells, for the
most part. And I did an entire video on this, which you can check out right here. Now the landscape of Eukaryopolis, it's
important to note, is kind of wet and squishy. It's a bit of a swampland.

Each eukaryotic cell is filled with a solution of water and nutrients called cytoplasm. And inside this cytoplasm is a sort of
scaffolding called the cytoskeleton, it's basically just a bunch of protein strands that reinforce the cell. Centrosomes are a
special part of this reinforcement; they assemble long microtubules out of proteins that act like steel girders that hold all
the city's buildings together. The cytoplasm provides the infrastructure necessary for all the organelles to do all of their
awesome, amazing business, with the notable exception of the nucleus, which has its own special cytoplasm called
"nucleoplasm" which is a more luxurious, premium environment befitting the cell's Beloved Leader. But we'll get to that in
a minute.

First, let's talk about the cell's highway system, the endoplasmic reticulum, or just ER, are organelles that create a network
of membranes that carry stuff around the cell. These membranes are phospholipid bilayers. The same as in the cell
membrane. There are two types of ER: there's the rough and the smooth. They are fairly similar, but slightly different
shapes and slightly different functions. The rough ER looks bumpy because it has ribosomes attached to it, and the
smooth ER doesn't, so it's a smooth network of tubes.

Smooth ER acts as a kind of factory-warehouse in the cell city. It contains enzymes that help with the creation of
important lipids, which you'll recall from our talk about biological molecules -- i.e. phosopholipids and steroids that turn
out to be sex hormones. Other enzymes in the smooth ER specialize in detoxifying substances, like the noxious stuff
derived from drugs and alcohol, which they do by adding a carboxyl group to them, making them soluble in water.

Finally, the smooth ER also stores ions in solutions that the cell may need later on, especially sodium ions, which are used
for energy in muscle cells. So the smooth ER helps make lipids, while the rough ER helps in the synthesis and packaging
of proteins. And the proteins are created by another typer of organelle called the ribosome. Ribosomes can float freely
throughout the cytoplasm or be attached to the nuclear envelope, which is where they're spat out from, and their job is
to assemble amino acids into polypeptides.

As the ribosome builds an amino acid chain, the chain is pushed into the ER. When the protein chain is complete, the ER
pinches it off and sends it to the Golgi apparatus. In the city that is a cell, the Golgi is the post office, processing proteins
and packaging them up before sending them wherever they need to go. Calling it an apparatus makes it sound like a bit of
complicated machinery, which it kind of is, because it's made up of these stacks of membranous layers that are
sometimes called Golgi bodies. The Golgi bodies can cut up large proteins into smaller hormones and can combine
proteins with carbohydrates to make various molecules, like, for instance, snot. The bodies package these little goodies
into sacs called vesicles, which have phosopholipid walls just like the main cell membrane, then ships them out, either to
other parts of the cell or outside the cell wall. We learn more about how vesicles do this in the next episode of Crash
Course.

The Golgi bodies also put the finishing touches on the lysosomes. Lysosomes are basically the waste treatment plants and
recycling centers of the city. These organelles are basically sacks full of enzymes that break down cellular waste and debris
from outside of the cell and turn it into simple compounds, which are transferred into the cytoplasm as new cell-building
materials.

Now, finally, let us talk about the nucleus, the Beloved Leader. The nucleus is a highly specialized organelle that lives in its
own double-membraned, high-security compound with its buddy the nucleolus. And within the cell, the nucleus is in
charge in a major way. Because it stores the cell's DNA, it has all the information the cell needs to do its job. So the
nucleus makes the laws for the city and orders the other organelles around, telling them how and when to grow, what to
metabolize, what proteins to synthesize, how and when to divide. The nucleus does all this by using the information
blueprinted in its DNA to build proteins that will facilitate a specific job getting done. For instance, on January 1st, 2012,
lets say a liver cell needs to help break down an entire bottle of champagne. The nucleus in that liver cell would start
telling the cell to make alcohol dehydrogenase, which is the enzyme that makes alcohol not-alcohol anymore. This protein
synthesis business is complicated, so lucky for you, we will have or may already have an entire video about how it
happens.

The nucleus holds its precious DNA, along with some proteins, in a weblike substance called chromatin. When it comes
time for the cell to split, the chromatin gathers into rod-shaped chromosomes, each of which holds DNA molecules.
Different species of animals have different numbers of chromosomes. We humans have 46. Fruit flies have 8. Hedgehogs,
which are adorable, are less complex than humans and have 90 Now the nucleolus, which lives inside the nucleus, is the
only organelle that's not enveloped by its own membrane--it's just a gooey splotch of stuff within the nucleus. Its main job
is creating ribosomal RNA, or rRNA, which it then combines with some proteins to form the basic units of ribosomes. Once
these units are done, the nucleolus spits them out of the nuclear envelope, where they are fully assembled into
ribosomes.
The nucleus then sends orders in the form of messenger RNA, or mRNA, to those ribosomes, which are the henchmen
that carry out the orders in the rest of the cell. How exactly the ribosomes do this is immensely complex and awesome, so
awesome, in fact, that we're going to give it the full Crash Course treatment in an entire episode. And now for what is,
totally objectively speaking of course, the coolest part of an animal cell: its power plants! The mitochondria are these
smooth, oblong organelles where the amazing and super-important process of respiration takes place. This is where
energy is derived from carbohydrates, fats and other fuels and is converted into adenosine triphosphate or ATP, which is
like the main currency that drives life in Eukaryopolis. You can learn more about ATP and respiration in an episode that we
did on that.

Now of course, some cells, like muscle cells or neuron cells need a lot more power than the average cell in the body, so
those cells have a lot more mitochondria per cell. But maybe the coolest thing about mitochondria is that long ago animal
cells didn't have them, but they existed as their own sort of bacterial cell. One day, one of these things ended up inside of
an animal cell, probably because the animal cell was trying to eat it, but instead of eating it, it realized that this thing was
really super smart and good at turning food into energy and it just kept it. It stayed around. And to this day they sort of
act like their own, separate organisms, like they do their own thing within the cell, they replicate themselves, and they
even contain a small amount of DNA.

What may be even more awesome -- if that's possible -- is that mitochondria are in the egg cell when an egg gets
fertilized, and those mitochondria have DNA. But because mitochondria replicate themselves in a separate fashion, it
doesn't get mixed with the DNA of the father, it's just the mother's mitochondrial DNA. That means that your and my
mitochondrial DNA is exactly the same as the mitochondrial DNA of our mothers. And because this special DNA is isolated
in this way, scientists can actually track back and back and back and back to a single "Mitochondrial Eve" who lived about
200,000 years ago in Africa. All of that complication and mystery and beauty in one of the cells of your body. It's
complicated, yes. But worth understanding.

Plant Cell

Plants are freaking great because they have this magical wizard power that allows them to take carbon dioxide out of the
air and convert it into wonderful, fresh, pure, oxygen for us to breathe. They're also way cooler than us because, unlike us
and every other animal on the planet, they don't require all kinds of Hot Pockets and fancy coffee drinks to keep them
going The only thing plants need to make themselves a delicious feast is sunlight and water. Just sunlight and water! Paula
Deen can't do that and she makes fried-egg bacon donut burgers. I'm telling you this is surprisingly good. This is a
different kind of magic. But you know, part of this is plants! And everything in it, in fact, everything that is in this
McDonalds in fact, everything that you have ever eaten in your life is either made from plants, or from something that ate
plants. So, let's talk about plants!

Plants probably evolved more than 500 million years ago. The earliest land-plant fossils date back more than 400 million
years ago. These plants were lycophytes which are still around today and which reproduce through making a bunch of
spores, shedding them, saying a couple of Hail Marys and hoping for the best. Some of these lycophytes went on to evolve
into "scale trees," which are now extinct, but huge, swampy forests of them used to cover the Earth. Some people call
these scale tree forests "coal forests" because there were so many of them and they were so dense and they covered the
whole Earth and they eventually fossilized into giant seams of coal, which are very important to our lifestyles today. So
this is now called the Carboniferous Period. See what they did there? Because Coal is made out of carbon, so they named
the epoch of geological history over how face-meltingly intense and productive these forests were. I would give my left
eyeball, three fingers on my left hand -- the middle ones, so that I could hang loose -- and my pinky toe if I were able to go
back and see these scale forests because they were freaking awesome.

Anyway, Angiosperms, or plants that use flowers to reproduce, didn't develop until the end of the Cretaceous Period,
about 65 million years ago, just as the dinosaurs were dying out. Which makes you wonder if in fact the first angiosperms
assassinated all the dinosaurs. I'm not saying that's definitely what happened, I'm just saying it's a little bit suspicious.
Anyway, on the cellular level, plant and animal cells are actually pretty similar. They're called eukaryotic cells, which
means they have a "good kernel." And that "kernel" is the nucleus. Not "new-cue-lus." And the nucleus can
be found in all sorts of cells.

Animal cells, plant cells, algae cells. You know, basically all of the popular kids. Eukaryotic cells are way more advanced
than prokaryotic cells. We have the eukaryotic cell and we have the prokaryotic cell. Prokaryotic basically means "before
the kernel." Pro-kernel. And then we have eukaryotic, which means "good kernel!" The prokaryotes include your bacteria
and your archaea, which you've probably met before in your lifetime, every time you've had strep throat, for example, or
if you've ever been in a hot spring or an oil well or something. They're everywhere. They covered the planet. They cover
you! But like I said, eukaryotes have that separately enclosed nucleus. That all important nucleus that contains its DNA
and is enclosed by a separate membrane Because the eukaryotic cell is a busy place -- there's chemical reactions going on
in all different parts of the cell -- it's important to keep those places divided up.

Eukaryotic cells also have these little stuff-doing factories called organelles. I guess we decided we would name everything
something weird... But, organelles. And they're suspended in cytoplasm, continuing with the really esoteric terminology
that you're going to have to know.

Cytoplasm is mostly just water, but it's some other stuff too. Well basically if you want to know about the structure of the
eukaryotic cell you should watch my video on animal cells. Let's just link to it right here. Plant and animal cells are very
similar environments. They control themselves in very similar ways, but obviously, plants and animals are very different
things. What are the differences in a plant cell that makes it so different from an animal? Well that's what we're going to
go over now.

First, plants are thought to have evolved from green algae, which evolved from some more primitive prokaryotes, and
something plants inherited from their ancestors was a rigid wall surrounding the plasma membrane of each cell.
So, this cell wall of plants is mainly made of cellulose and lignin, which are two really tough compounds. Cellulose is by far
the most common and easy to find complex carbohydrate in nature, although if you were to include simple carbohydrates
as well, glucose would win that one. And this is because, fascinating fact: cellulose is just a chain of glucose molecules!

If you want to jog your memory about carbohydrates and other organic molecules, you can watch this episode right here.
Anyway, as it happens, you know who needs carbohydrates to live? Animals. But you know what's a real pain in the ass to
digest? Cellulose. Plants weren't born yesterday. Cellulose is a far more complex structure than you will generally find in a
prokaryotic cell, but it's also one of the main things that differentiates a plant cell from an animal cell.

Animals cells don't have this rigid cell wall--they have just a flexible membrane that frees them up to move around and eat
plants and stuff. However, the cell wall gives structure to a plant's leaves, roots and stems, and it also protects it to a
degree. Which is why trees aren't squishy and don't giggle when you poke them. The combination of lignin and cellulose is
what makes trees, for example, able to grow really, really freaking tall. Both of these compounds are extremely strong and
resistant to deterioration. When we eat food, lignin and cellulose is what we call "roughage" because we can't digest it.
It's still useful for us in certain aspects of our digestive system, but it's not nutritious. Which is why eating a stick is really
unappetizing. And like, your shirt. This is a 100% plant shirt, but it doesn't taste good. We can't go around eating wood like
a beaver or grass like a cow because our digestive systems just aren't set up for that.

However, other animals that don't have access to delicious donut burgers have either developed gigantic stomachs like
sloths or multiple stomachs like goats in order to make a living eating cellulose. These animals have a kind of bacteria in
their stomach that actually does the digestion of the cellulose for it. It breaks the cellulose into individual glucose
molecules, which can then be used for food. But other animals, like humans -- mostly carnivores -- don't have any of that
kind of bacteria, which is why it's so difficult for us to digest sticks. Ah! But there is another reason why cellulose and
lignin are very very useful to us as humans: It burns, my friends!

This is basically what would happen in our stomachs. It's oxidizing. It's producing the energy that we would get out of it if
we were able to, except it's doing it very very quickly. And this is the kind of energy, like, this energy that's coming out of it
right now, is the energy that would be useful to us if we were cows. But we're not. So instead, we just use it to keep
ourselves warm on the cold winter nights.

Anyway, while we animals are walking around, spending our lives searching for ever more digestible plant materials,
plants don't have to do any of that. They just sit there and they make their own food. And you know how they do that?
They do it with photosynthesis! Another thing that plant cells have that animal cells just don't have are plastids, organelles
that plants use to make and store compounds that they need. And you wanna know something super interesting about
plastids? They and their fellow organelles, the mitochondria that generate energy for the cell, actually started as bacteria
that were absorbed into plant cells very early in their evolution like maybe some protist-like cell absorbed a bacteria, and
it found that instead of digesting that bacteria for the energy that it has, it could use that bacteria. That bacteria could
create energy for the cell or convert light into lovely glucose compounds, which is crazy!

Nobody's really, precisely sure how this happened, but they know that it did happen because plastids and mitochondria
both have double membranes. One from the original bacteria, and one from the cell as it wrapped around it. Cool, huh?
Anyway, the most important of the plastids are chloroplasts, which convert light energy from the sun into the sugar and
into oxygen, which the plant doesn't need, so it just gets rid of it. All the green parts of a plant that you see -- the leaves,
the non-woody stems, the unripened oranges -- are all filled with cells which are filled with chloroplasts, which are making
food and oxygen for you.

Another big difference between a plant cell and an animal cell, is the large, central vacuole. Plant cells can push water into
vacuoles which provides turgor pressure from inside the cell, which reinforces the already stiff cellulose wall and makes
the plant rigid like a crunchy piece of celery or something. Usually when soil dries out or a celery stalk sits in your
refrigerator for too long, the cells lose some water, turgor pressure drops, and the plant wilts or gets all floppy.

So, the vacuoles are also kind of a storage container for the cell. It can contain water, which plants need to save up, just in
case. And also other compounds that the cell might need. It can also contain and export stuff the cell doesn't need
anymore, like wastes. Some animal cells also have vacuoles, but they aren't as large and they don't have this very
important job of giving the animal shape. So now, let's do this. Let's just go over the basics of plant cell anatomy:

1. They've got a cell wall that's made out of cellulose and so it's really rigid and not messing around.
2. They've got a nucleus in its own little baggie that's separate from all the other organelles. This is basically the
headquarters of any eukaryotic cell: it stores all the genetic information for the plant and also acts as the cell's activities
director, telling it how to grow, when to split, when to jump and how high...that sort of thing. Animal cells have this kind
of nucleus too, but prokaryotes don't. Which is why they're stuck hanging around in oil wells and stuff.
3. They've got plastids, including chloroplasts, which are awesome green food-making machines.
4. They've got a central vacuole that stores water and other stuff and helps give the cell structural support.

And so, stack these cells on top of one another like apartments in an apartment building and you've got a plant! And all of
these unique features are what make it possible for plants to put food on
our table and air in our lungs. So next time you see a plant, just go ahead and shake its
hand and thank it for its hard work and its service.

Potrebbero piacerti anche