Storm in A Teacup - The Physics of Everyday - Helen Czerski
Storm in A Teacup - The Physics of Everyday - Helen Czerski
Storm in A Teacup - The Physics of Everyday - Helen Czerski
Storm in a Teacup is about linking the little things we see every day with
the big world we live in. It’s a romp through our physical surroundings,
showing how playing with things such as popcorn, coffee stains and
refrigerator magnets can shed light on Scott’s expeditions, medical tests and
solving our future energy needs.
Each chapter begins with something small in the everyday, something that
we will have seen many times but may have never thought about. By the
end of each chapter, we’ll see the same patterns explaining some of the
most important science and technology of our time. Each mini-quest is
rewarding in itself, but the real pay-off comes when the pieces are put
together.
Cover
About the Book
Title Page
Dedication
Epigraph
Introduction
References
Acknowledgements
Index
About the Author
Copyright
STORM IN A TEACUP
The Physics of Everyday Life
HELEN CZERSKI
To my parents,
Jan and Susan
While I was a university student, I spent a while
doing physics revision at my Nana’s house. Nana, a
down-to-earth northerner, was very impressed when
I told her that I was studying the structure of the
atom.
WE LIVE ON the edge, perched on the boundary between planet Earth and the
rest of the universe. On a clear night, anyone can admire the vast legions of
bright stars, familiar and permanent, landmarks unique to our place in the
cosmos. Every human civilization has seen the stars, but no one has touched
them. Our home here on Earth is the opposite: messy, changeable, bursting
with novelty and full of things that we touch and tweak every day. This is
the place to look if you’re interested in what makes the universe tick. The
physical world is full of startling variety, caused by the same principles and
the same atoms combining in different ways to produce a rich bounty of
outcomes. But this diversity isn’t random. Our world is full of patterns.
If you pour milk into your tea and give it a quick stir, you’ll see a swirl, a
spiral of two fluids circling each other while barely touching. In your
teacup, the spiral lasts just a few seconds before the two liquids mix
completely. But it was there for long enough to be seen, a brief reminder
that liquids mix in beautiful swirling patterns and not by merging
instantaneously. The same pattern can be seen in other places too, for the
same reason. If you look down on the Earth from space, you will often see
very similar swirls in the clouds, made where warm air and cold air waltz
around each other instead of mixing directly. In Britain, these swirls come
rolling across the Atlantic from the west on a regular basis, causing our
notoriously changeable weather. They form at the boundary between cold
polar air to the north and warm tropical air to the south. The cool and warm
air chase each other around in circles, and you can see the pattern clearly on
satellite images. We know these swirls as depressions or cyclones, and we
experience rapid changes between wind, rain and sunshine as the arms of
the spiral spin past.
A rotating storm might seem to have very little in common with a stirred
mug of tea, but the similarity in the patterns is more than coincidence. It’s a
clue that hints at something more fundamental. Hidden beneath both is a
systematic basis for all such formations, one discovered and explored and
tested by rigorous experiments carried out by generations of humans. This
process of discovery is science: the continual refinement and testing of our
understanding, alongside the digging that reveals even more to be
understood.
Sometimes a pattern is easy to spot in new places. But sometimes the
connection goes a little bit deeper and so it’s all the more satisfying when it
finally emerges. For example, you might not think that scorpions and
cyclists have much in common. But they both use the same scientific trick
to survive, although in opposite ways.
A moonless night in the North American desert is cold and quiet. Finding
anything out here seems close to impossible, since the ground is lit only by
dim starlight. But to find one particular treasure, you equip yourself with a
special torch and set out into the darkness. The torch needs to be one that
produces light that is invisible to our species: ultraviolet light, or ‘black
light’. As the beam roams across the ground, it’s impossible to tell exactly
where it’s pointing because it’s invisible. Then there’s a flash, and the
darkness of the desert is punctured by a surprised scuttling patch of eerie
bright blue-green. It’s a scorpion.
This is how enthusiasts find scorpions. These black arachnids have
pigments in their exoskeleton that take in ultraviolet light that we can’t see
and give back visible light that we can see. It’s a really clever technique,
although if you’re scared of scorpions to start with, your appreciation might
be a little muted. The name for this trick of the light is fluorescence. The
blue-green scorpion glow is thought to be an adaptation to help the
scorpions find the best hiding places at dusk. Ultraviolet light is around all
the time, but at dusk, when the sun has just slipped below the horizon, most
of the visible light has gone and only the ultraviolet is left. So if the
scorpion is out in the open, it will glow and be easy to spot because there
isn’t much other blue or green light around. If the scorpion is even slightly
exposed, it can detect its own glow and so it knows it needs to do a better
job of hiding. It’s an elegant and effective signalling system – or was until
the humans bearing ultraviolet torches turned up.
Fortunately for the arachnophobes, you don’t need to be in a scorpion-
populated desert at night to see fluorescence – it’s pretty common on a dull
morning in the city as well. Look again at those safety-conscious cyclists:
their high-visibility jackets seem oddly bright compared with the
surroundings. It looks as though they’re glowing, and that’s because they
are. On cloudy days, the clouds block the visible light, but lots of ultraviolet
still gets through. The pigments in the high-visibility jackets are taking in
the ultraviolet and giving back visible light. It’s exactly the same trick the
scorpions are playing, but for the opposite reason. The cyclists want to
glow; if they’re emitting that extra light, they’re easier to see and so safer.
This sort of fluorescence is pretty much a free lunch for humans; we’re not
aware of the ultraviolet light in the first place, so we don’t lose anything
when it gets turned into something we can use.
It’s fascinating that it happens at all, but the real joy for me is that a
nugget of physics like that isn’t just an interesting fact: it’s a tool that you
carry with you. It can be useful anywhere. In this case, the same bit of
physics helps both scorpions and cyclists survive. It also makes tonic water
glow under ultraviolet light, because the quinine in it is fluorescent. And it’s
how laundry brighteners and highlighter pens work their magic. Next time
you look at a highlighted paragraph, bear in mind that the highlighter ink is
also acting as an ultraviolet detector; even though you can’t see the
ultraviolet directly, you know it’s there because of this glow.
I studied physics because it explained things that I was interested in. It
allowed me to look around and see the mechanisms making our everyday
world tick. Best of all, it let me work some of them out for myself. Even
though I’m a professional physicist now, lots of the things I’ve worked out
for myself haven’t involved laboratories or complicated computer software
or expensive experiments. The most satisfying discoveries have come from
random things I was playing with when I wasn’t meant to be doing science
at all. Knowing about some basic bits of physics turns the world into a
toybox.
There is sometimes a bit of snobbery about the science found in kitchens
and gardens and city streets. It’s seen as something to occupy children with,
a trivial distraction which is important for the young, but of no real use to
adults. An adult might buy a book about how the universe works, and that’s
seen as being a proper adult topic. But that attitude misses something very
important: the same physics applies everywhere. A toaster can teach you
about some of the most fundamental laws of physics, and the benefit of a
toaster is that you’ve probably got one, and you can see it working for
yourself. Physics is awesome precisely because the same patterns are
universal: they exist both in the kitchen and in the furthest reaches of the
universe. The advantage of looking at the toaster first is that even if you
never get to worry about the temperature of the universe, you still know
why your toast is hot. But once you’re familiar with the pattern, you will
recognize it in many other places, and some of those other places will be the
most impressive achievements of human society. Learning the science of
the everyday is a direct route to the background knowledge about the world
that every citizen needs in order to participate fully in society.
Have you ever had to tell apart a raw egg and a boiled egg, without
taking their shells off? There’s an easy way to do it. Put the egg down on a
smooth, hard surface and set it spinning. After a few seconds, briefly touch
the outside of the shell with one finger, just enough to stop the egg’s
rotation. The egg might just sit there, stationary. But after a second or two,
it might slowly start to spin again. Raw and boiled eggs look the same on
the outside, but their insides are different and that gives the secret away.
When you touched the cooked egg, you stopped a whole solid object. But
when you stopped the raw egg, you only stopped the shell. The liquid inside
never stopped swirling around, and so after a second or so, the shell started
to rotate again, because it was being dragged around by its insides. If you
don’t believe me, go and find an egg and try it. It is a principle of physics
that objects tend to continue the same sort of movement unless you push or
pull on them. In this case, the total amount of spin of the egg white stays the
same because it had no reason to change. This is known as conservation of
angular momentum. And it doesn’t just work in eggs.
The Hubble Space Telescope, an orbiting eye that has been whooshing
round our planet since 1990, has produced many thousands of spectacular
images of the cosmos. It has sent back pictures of Mars, the rings of
Uranus, the oldest stars in the Milky Way, the wonderfully named Sombrero
Galaxy and the giant Crab Nebula. But when you’re floating freely in
space, how do you hold your position as you gaze on such tiny pinpoints of
light? How do you know precisely which way you’re facing? Hubble has
six gyroscopes, each of which is a wheel spinning at 19,200 revolutions per
second. Conservation of angular momentum means that those wheels will
keep spinning at that rate because there is nothing to slow them down. And
the spin axis will stay pointed in precisely the same direction, because it has
no reason to move. The gyroscopes give Hubble a reference direction, so
that its optics can stay locked on a distant object for as long as necessary.
The physical principle used to orient one of the most advanced technologies
our civilization can produce can be demonstrated with an egg in your
kitchen.
This is why I love physics. Everything you learn will come in useful
somewhere else, and it’s all one big adventure because you don’t know
where it will take you next. As far as we know, the physical laws we
observe here on Earth apply everywhere in the universe. Many of the nuts
and bolts of our universe are accessible to everyone. You can test them for
yourself. What you can learn with an egg hatches into a principle that
applies everywhere. You step outside armed with your hatchling, and the
world looks different.
In the past, information was treasured more than it is now. Each nugget
was hard-earned and valuable. These days, we live on the shore of an ocean
of knowledge, one with regular tsunamis that threaten our sanity. If you can
manage your life as you are, why seek more knowledge and therefore more
complications? The Hubble Space Telescope is all very nice, but unless it’s
also going to look downwards once in a while to find your keys when
you’re late for a meeting, does it make any difference?
Humans are curious about the world, and we get a lot of joy from
satisfying our curiosity. The process is even more rewarding if you work
things out for yourself, or if you share the journey of discovery with others.
And the physical principles you learn from playing also apply to new
medical technologies, the weather, mobile phones, self-cleaning clothes and
fusion reactors. Modern life is full of complex decisions: is it worth paying
more for a compact fluorescent light bulb? Is it safe to sleep with my phone
next to my bed? Should I trust the weather forecast? What difference does it
make if my sunglasses have polarizing lenses? The basic principles alone
often won’t provide specific answers, but they’ll provide the context needed
to ask the right questions. And if we’re used to working things out for
ourselves, we won’t feel helpless when the answer isn’t obvious on the first
try. We’ll know that with a bit of extra thinking, we can clarify things.
Critical thinking is essential to make sense of our world, especially with
advertisers and politicians all telling us loudly that they know best. We need
to be able to look at the evidence and work out whether we agree with
them. And there’s more than our own daily lives at stake. We are
responsible for our civilization. We vote, we choose what to buy and how to
live, and we are collectively part of the human journey. No one can
understand every single detail of our complex world, but the basic
principles are fantastically valuable tools to take with you on the way.
Because of all this, I think that playing with the physical toys in the
world around us is more than ‘just fun’, even though I’m a huge fan of fun
for its own sake. Science isn’t just about collecting facts; it’s a logical
process for working things out. The point of science is that everyone can
look at the data and come to a reasoned conclusion. At first, those
conclusions may differ, but then you go and collect more data that helps
you decide between one description of the world and another, and
eventually the conclusions converge. This is what separates science from
other disciplines – a scientific hypothesis must make specific testable
predictions. That means that if you have an idea about how you think
something works, the next thing to do is to work out what the consequences
of your idea would be. In particular, you have to look hard for consequences
that you can check for, and especially for consequences that you can prove
wrong. If your hypothesis passes every test we can think of, we cautiously
agree that this is probably a good model for the way the world works.
Science is always trying to prove itself wrong, because that’s the quickest
route to finding out what’s actually going on.
You don’t have to be a qualified scientist to experiment with the world.
Knowing some basic physical principles will set you on the right track to
work a lot of things out for yourself. Sometimes, it doesn’t even have to be
an organized process – the jigsaw pieces almost slot themselves into place.
One of my favourite voyages of discovery started with disappointment: I
made blueberry jam and it turned out pink. Bright fuchsia pink. It happened
a few years ago, when I was living in Rhode Island, sorting out the last bits
and pieces before moving back to the UK. Most things were done, but there
was one last project that I was adamant about fitting in before I left. I had
always loved blueberries – they were slightly exotic, delicious, and
beautifully and bizarrely blue. In most places I’ve lived they come in
frustratingly small quantities, but in Rhode Island they grow in abundance. I
wanted to convert some of the summer blueberry bounty into blue jam to
take back to the UK. So I spent one of my last mornings there picking and
sorting blueberries.
The most important and exciting thing about blueberry jam is surely that
it is blue. I thought so, anyway. But nature had other ideas. The pan of
bubbling jam was many things, but blue was not one of them. I filled the
jam jars, and the jam really did taste lovely. But the lingering
disappointment and confusion followed me and my pink jam back to the
UK.
Six months later, I was asked by a friend to help with a historical
conundrum. He was making a TV programme about witches, and he said
that there were records of ‘wise women’ boiling verbena petals in water and
putting the resulting liquid on people’s skin as a way of telling whether they
were bewitched. He wondered whether they were measuring something
systematically, even if it wasn’t what they intended. I did a bit of research
and found that maybe they were.
Purple verbena flowers, along with red cabbage, blood oranges and lots
of other red and purple plants, contain chemical compounds called
anthocyanins. These anthocyanins are pigments, and they give the plants
their bright colours. There are a few different versions, so the colour varies
a bit, but they all have a similar molecular structure. That’s not all, though.
The colour also depends on the acidity of the liquid that the molecule is in –
what’s called its ‘pH value’. If you make that environment a little more
acidic or a little more alkaline, the molecules change shape slightly and so
their colour changes. They are indicators, nature’s version of litmus paper.
You can have lots of fun in the kitchen with this. You need to boil the
plant to get the pigment out, so boil a bit of red cabbage in water, and then
save the water (which is now purple). Mix some with vinegar, and it goes
red. A solution of laundry powder (a strong alkali) makes it go yellow or
green. You can generate a whole rainbow of outcomes just from what’s in
your kitchen. I know: I did it. I love this discovery because these
anthocyanins are everywhere, and accessible to anyone. No chemistry set
required!
So maybe these wise women were using the verbena flowers to test for
pH, not bewitchment. Your skin pH can vary naturally, and putting the
verbena concoction on skin could produce different colours for different
people. I could make cabbage water go from purple to blue when I was nice
and sweaty after a long run, but it didn’t change colour when I hadn’t been
exercising. The wise women may have noticed that different people made
the verbena pigments change in different ways, and put their own
interpretation on it. We’ll never know for sure, but it seems to me to be a
reasonable hypothesis.
So much for history. And then I remembered the blueberries and the jam.
Blueberries are blue because they contain anthocyanins. Jam has only four
ingredients: fruit, sugar, water and lemon juice. The lemon juice helps the
natural pectin from the fruit do its job of making the jam set. It does that
because … it’s acid. My blueberry jam was pink because the boiled
blueberries were acting as a saucepan-sized litmus test. It had to be pink for
the jam to set properly. The excitement of working that out almost made up
for the disappointment of never having made blue jam. Almost. But the
discovery that there’s a whole rainbow of colour to be had from just one
fruit is the sort of treasure that’s worth the sacrifice.
This book is about linking the little things we see every day with the big
world we live in. It’s a romp through the physical world, showing how
playing with things like popcorn, coffee stains and refrigerator magnets can
shed light on Scott’s expeditions, medical tests and solving our future
energy needs. Science is not about ‘them’, it’s about ‘us’, and we can all go
on this adventure in our own way. Each chapter begins with something
small in the everyday world, something that we will have seen many times
but may never have thought about. By the end of each chapter, we’ll see the
same patterns explaining some of the most important science and
technology of our time. Each mini-quest is rewarding in itself, but the real
payoff comes when the pieces are put together.
There’s another benefit to knowing about how the world works, and it’s
one that scientists don’t talk about often enough. Seeing what makes the
world tick changes your perspective. The world is a mosaic of physical
patterns, and once you’re familiar with the basics, you start to see how
those patterns fit together. I hope that as you read this book, the scientific
hatchlings from the chapters along the way will grow into a different way of
seeing the world. The final chapter of this book is an exploration of how the
patterns interlock to form our three life-support systems – the human body,
our planet and our civilization. But you don’t have to agree with my
perspective. The essence of science is experimenting with the principles for
yourself, considering all the evidence available and then reaching your own
conclusions.
The teacup is only the start.
1
Popcorn and rockets
The gas laws
EXPLOSIONS IN THE kitchen are generally considered a bad idea. But just
occasionally, a small one can produce something delicious. A dried corn
kernel contains lots of nice food-like components – carbohydrates, proteins,
iron and potassium – but they’re very densely packed and there’s a tough
armoured shell in the way. The potential is tantalizing, but to make it edible
you need some extreme reorganization. An explosion is just the ticket, and
very conveniently, this seed carries the seeds of its own destruction within
it. Last night, I did a bit of ballistic cooking and made popcorn. It’s always
a relief to discover that a tough, unwelcoming exterior can conceal a softer
inside – but why does this one make fluff instead of blowing itself to bits?
Once the oil in the pan was hot, I added a spoonful of kernels, put the lid
on, and left it while I put the kettle on to make tea. Outside, a huge storm
was raging, and chunky raindrops were hammering against the window.
The corn sat in the oil and hissed gently. It looked to me as though nothing
was happening, but inside the pan, the show had already started. Each corn
kernel contains a germ, which is the start of a new plant, and the
endosperm, which is there as food for the new plant. The endosperm is
made up of starch packaged into granules, and it contains about 14 per cent
water. As the kernels sat in the hot oil, that water was starting to evaporate,
turning into steam. Hotter molecules move faster, so that as each kernel
heated up, there were more and more water molecules whooshing around
inside it as steam. The evolutionary purpose of a corn kernel’s shell is to
withstand assault from outside, but it now had to contain an internal
rebellion – and it was acting like a mini pressure cooker. The water
molecules that had turned to steam were trapped with nowhere to go, so the
pressure inside was building up. Molecules of gas were continually
bumping into each other and into the walls of the container, and as the
number of gas molecules increased and they moved faster, they were
hammering harder and harder on the inside of the shell.
Pressure cookers work because hot steam cooks things very effectively,
and it’s no different inside popcorn. As I searched for teabags, the starch
granules were being cooked into a pressurized gelatinous goo, and the
pressure kept going up. The outer shell of a popcorn kernel can withstand
this stress, but only up to a point. When the temperature inside approaches
180°C and the pressure gets up to nearly ten times the normal pressure of
the air around us, the goo is on the edge of victory.
I gave the pan a little shake and heard the first dull pop echoing round the
inside. After a couple of seconds, it sounded as though a mini machine gun
was being fired in there, and I could see the lid lifting as it got hit from
underneath. Each individual pop also came with a fairly impressive puff of
steam from the edge of the pan lid. I left it for a moment to pour a cup of
tea, and in those few seconds, the barrage from underneath shifted the lid
and fluff started taking flight.
At the moment of catastrophe, the rules change. Until that point, a fixed
amount of water vapour is confined, and the pressure it exerts on the inside
of the shell increases as the temperature increases. But when the hard shell
finally succumbs, the insides are exposed to the atmospheric pressure in the
rest of the pan and there is no volume limit any more. The starchy goo is
still full of hot hammering molecules but nothing is pushing back from the
other side. So it expands explosively, until the pressure inside matches the
pressure outside. Compact white goo becomes expansive white fluffy foam,
turning the entire kernel inside-out; and as it cools, it solidifies. The
transformation is complete.
Tipping the popped corn out revealed a few casualties left behind. Dark
burnt unpopped corn rattled sadly round the bottom of the pan. If the outer
shell is damaged, water vapour escapes as it is heated, and the pressure
never builds up. The reason that popcorn pops and other grains don’t is that
all the others have porous shells. If a kernel is too dry, perhaps because it
was harvested at the wrong time, there isn’t enough water inside it to build
up the pressure needed to burst the shell. Without the violence of an
explosion, inedible corn remains inedible.
I took the bowl of perfectly cooked fluff and the tea over to the window
and stood watching the storm. Destruction doesn’t always have to be a bad
thing.
There is beauty in simplicity. And it’s even more satisfying when that
beauty condenses out of complexity. For me, the laws that tell us how gases
behave are like one of those optical illusions where you think you’re seeing
one thing, and then you blink and look again and see something completely
different.
We live in a world made of atoms. Each of these tiny specks of matter is
coated with a distinctive pattern of negatively charged electrons, chaperones
to the heavy and positively charged nucleus within. Chemistry is the story
of those chaperones sharing duties between multiple atoms, shifting
formation while always obeying the strict rules of the quantum world, and
holding the captive nuclei in larger patterns called molecules. In the air I’m
breathing as I type this, there are pairs of oxygen atoms (each pair is one
oxygen molecule) moving at 900 mph bumping into pairs of nitrogen atoms
going at 200 mph, and then maybe bouncing off a water molecule going at
over 1,000 mph. It’s horrifically messy and complicated – different atoms,
different molecules, different speeds – and in each cubic centimetre of air
there are about 30,000,000,000,000,000,000 (3 × 1019) individual
molecules, each colliding about a billion times a second. You might think
that the sensible approach to all that is to quit while you’re ahead and take
up brain surgery or economic theory or hacking supercomputers instead.
Something simpler, anyway. So it’s probably just as well that the pioneers
who discovered how gases behave had no idea about any of it. Ignorance
has its uses. The idea of atoms wasn’t really a part of science until the early
1800s and absolute proof of their existence didn’t turn up until around
1905. Back in 1662, all that Robert Boyle and his assistant Robert Hooke
had was glassware, mercury, some trapped air and just the right amount of
ignorance. They found that as the pressure on a pocket of air increased, its
volume decreased. This is Boyle’s Law, and it says that gas pressure is
inversely proportional to volume. A century later, Jacques Charles found
that the volume of a gas is directly proportional to its temperature. If you
double the temperature, you double the volume. It’s almost unbelievable.
How can so much atomic complication lead to something so simple and so
consistent?
*
One last intake of air, one calm flick of its fleshy tail, and the giant leaves
the atmosphere behind. Everything this sperm whale needs to live for the
next forty-five minutes is stored in its body, and the hunt begins. The prize
is a giant squid, a rubbery monster armed with tentacles, vicious suckers
and a fearsome beak. To find its prey, the whale must venture deep into the
real darkness of the ocean, to the places never touched by sunlight. Routine
dives will reach 500–1,000 metres, and the measured record is around 2 km.
The whale probes the blackness with highly directional sonar, waiting for
the faint echo that suggests dinner might be close. And the giant squid
floats unaware and unsuspecting, because it is deaf.
The most precious treasure the whale carries down into the gloom is
oxygen, needed to sustain the chemical reactions that power the swimming
muscles, and the whale’s very life. But the gaseous oxygen supplied by the
atmosphere becomes a liability in the deep – in fact, as soon as the whale
leaves the surface, the air in its lungs becomes a problem. For every
additional metre it swims downwards, the weight of one extra metre of
water presses inwards. Nitrogen and oxygen molecules are bouncing off
each other and the lung walls, and each collision provides a minuscule
push. At the surface, the inward and outward pushes on the whale balance.
But as the giant sinks, it is squashed by the additional weight of the water
above it, and the push of the outside overwhelms the push from the inside.
So the walls of the lungs move inwards until equilibrium, the point where
the pushes are balanced once again. A balance is reached because as the
whale’s lung compresses, each of the molecules has less space and
collisions between them become more common. That means that there are
more molecules hammering outwards on each bit of the lungs, so the
pressure inside increases until the hammering molecules can compete
equally with those outside. Ten metres of water depth is enough to exert
additional pressure equivalent to a whole extra atmosphere. So even at that
depth, while it could still easily see the surface (if it were looking), the
whale’s lungs reduce to half the volume that they were. That means there
are twice as many molecular collisions on the walls, matching the doubled
pressure from outside. But the squid might be 1 km below the surface, and
at that depth the vast pressure of the water means that the lungs should
collapse to a mere 1 per cent of the volume they have at the surface.
Eventually, the whale hears the reflection of one of its loud clicks. With
shrunken lungs, and only sonar to guide it, it must now prepare for battle in
the vast darkness. The giant squid is armed, and even if it eventually
succumbs, the whale may well swim away with horrific scars. Without
oxygen from its lungs, how does it even have the energy to fight?
The problem of the shrunken lungs is that if their volume is only one-
hundredth of what it was at the surface, the pressure of the gas in there will
be one hundred times greater than atmospheric pressure. At the alveoli, the
delicate part of the lungs where oxygen and carbon dioxide are exchanged
into and out of the blood, this pressure would push both extra nitrogen and
extra oxygen to dissolve in the whale’s bloodstream. The result would be an
extreme case of what divers call ‘the bends’, and as the whale returned to
the surface the extra nitrogen would bubble up in its blood, doing all sorts
of damage. The evolutionary solution is to shut off the alveoli completely,
from the moment the whale leaves the surface. There is no alternative. But
the whale can access its energy reserves because its blood and muscles can
store an extraordinary amount of oxygen. A sperm whale has twice as much
haemoglobin as a human, and about ten times as much myoglobin (the
protein used to store energy in the muscles). While it was at the surface, the
whale was recharging these vast reservoirs. Sperm whales are never
breathing from their lungs when they make these deep dives. It’s far too
dangerous. And they’re not just using their one last breath while they’re
underwater. They’re living – and fighting – on the surplus that’s stored in
their muscles, the cache gathered during the time they spent at the surface.
No one has ever seen the battle between a sperm whale and a giant squid.
But the stomachs of dead sperm whales contain collections of squid beaks,
the only part of the squid that can’t be digested. So each whale carries its
own internal tally of fights won. As a successful whale swims back towards
the sunlight, its lungs gradually re-inflate and reconnect with its blood
supply. As the pressure decreases, the volume once again increases until it
has reached its original starting point.
Oddly, the combination of complex molecular behaviour with statistics
(not usually associated with simplicity) produces a relatively
straightforward outcome in practice. There are indeed lots of molecules and
lots of collisions and lots of different speeds, but the only two important
factors are the range of speeds that the molecules are moving at, and the
average number of times they collide with the walls of their container. The
number of collisions, and the strength of each collision (due to the speed
and mass of that molecule) determine the pressure. The push made by all
that compared with the push from the outside determines the volume. And
then the temperature has a slightly different effect.
Reaching the South Pole was a major landmark in human history. The great
polar explorers – Amundsen, Scott, Shackleton and others – are legendary
figures, and the books about their achievements and failures are some of the
greatest adventure stories of all time. And as if it wasn’t enough to deal
with unimaginable cold, lack of food, fierce oceans and clothing that wasn’t
up to the task, the mighty ideal gas law was against them, quite literally.
The centre of Antarctica is a high, dry plateau. It is covered in deep ice,
but it hardly ever snows. The bright white surface reflects almost all of the
feeble sunlight back into space, and temperatures can drop below −80°C. It
is quiet. At an atomic level, the atmosphere here is sluggish, because the air
molecules have little energy (due to the cold) and are moving relatively
slowly. Air from above descends on to the plateau, and the ice steals its
heat. Cold air becomes colder. The pressure is fixed, so this air shrinks in
volume and becomes more dense. The molecules are closer together,
moving more slowly, unable to push outwards hard enough to compete with
the air around them pushing inwards. As the land slopes away from the
centre of the continent towards the ocean, so this cold, dense air also
slithers away from the centre along the surface, unstoppable, like a slow
waterfall of air. It is funnelled through vast valleys, picking up speed as the
funnels descend outwards, always outwards towards the ocean. These are
the katabatic winds of Antarctica, and if you want to walk to the South
Pole, they will be in your face all the way. It’s hard to think of a worse trick
nature could have played on those polar explorers.
‘Katabatic’ is just a name for this sort of wind, and it’s found in many
places, not all of them cold. As they descend, those sluggish molecules do
warm up, just a little bit. And the consequences of that warming can be
dramatic.
In 2007, I was living in San Diego and working at the Scripps Institution
of Oceanography. As a northerner I was slightly suspicious of the eternal
sunshine, but I got to swim in a 50-metre outdoor pool every morning so I
couldn’t really complain. And the sunsets were amazing. San Diego is on
the coast with a clear view west across the Pacific Ocean, and the evening
skyline was reliably stunning.
I really missed the seasons, though. It seemed as though time never
moved on, almost like living in a dream. But then the Santa Ana winds
came, and it went from sunny and warm and cheerful to sullenly hot and
dry. The Santa Ana winds come every autumn, as air pours off the high
deserts and flows over the coast of California out towards the ocean. These
are also katabatic winds, just like the ones in Antarctica. But by the time
they reach the ocean, the air is much hotter at the coast than it was on the
high plateau. One memorable day, I was driving north up the I-5 freeway,
towards one of the big valleys that funnelled the hot air out to sea. There
was a river of low cloud sitting in the valley. My boyfriend at the time was
driving. ‘Can you smell smoke?’ I asked. ‘Don’t be silly,’ he said. But the
next morning, I woke up in a weird world. There were huge wildfires to the
north of San Diego, marching across the valleys, and there was ash in the
air. A campfire had got out of control in the hot, dry conditions, and the
winds were blowing the fire towards the coast. That river of cloud had been
smoke. People went to work, and either were sent home or sat in huddles
listening to the radio and wondering whether their houses were safe. We
waited. The horizon was hazy because of ash clouds you could see from
space, but the sunsets were spectacular. After three days, the smoke started
to lift. People I knew had lost their houses to the flames. Everything had a
layer of ash on it and health officials were advising against any outdoor
exercise for a week.
Up on the high plateau, hot desert air had cooled, become more dense,
and slithered downslope, just like the winds that faced Scott in Antarctica.
But the wildfires started because that air wasn’t only dry, it was hot. Why
would it get hotter as it came downhill? Where does the energy come from?
The ideal gas law still applies – this was a fixed mass of air, and it was
moving so quickly that there was no time for it to exchange energy with its
surroundings. As that stream of dense air made its way downhill, the
atmosphere that was already at the bottom of the hill pushed on it, because
the pressure down there was higher. Pushing on something is a way of
giving it energy. You can imagine individual air molecules hitting the wall
of a balloon that is moving towards them. They’ll bounce off with more
energy than they had to start with, because they’re bouncing off a moving
surface. So the volume of the air in the Santa Ana winds decreased because
it was squeezed inwards by the surrounding atmosphere. That squeezing
gave the travelling air molecules extra energy, and so the temperature of the
wind increased. It’s called adiabatic heating. Every year, when the Santa
Ana winds come, everyone in California is extra vigilant about open fires.
After a few days of such hot, dry air stealing the moisture from the
landscape, sparks can easily turn into wildfires. And the heat doesn’t just
come from the California sun – it also comes from the extra energy given to
the gas molecules as they are compressed by denser air closer to the ocean.
Anything that changes the average speed of air molecules will change the
temperature.
The same thing happens in reverse when you squirt whipped cream out
of a can. The air that comes out in the cream has expanded suddenly and
pushed on its surroundings, so it has given away energy and cooled down.
The nozzle of the squirty cream canister feels cold to the touch for this
reason – the gas that’s coming through it is giving away its energy as it
reaches the free atmosphere. Less energy is left behind, so the can feels
cool.
Air pressure is just a measure of how hard all those tiny molecules are
hammering on a surface. Normally we don’t notice it much because the
hammering is the same from every direction – if I hold up a piece of paper,
it doesn’t move because it’s getting pushed equally from both sides. Each
one of us is getting pushed by air all the time, and we hardly feel it at all. So
it took people a long time to work out how hard that push actually is, and
when it came along, the answer was a bit of a shock. The magnitude of the
discovery was easy to appreciate because the demonstration was unusually
memorable. It’s not often that an important scientific experiment is also set
up to be a theatrical spectacle, but this one had all the proper ingredients:
horses, suspense, an astonishing outcome and the Holy Roman Emperor
looking on.
The difficulty was that to work out how hard air is pushing on something,
you really need to take away all the air on the other side of it, leaving
behind a vacuum. In the fourth century BC, Aristotle had declared that
‘nature abhors a vacuum’, and that was still the prevailing view nearly a
thousand years later. Creating a vacuum seemed out of the question. But
some time around 1650, Otto von Guericke invented the first vacuum
pump. Instead of writing a technical paper about it and disappearing into
fn2
obscurity, he chose spectacle to make his point. It probably helped that he
was a well-known politician and diplomat, and was on good terms with the
rulers of his day.
On 8 May 1654, Ferdinand III, the Holy Roman Emperor and overlord of
a large part of Europe, joined his courtiers outside the Reichstag in Bavaria.
Otto brought out a hollow sphere, 50 cm in diameter and made of thick
copper. It was split into two separate halves with a smooth, flat surface
where they touched. Each half had a loop attached to the outside, so that
two ropes could be tied on and used to pull the halves apart. He greased the
flat surfaces, pushed the two sides together, and used his new vacuum pump
to remove the air from the inside of the sphere. There was nothing to hold it
together, but after the air had been removed, the two halves behaved as
though they were glued to each other. Otto had realized that the vacuum
pump gave him a way to see how strongly the atmosphere could push.
Billions of minuscule air molecules were hammering all over the outside of
the sphere, pushing the halves together. But there was nothing inside to
fn3
push back. You could only pull the two hemispheres apart if you could
pull harder than the air could push.
Then the horses were assembled. A team was hitched to each side of the
sphere, pulling in opposite directions in a giant tug of war. As the Emperor
and his retinue looked on, the horses strained against the invisible air. The
only thing holding the sphere together was the force of air molecules hitting
something the size of a large beach ball. But the strength of thirty horses
could not pull the sphere apart. When the tug of war had finished, Otto
opened the valve to let air into the sphere, and the two halves just fell open.
There was no question about the winner. Air pressure was far stronger than
anyone had suspected. If you take all the air out of a sphere that size and
hang it vertically, the upward push of the air could theoretically support
2,000 kg, the weight of a large adult rhino. That means that if you draw a
circle 50 cm in diameter on the floor, the push of the air on just that bit of
floor is also equal to the weight of a 2,000 kg rhino. Those tiny invisible
molecules are hitting us very hard indeed. Otto did this demonstration many
times for different audiences, and the sphere became known as a
Magdeburg sphere, named after his home town.
Otto’s experiments became famous partly because others wrote about
them. His ideas first reached the scientific mainstream in a book by Gaspar
Schott, published in 1657. It was reading about Otto’s vacuum pump that
inspired Robert Boyle and Robert Hooke to carry out their experiments on
gas pressure.
You can try a version of this for yourself, without the need for either
horses or emperors. Find a square of thick, flat cardboard that’s large
enough to cover the mouth of a glass. It’s best to try this over a sink, just in
case. Fill up the glass with water right to the rim and put the cardboard on
top. Push it flat against the rim of the glass so there’s no air left between the
surface of the water and the cardboard. Then turn the glass over – and
remove your hand. The cardboard, supporting the entire weight of the
water, will stay put. It stays there because air molecules are hitting it from
the underside, pushing the cardboard upwards. That push is easily enough
to hold the water up.
The battering of air molecules isn’t just useful for keeping things in
place. It can also be used to move things around, and humans weren’t the
first ones to take advantage of that. Let’s meet an elephant, one of the
Earth’s most impressive experts at manipulating its environment with air.
An African bush elephant is a majestic giant, usually found ambling
peacefully across dusty dry savannah. Elephant family life is based around
groups of females. An elder stateswoman, the matriarch, leads each group
as they roam in search of food and water, relying on her memory of the
landscape to make decisions. But these animals don’t just depend on their
heft to survive. Each elephant may have a heavy lumbering body, but to
make up for it, it’s got one of the most delicate and sensitive tools in the
animal kingdom: a trunk. As a family group moves around, they’re
constantly exploring the world with this odd appendage, signalling, sniffing,
eating and snorting.
An elephant’s trunk is fascinating in many ways. It’s a network of
interlocking muscles, capable of bending and lifting and picking up objects
with incredible dexterity. If that were all it is, it would be useful enough, but
it’s made even better by the two nostrils that run down the length of the
trunk. These nostrils are flexible pipes that join the snuffling trunk tip to the
elephant’s lungs, and this is where the real fun starts.
As our elephant and her family group approach a watering hole, the ‘still’
air around them is bumping and jostling just like anywhere else, hammering
against their wrinkly grey skin, against the ground and against the water
surface. The matriarch is slightly ahead of the group, swinging her trunk as
she saunters into the pool and sends ripples through her reflection. She dips
her trunk into the water, closes her mouth, and the huge muscles around her
chest lift and expand her ribcage. As her lungs expand, the air molecules
inside spread out to take up the new space. But that means that right down
inside the tip of the trunk, where the cool water touches the air in her
nostrils, there are fewer air molecules hitting the water. The ones that are
there are going just as fast, but there aren’t as many collisions. The
consequence is that the pressure inside her lungs has dropped. Now the
atmosphere is winning in the shoving contest between the air molecules
hitting the pool of water and the air molecules inside the matriarch. The
push from inside can’t match the push from outside any more, and the water
is just the stuff in the middle of the competition. So the atmosphere pushes
water up the elephant’s trunk, because what’s inside can’t push back. Once
the water has taken up some of the extra space, the air molecules inside are
as close together as they were at the start, and the water doesn’t move any
further.
Elephants can’t drink through their trunks – if they tried, they’d cough
just as you would if you tried to drink through your nose. So once the
matriarch has perhaps 8 litres of water inside her trunk, she stops expanding
her ribcage. Curling her trunk up and under, she points the tip into her
mouth. Then she uses her chest muscles to squash her chest, decreasing the
size of her lungs. As the air molecules inside are squashed closer together,
the water surface halfway up her trunk gets hit much more often. The battle
of the air inside and the air outside reverses, and the water is pushed out of
the trunk into the elephant’s mouth. Our matriarch is controlling the volume
of her lungs to control how hard the air inside her is pushing on the outside.
If she shuts her mouth, the only place where anything can move is at her
trunk, and whatever is at the tip of her trunk will get pushed in or pushed
out. An elephant’s trunk and lungs together are a combined tool for
manipulating air so that the air, rather than the elephant herself, does the
pushing.
fn4
We do the same thing when we suck a drink up a straw. As we expand
our lungs, the air inside is spread more thinly. There are fewer air molecules
inside the straw to push on the water surface. And so the atmosphere
pushing on the rest of the drink pushes the drink up the straw. We call this
sucking, but we’re not pulling on the drink. The atmosphere is pushing it up
the straw, doing the work for us. Even something as heavy as water can be
shunted about if the hammering of the air molecules is harder on one side
than the other.
However, sucking air up a trunk or a straw has limits. The bigger the
pressure difference between the two ends, the harder the push will be. But
the biggest difference you can possibly make when you’re sucking is the
difference between the pressure of the atmosphere and zero. Even with a
perfect vacuum pump instead of lungs, you couldn’t drink through a vertical
straw that was longer than 10.2 metres, because our atmosphere can’t push
water any higher than that. So to exploit to the full the ability of gas
molecules to push things around, you need to get them working at higher
pressures. The atmosphere can push hard, but if you force another gas to be
hotter and put it under greater pressure, it can push harder. Get enough tiny
gas molecules hitting something often enough and fast enough, and you can
move a civilization.
A steam locomotive is a dragon made of iron, a hissing, breathing,
muscular beast. Less than a century ago these dragons were everywhere,
hauling the products of industry and the needs of society across whole
countries and expanding the world of their passengers. They were mundane
and noisy and polluting, but they were beautiful pieces of engineering.
When they became obsolete, the dragons were not allowed to die because
society just couldn’t let them go. They’ve been kept alive by volunteers,
enthusiasts and a deep well of affection. I grew up in the north of England,
and so my childhood years were steeped in the history of the Industrial
Revolution: mills, canals, factories and, more than anything else, steam. But
I live in London now, and so it’s easy to forget. A trip along the Bluebell
steam railway with my sister brought it all back.
It was a chilly winter day, absolutely perfect for a journey propelled by
steam with the promise of tea and scones at the other end. We didn’t spend
too long at the station where we started, but when we arrived at Sheffield
Park, we stepped off the train into a slow but steady hum of activity. The
engines were constantly being tended by an ever-changing swarm of
humans who seemed tiny beside their iron beasts. The humans involved
with the engines were easy to identify: blue overalls, peaked caps, a jolly
demeanour, an optional beard and in between engine-tending duties usually
to be found leaning on something. As my sister pointed out, an awful lot of
them seemed to be called Dave. The beauty of a steam engine is that the
principle behind it is fantastically simple, but the raw power produced
needs to be goaded, tamed and nurtured. A steam engine and its humans are
a team.
Standing on the ground, looking up at one large, black engine, it was hard
to comprehend that in its heart, this was basically a furnace on wheels
heating a giant kettle. One of the Daves invited us into the cab. We climbed
up the ladder just behind the engine and found ourselves in a grotto full of
brass levers, dials and pipes. There were also two white enamel mugs and a
sandwich tucked behind one of the pipes. But the best thing about the cab
was that we could see right into the belly of the beast. The giant furnace at
the heart of a steam engine is filled with fiery coals burning an intense
yellow. The fireman gave me a shovel and told me to feed it, and so I
obediently scooped coal from the tender behind me into the glowing mouth.
The engine is hungry. On one 18-km journey, it will burn through 500 kg of
coal. That half-tonne of solid black gold is converted into gas – carbon
dioxide and water; and the burning releases enormous amounts of energy,
so those gases are extremely hot. This is the start of the energy conversion
that powers the train.
When you look at a steam engine, the main feature is the long cylinder of
the ‘engine’ itself, stretching from the cabin to the funnel. I’d never really
thought about what was in there, but it’s full of tubes. The tubes are
carrying the hot gas from the firebox through the engine, and this is the
kettle. Most of the space around the tubes is taken up by water, a giant bath
of bubbling, boiling liquid. As this is heated by the tubes, it produces steam,
hot water molecules that are zooming about in the space right at the top of
the engine at very high speeds. This is what most of the steam engine is:
furnace and kettle, producing vast clouds of hot water vapour. This dragon
isn’t breathing fire, it’s breathing billions of energetic molecules, all
whizzing about at gigantic speeds but trapped in the engine. The
temperature of that gas is about 180°C and the pressure in the top of the
kettle is about ten times as high as atmospheric pressure. The molecules are
thumping hard against the walls of the engine, but they can only escape
after being put to work.
We climbed down from the cab and walked up to the front. The towering
engine, the half-tonne of coal, the giant kettle and the human teamwork
were all in service to what we found there: two cylinders containing pistons,
each about 50 cm in diameter and 70 cm long. It’s down here at the front,
dwarfed by the dragon above, that the real work is done. The hot, high-
pressure steam is fed into one cylinder at a time. The atmospheric pressure
on the other side of the piston is no match for the ten atmospheres that the
dragon has breathed out. The hammering molecules shove the piston along
the cylinder, and then are finally released to the atmosphere with a satisfied
‘chuff’. This is what you’re hearing when a steam engine’s familiar ‘chuff,
chuff, chuff’ comes towards you. It’s the release to the atmosphere of water
vapour whose work is done. The piston drives the wheels, and the wheels
grip the rails and drag the carriages. We know that steam engines need vast
quantities of coal to keep them going, but almost no one talks about the
water used on every journey. The 500 kg of coal that is shovelled into this
engine on each trip is used to convert 4,500 litres of water to gas, and then
that gas pushes on a piston and is lost to the atmosphere one ‘chuff’ at a
fn5
time.
Finally it was time to leave the engine and get back in one of the
carriages to be carried home. The return journey felt different. The billows
of steam whooshing past the windows had made their contribution to our
excursion. Instead of appearing loud and intrusive, the engine pulling us
along seemed relatively quiet and calm considering what was going on
inside it. It would be lovely if someone could make a glass steam
locomotive one day, so that we could all see the beast at work.
The steam revolution in the early 1800s was all about using the push of
gas molecules to do something useful. All you need is a surface with gas
molecules hitting one side harder than on the other. That push could lift the
lid of a pan as you cook, or it could be used to transport food and fuel and
people, but it comes from the same basic principles. We don’t use steam
engines any more, but we do still use that push. A steam engine is
technically an ‘external combustion engine’ because the furnace is separate
from the kettle. In a car engine, the burning happens in the cylinder –
gasoline burns right next to the piston and the burning itself produces hot
gas to shove the piston along. That’s classed as an internal combustion
engine. Every time you get into a car or a bus, you’re being carried along by
the push of gas molecules.
It’s easy to play with the effects of pressure and volume, especially if you
can find a wide-necked bottle and a hard-boiled egg with its shell taken off.
The neck of the bottle needs to be just a bit narrower than the egg, so that
the egg will perch on top of it without falling in. Light some paper, drop it
into the bottle, let it burn for a few seconds and then put the egg back on
top. After a while, you’ll see the egg squeeze itself down inside the bottle.
That’s a bit weird, and it’s inconvenient that now you have an egg in a
bottle and it won’t come out. There are a few solutions, but one of them is
to turn the bottle upside-down so that the egg is sitting in the neck, and then
run the bottle under a hot tap. After a while, the egg will come whooshing
out.
The game here is that you have a fixed mass of gas (in the bottle) and a
way to tell whether the pressure inside is higher or lower than the pressure
of the atmosphere. If the egg is blocking the neck, the volume of gas inside
is fixed. If you increase the temperature by setting fire to something, the
pressure inside will increase and air will escape around the sides of the egg
(if the egg is sitting on top). When it cools down again, the pressure inside
will decrease (since the volume is fixed) and the egg will be pushed inside,
because the push from outside is now greater than the push back from
inside. You can get the egg to move just using the heating and cooling of air
in a container with a fixed volume.
The high pressures in a steam engine are controlled and stable, ideal for
pushing pistons and making wheels turn. But that’s not the end of it. Why
waste energy on intermediate stages between the gas and the wheels? Why
not just let the hot high-pressure gases shove your vehicle forward directly?
That’s how guns, cannons and fireworks have always worked, although the
early ones were all notoriously unreliable. But by the early 1900s,
technology and ambition had moved on. Along came the rocket, the most
extreme form of direct propulsion ever invented.
It wasn’t until after the First World War that the necessary technology
reached any degree of reliability, but by the 1930s you could launch a
rocket that would probably go in the right direction and probably wouldn’t
kill anyone. Most of the time. As with many new technologies, inventors
made it work before anyone knew what to do with it. And out of the fertile
pond of enthusiastic human inventiveness came something very new and
modern-sounding and utterly doomed: rocket post.
In Europe, rocket post only really happened because of one man: Gerhard
Zucker. A few inventors at that time were tinkering with rockets, but
Zucker led the field in dogged persistence and unfailing optimism in the
face of continual discouragement. This young German was obsessed by
rockets, and since the military weren’t interested in what he was doing he
looked to the civilian world for an excuse to continue. Sending mail by
rocket sounded to him like something the world was crying out for – fast,
capable of crossing the sea and covered in the glitter of novelty. The
Germans tolerated his early (unsuccessful) experiments and then decided
they’d had enough, so Zucker came to the UK. Here he found friends and
support in the stamp-collecting community, who liked the idea of a new
kind of novelty stamp to go with a new kind of novelty mail delivery
system. Things were looking up. After a small-scale test in Hampshire,
Zucker was sent up to Scotland in July 1934 to test sending his mail rocket
between two islands, Scarp and Harris.
Zucker’s rocket wasn’t particularly sophisticated. The main body of it
was a large metal cylinder about a metre long. Inside, a narrow copper tube
with a nozzle at its back end was filled with packed powder explosive. The
space in between the inner tube and outer cylinder was filled with letters,
and there was a pointy nose on the front with a spring in it, presumably to
help soften the landing. Rather sweetly, on Zucker’s diagram of the setup,
the thin layer between the explosive and the highly flammable letters is
labelled ‘asbestos packing round cartridge, to prevent damage to mails’.
The rocket was laid down on its side on a slanted trestle, pointed upwards
and sideways. At the moment of launch, a battery would ignite the
explosive, and the burning would produce vast quantities of hot, high-
pressure gas. The gas molecules, now moving at high speed, would bounce
off the inside of the front end of the rocket, driving it forward, but there
would be no equivalent push at the back end – gas would just escape
through the nozzle to the atmosphere. This imbalance in pushing could
drive the rocket forward very quickly. The explosive burn would continue
for a few seconds, enough to push the rocket high into the air and over the
channel between the islands. There didn’t seem to be too much concern
about how and where it would land, but that was one reason for trying it out
in a very remote part of Scotland surrounded by sea.
Zucker collected 1,200 letters to send as part of the trial, each adorned
with a special stamp that said ‘Western Isles Rocket Post’. He packed as
many as would fit inside his rocket and set up the trestle, watched by a
bemused crowd of locals and an early BBC TV camera. The moment had
come.
When the launch button was pressed, the battery ignited the explosive.
The rapid burning generated the expected mixture of hot gases inside the
copper tube, and the energetic molecules hammered on the front of the
rocket, shoving it up the trestle at high speed. But after only a couple of
seconds, there was a loud, dull thud and the rocket disappeared behind a
plume of smoke. As the smoke cleared, hundreds of letters could be seen
fluttering to the ground. The asbestos had done its job, but the rocket
hadn’t. Hot, high-pressure gas is hard to control, and the energetic
molecules had broken the casing. Zucker blamed the explosive cartridge,
and set about collecting the letters and preparing for a second trial.
A few days later, 793 surviving letters from the first rocket and also 142
new ones were packed into a second rocket. This one was launched from
the other island, Harris, back towards Scarp. But Zucker was out of luck.
The second rocket also exploded on the launch pad, this time with an even
louder bang. The surviving letters were collected up again and sent to their
recipients by the conventional mail system, with singed edges as souvenirs.
The trial was abandoned. For the next few years, Zucker stubbornly carried
fn6
on, always convinced that next time, it would work. But it never did, at
least not for mail. Zucker pushed hard against the unknown, and it’s only
hindsight that tells us that it wasn’t the right time or the right place or the
right idea. If it had been all three, we’d hail him as a genius. But small-scale
rocketry was just too unreliable and fiddly to deliver messages better and
faster than motorized transport and the telegraph. In a way he was right:
using hot, high-pressure gases as a propellant has enormous potential to get
things from A to B. But it was others who took the principle, found a
suitable application, and solved the practical problems until it became a
success. Rocket development became the preserve of the military, with the
German V1 and V2 rockets used in the Second World War showing the
way, and civilian space programmes taking over after that.
These days, we are all familiar with images of giant rockets carrying
huge cargoes of people and equipment to the International Space Station, or
taking satellites into orbit. Rockets can seem frighteningly powerful, and
the modern control systems that now make them safe and reliable are a
huge human achievement. But the basic mechanism behind every Saturn V
rocket, every Soyuz and Arianne and Falcon 9 that has ever flown, is the
same as it was for Gerhard Zucker’s primitive mail rocket. If you make
enough hot, high-pressure gas quickly enough, you can make use of the
huge cumulative force that comes from billions of individual molecules
bumping into things. The flight pressure in the first stage of a Soyuz rocket
is about sixty times greater than atmospheric pressure, so the push is sixty
times stronger than the normal push of the air. But it’s exactly the same type
of push: just molecules bumping into things. Vast quantities of them
colliding often enough and fast enough can send a man to the Moon. Never
underestimate things that are too small to see!
Gas molecules are always with us. The Earth has an atmosphere that
surrounds us, bumps into us, pushes on us and also keeps us alive. The
wonderful thing about our atmosphere is that it isn’t static – it’s constantly
shifting around and changing. The air is invisible to us, but if we could see
it, we’d see huge masses of it heating up and cooling down, expanding and
contracting, always moving. What our atmosphere is doing is dictated by
the gas laws we’ve seen at work in this chapter, just like any other
collection of gas molecules. Even though it’s not contained in a whale’s
lungs or a steam engine, it’s still pushing. But since its surroundings are
also air, that means that it’s constantly pushing itself around, readjusting to
conditions. We can’t see the details, but we have a name for the
consequences: weather.
The best place to watch a storm is a vast open plain. The day before, the
air can be calm and the expanse of blue up above seems to go on for ever.
Invisible air molecules crowd together close to the ground and spread out
further up, always pushing, hassling, readjusting and flowing. Air is
shunted from regions of high pressure to regions of low pressure,
responding to heating and cooling, always on the way to somewhere else.
But the adjustments are slow and peaceful, and there’s no hint of the vast
amounts of energy carried by the molecules.
The day of the storm dawns just as the one before it did, but the sky is
clearer, so the ground is heating up more quickly. The air molecules take
some of that energy and speed up. By early afternoon, a deep wall of cloud
is approaching and expanding as it moves, until it stretches across the
horizon. Energy is on the move. A pressure difference is pushing this slab
of gaseous architecture across the plain. The drama comes because this
giant structure is unstable. Although the air molecules are shoving hard on
each other, they haven’t had time to rearrange themselves into a more
balanced situation. Alongside that, vast amounts of energy are being
shunted around, so the situation is constantly changing. Hot air warmed by
the ground is pushing upwards into the cloud, pummelling its way through
and building towers that stretch high above the wall.
As the thundercloud arrives overhead, the expansive blue is replaced by a
dark low lid on the landscape. On the ground, we are boxed in by the clash
that is going on above. We can’t see the air molecules, but we can see the
clouds churning and surging. And this is only a hint of the violence going
on within them as air packets are buffeted and pummelled, because the
imbalances of pressure are so strong that readjustment is a rapid and
energetic process. As energy is exchanged by the air molecules, water
droplets cool and grow and the first large raindrops start to fall. Strong
winds stream past us, as the air molecules rush around even at ground level.
Big storm clouds remind us how much energy there is up there in the
blue sky. We see hints of the bumping and shoving, and it looks extreme –
but it’s only the merest hint of the real bumping and shoving happening at a
molecular level above our heads. Air molecules may absorb energy from
the Sun, lose energy to the ocean, gain energy from condensation as clouds
form or lose energy by radiating it away to space, and they are constantly
adjusting according to the ideal gas law. Our spinning planet with its rough
and multicoloured surface makes the adjustments more complicated, and so
do clouds, tiny particulates, and the specific gases present. A weather
forecast is really just a way of keeping track of the battles above our heads
and picking out the ones that will affect us most down here on the ground.
But the action right at the root of it all is the same as that used by an
elephant, a rocket and a steam engine. It’s all just the gas laws in action.
The same bit of physics that makes popcorn pop also makes the weather
work.
2
What goes up must come down
Gravity
My scientific research is all about the physics of the ocean surface. I’m an
experimentalist, and so part of my job is to go out on the ocean and measure
what’s happening at this messy, beautiful boundary between the air and the
sea. And that means spending weeks working on a research ship, a floating,
functional, mobile scientific village. The problem with living on a ship is
that you have to live with gravity basically having gone wrong. ‘Down’
becomes an uncertain concept. Things may fall at the same speed and in the
same direction as if you’d dropped them on land, but then again, they may
not. If you spot a loose object just sitting on a table, you tend to find
yourself watching it suspiciously because there is no guarantee that it’s
going to stay put. Life at sea is full of elastic bungees, string, rope, sticky
grippy mats, locked drawers – anything that helps to keep life organized
when there’s a capricious force pulling things in unpredictable directions,
like a scientific poltergeist. My specific research topic is the bubbles
produced by breaking waves in storms, and so I’ve spent months living at
sea in some pretty nasty conditions. I actually quite like it – you adapt very
quickly – but it’s a good lesson in how much we take gravity for granted.
On one research ship in the Antarctic, the ship’s purser used to marshal the
unreasonably enthusiastic among us for circuit training three times each
week. We’d gather in the hold, an echoing iron space down in the guts of
the ship, and obediently jump and lift and skip for an hour. It was probably
the most effective circuit training I’ve ever done, because you never knew
what force you were going to have to resist. The first three sit-ups might be
ridiculously easy, because the ship was heaving downwards, effectively
reducing gravity. You’d just start feeling really good about yourself when
the penalty arrived as the ship reached the bottom of the trough of the wave.
At that point, gravity was effectively 50 per cent stronger, and suddenly it
felt as though your tummy muscles had to fight against strips of elastic
pinning you to the floor. Four more sit-ups and gravity would vanish again
… Anything involving jumping was even worse because you were never
quite sure where the floor was. And then afterwards, in the shower, you’d
spend your time chasing the water flow around the shower cubicle, as the
rolling of the ship made it impossible to predict where it was going to fall.
Of course, there was nothing wrong with gravity itself. Everything on
that ship was being pulled towards the centre of the Earth with the same
force. But when you feel the force of gravity, you’re resisting an
acceleration. If your surroundings are accelerating all by themselves as the
giant tin can you’re living in is tossed around by nature, your body can’t tell
the difference between gravitational acceleration and any other acceleration
that’s going on. So you get ‘effective gravity’, which is what you
experience overall, without worrying about where it’s coming from. That’s
why that odd feeling you get in a lift only happens at the beginning and end
of the ride – when the lift is accelerating towards its top speed, or
decelerating (a negative acceleration) towards a halt. Your body can’t tell
the difference between the acceleration of the lift and the acceleration due
fn3
to gravity, so you experience a reduced or increased ‘effective gravity’.
For a fraction of a second, you can experience what it might be like to live
on a planet with a different gravity field.
Fortunately for us, we’re completely free of these complications most of
the time. Gravity is constant and points towards the centre of the Earth.
‘Down’ is the direction in which things fall. Even plants know that.
My mother is a keen gardener, so when I was growing up I had a lot of
opportunities to plant seeds, chop up weeds, wrinkle my nose in disgust at
slugs and turn over compost heaps. I remember being fascinated by
seedlings because they clearly knew up from down. Down in the darkness
of the soil, a seed case would open and new roots would creep downwards
while a nascent shoot explored upwards. You could pull up a young
seedling and see that there had been no hesitation or exploration. The root
just went straight down and the shoot went straight up. How did it know?
When I was a bit older, I found the answer and it’s delightfully simple. It
turns out that inside the seed there are specialized cells called statocytes that
are mini plant snow-globes. Inside each one there are specialized starch
grains that are more dense than the rest of the cell, and they settle towards
the bottom of the cells. Protein networks can sense where they are, and so
the seed, and later the plant, knows which way is up. Next time you plant a
seed, turn it over and think about the mini snow-globe inside, and then plant
it whichever way up you like, because the plant can solve the puzzle.
Gravity is a fantastically useful tool. Plumb lines and spirit levels are
cheap and accurate. ‘Down’ is universally accessible. But if everything is
pulling on everything else, what about the mountain that I can see in the
distance? Isn’t that pulling on me? What’s so special about the centre of the
planet?
I love coastlines for all sorts of reasons (waves, bubbles, sunsets and sea
breezes), but what I love most of all is the liberating, luxurious feeling of
taking in the vast expanse of the sea. When I lived in California, I shared a
tiny house very close to the beach, so close that we could hear the waves at
night. There was an orange tree in the back garden, and a porch for
watching the world go by. The ultimate luxury at the end of a busy day was
to walk to the end of the road, sit on the smooth, worn rocks and look out at
the Pacific Ocean. When I did that sort of thing in England as a kid, I was
just watching for fish or birds or big waves. But when I watched the ocean
in San Diego, I was imagining the planet. The Pacific Ocean is vast, taking
up a full third of the circumference of the Earth at the equator. Looking out
towards the sunset, I could imagine the giant ball of rock I was living on,
Alaska and the Arctic far away to my right, in the north, and the full length
of the Andes running all the way to Antarctica on my left, south of me. I
could almost give myself vertigo watching it all in my head. And once, it
occurred to me that I was directly experiencing all of those places. Each one
of them was tugging on me, and I was tugging on them. Every bit of mass is
pulling on every other bit of mass. Gravity is a fantastically weak force –
even a small child can generate the force to resist the gravitational pull of a
whole planet. But nevertheless, each of those minuscule tugs is still there.
Together, an uncountable number of minute tugs adds up to a single force,
the gravity that we experience.
This was the step taken by the great scientist Isaac Newton when he
published his Law of Universal Gravitation in Philosophiae Naturalis
Principia Mathematica – the famous Principia – in 1687. Using the rule
that the gravitational force between two things is inversely proportional to
the square of the distance separating them, he showed that if you added up
the pull of every single bit of a planet, quite a lot of those sideways pulls
cancelled each other out, and the result was a single downward force,
pointing towards the centre of the planet and proportional to the Earth’s
mass and the mass of the thing being pulled. A mountain that’s twice as far
away will only pull on you with a quarter of the force. So distant objects
matter less. But they still count. Sitting looking out at the sunset, I was
being pulled sideways to the north and a bit downwards by Alaska and
sideways to the south plus a bit downwards by the Andes. But the pulls to
the north and the south cancelled each other out, and what was left over was
downwards.
So even though we’re all being pulled on (right now) by the Himalaya,
the Sydney Opera House, the Earth’s inner core and lots of marine snails,
we don’t need to know the details. The complexities sort themselves out,
and leave us with a simple tool. To predict the Earth’s pull on me, I just
need to know how far away the centre of it is, and the mass of the whole
planet. The beauty of Newton’s theory was that it was simple, it was
elegant, and it worked.
But it’s still true that forces are weird. In spite of its brilliance, Isaac
Newton’s explanation of gravity had one major flaw: there was no
mechanism. It’s straightforward to state that the Earth is pulling on an
fn4
apple, but what is doing the pulling? Are there invisible strings? Pixies?
This wasn’t sorted out satisfactorily until Einstein worked out the Theory of
General Relativity, but for the 230 years in the middle, Newton’s model of
gravity was accepted (and is still widely used today) because it worked so
incredibly well.
You can’t see forces, but almost every kitchen has a device in it for
measuring them. That’s because you need something important for cooking
(and especially for baking) that no glossy recipe book ever mentions. It’s
necessary because quantities matter: you have to measure ‘stuff’, and you
have to do it accurately. The unmentioned critical ingredient that lets you do
this is simple: something (anything) the size of a planet. It’s very fortunate
for all fans of Eccles cakes, Victoria sponge and chocolate gateau that we’re
sitting right on top of one.
I’ve got a book full of hand-written recipes that I’ve been adding to since
I was eight or nine years old, and I love being able to go straight back to the
ones of my childhood. Carrot cake is one of those, scribbled on a page
smudged with the years, and the recipe starts with an instruction to procure
200 g of plain flour. So the baker does something very clever that we all
take completely for granted. They put some flour in a bowl and measure
directly how much the Earth is pulling on it. That’s what scales do. You put
them in the gap between the vast planet and the tiny bowl, and measure the
squeeze. The pull between an object and our planet is directly proportional
to the mass of both the object and the Earth. Since the mass of the Earth
isn’t changing, that pull depends solely on the mass of the flour that went
into the bowl. Scales measure weight, which is the force between flour and
planet. But the weight is just the mass of the flour multiplied by the strength
of gravity, which is a constant in our kitchens. So if you measure the weight
and you know the strength of gravity, you can work out the mass of flour in
the bowl. Next you need 100 g of butter, so you put butter into the bowl
until the squeezing force is half what it was before. This is a fantastically
useful and very simple technique for getting at how much stuff you have,
and it works for everyone on the planet. Heavy objects are heavy only
because they consist of more ‘stuff’, so Earth is pulling harder on them.
Nothing is heavy in space because the local gravity is too weak to pull
noticeably on things, unless you’re very close to a planet or a star.
But what those kitchen scales are really telling you is that gravity, the
grand force that holds together our planet and our solar system and
dominates our civilization, is unbelievably weak and weedy. The Earth has
a mass of 6 × 1024 kg (6 thousand billion billion tons, if you prefer those
units), and it can only pull on your bowl of flour with the force of a small
elastic band. It’s just as well, otherwise life wouldn’t be able to exist, but it
does put things into perspective a bit. Every time you pick up an object, you
are resisting the gravitational pull of a whole planet. The solar system is
large because gravity is weak. Gravity does have one major advantage over
all the other fundamental forces, though, and that’s its reach. It may be
weak, and getting even weaker as you travel further from Earth, but it
stretches out across the vast distances of space, tugging on other planets and
suns and galaxies. Each tug is tiny, but it’s this frail force field that gives
our universe its structure.
Still, even picking up the finished carrot cake does take some effort.
When it’s sitting on the table, the table surface is pushing upwards on the
cake just enough to exactly balance out the pull between cake and planet.
To pick it up, you have to provide that much force plus a tiny bit more,
enough so that the overall force on the cake is upwards. Our lives are
controlled not by what individual forces are acting, but on what’s left over
on balance. And that simplifies things a lot. Massive forces can be made
irrelevant by setting them in opposition to other massive forces. The easiest
place to start thinking about this is with solid objects, because they keep
their shape while they’re being pulled. And London’s Tower Bridge is very
solid indeed.
Gravity can be a nuisance, because sometimes you want to hold things in
the air. To do that, you need to resist the downward pull. If you couldn’t,
everything would slither around on the floor. Fluids flow downwards, and
that’s just the way it is. For solids, things are different. One single concept,
the pivot, lets us effectively neutralize gravity by turning stupidly heavy
things into one half of a see-saw. The mysterious other half is often
cunningly hidden away, and there’s no better example of this than the two
graceful towers of Tower Bridge in London. Built on two man-made
islands, each a third of the way across the Thames, these towers stand guard
at the entrance into London from the sea, and carry the road linking the
north of the city to the south.
The pavement is a noisy circus of tourists engaged in camera
choreography, while London’s taxis, souvenir merchants, coffee stalls, dog-
walkers and buses just get on with things in the background. Our tour guide
strides through all the chaos and we toddle along behind him like a line of
obedient ducklings. He opens an iron gate at the base of one of the towers,
ushers us round the corner into a sort of posh garden shed made of stone,
and suddenly it’s calm. You can almost hear the sigh of relief as his flock
realize they’ve survived the tourist gauntlet and have arrived at their
reward: the brass dials, giant levers and reassuringly robust-looking valves
of solid Victorian engineering. The pretty and delicate fairy-castle exterior
of Tower Bridge is famous around the world, but we are here to see what’s
lurking inside: the massive steel guts of this elegant and powerful beast.
London has been a port for two thousand years, and the nice thing about
having a city on a river is that you have two banks to play with, not just one
bit of coastline. But the Thames is both a vital highway for anything that
floats and a massive obstacle for anything that walks or rolls. Bridges came
and went over the centuries, and by the 1870s the city was crying out for
another one. The problem was: how do you satisfy the owner of the horse
and cart without cutting off the river to the tall ships? Tower Bridge is the
ingenious solution.
The small stone shed squats on top of a spiral staircase that leads
downwards into a series of improbably large brick grottos, hidden inside the
foundations of the tower. It’s like the wardrobe that leads into Narnia,
except that this is Narnia for engineers. The first grotto contains the original
hydraulic pumps and the next (much bigger) one is mostly filled by a
wooden monster: a two-storey-high barrel which used to act as a temporary
energy store – a non-electric battery. But it’s the third one, the largest of all,
that I’ve really come to see. This is the chamber that houses the
counterweight.
The road between the two towers splits into two separate halves. About a
thousand times each year, a ship or a boat arrives at the bridge, and the
traffic is stopped. Each half of the road swings upwards, and on the other
side of the pivot, in this dark chamber underneath the tower, the hidden half
of the bridge swings downwards. I look upwards at the underside of this
seesaw, and ask what exactly is hanging above us. Glen, our guide, is quite
cheery about it. ‘Oh, there’s about 460 tons of lead ingots and bits of pig
iron up there,’ he says. ‘It rattles around loose – you can hear it when the
bridge opens. When they change anything on the bridge, they usually add a
bit or take some away so it stays perfectly balanced.’ We are apparently
standing directly underneath the biggest beanbag in the world.
It’s the balance that is the key. Nothing ever lifts the bridge. All those
engines do is tilt it a bit – what’s on one side of the pivot point is exactly
balanced by what’s on the other side. This means that very little energy is
required to move it – only just enough to overcome the friction of the
bearings. Gravity effectively goes away as a problem, because the pull
downward on one side is exactly balanced by the pull downward on the
other side. We can’t beat gravity, but we can use it against itself. And you
can make a see-saw as large as you like, as the Victorians recognized.
After the tour, I walked a little distance along the river and then turned
round to look back at the bridge. My view of it had completely changed,
and I absolutely loved seeing it differently. The Victorians didn’t have
electricity on tap, computers to control things, or swanky new materials like
plastic and reinforced concrete. But they were masters of simple physical
principles, and the simplicity of the bridge really gets to me. It’s precisely
because it’s based on something so simple that it’s still working after 120
years, with almost no alteration. The gothic revival architecture (which is
apparently the technical term for ‘fairy-castle style’) is just wallpaper
covering up a giant see-saw. If they ever build one like this again, I hope
that they make some of it transparent, so that everyone can see the genius of
it.
This trick for reducing the problems of gravity can be seen all over the
place. For example, imagine a pivot point 4 metres above ground with two
6-metre-long halves of a see-saw balancing each other out on either side.
This isn’t a bridge. This is a Tyrannosaurus rex, the iconic carnivore of the
Cretaceous world. Two chunky legs hold it up, and the pivot point is at its
hips. The reason it didn’t spend its life falling flat on its face is that the
large heavy head with its terrifying teeth was balanced out by a long,
muscular tail. But life as a walking see-saw comes with a problem. Even a
very determined T. rex would sometimes have needed to change direction,
and they were lousy at it. It’s been estimated that it could take one or two
seconds for them to turn through 45 degrees, making them a bit more
cumbersome than the clever, agile T. rex of Jurassic Park. What could limit
a huge, strong dinosaur in such a way? Physics to the rescue ...
Spinning ice-skaters bring many things to the world – aesthetics, grace
and astonishment at what the human body is capable of. But if you hang
around physicists explaining things often enough, you could be forgiven for
thinking that an ice-skater’s sole contribution is to show everyone that
sticking your arms out makes you spin more slowly than when they’re
tucked in. They’re a useful example because ice is more or less free of
friction, and so once somebody is spinning, they have a fixed ‘amount’ of
spin. There’s nothing to slow them down. So it’s really interesting that
when they change their shape, they also change their speed. It turns out that
when things are further from the axis of spin, they have to travel further on
fn5
each turn, and so effectively take up more of the available ‘spin’. If you
stick your arms out, they’re further away from the axis, and so the speed of
rotation goes down to compensate. And this is basically the problem that
the T. rex had. It could only generate so much turning force (‘torque’) with
its legs, and because its huge head and tail were sticking out just like very
fat, heavy, scaly versions of the skater’s arms, it could only turn slowly.
Any small agile mammal (for example, one of our very distant ancestors)
would be a lot safer once it had worked that out.
The same thought also explains why we put our arms out sideways when
we think we’re about to fall over. If I’m standing upright and I start to fall
to my right, I’m rotating around my ankles. If I stick my arms out or up
before I start to fall, the same tipping force won’t move me as far, and so
I’ve got more time to make adjustments to stay standing up. That’s why
gymnasts on the balance beam almost always have their arms out
horizontally – it’s increasing their moment of inertia, so they’ve got more
time to correct their posture before they fall too far. Having your arms out
also lets you rotate yourself by lifting or dropping them, and that helps your
balance, too.
In 1876, Maria Spelterina became the only woman ever to cross Niagara
Falls on a tightrope. There’s a photograph of her halfway across, serenely
balanced and with peach baskets attached to her feet (to increase the
drama). But the most obvious prop in the photo is the long horizontal pole
she’s carrying, the best aid to balance. Arms will only reach out so far, but
this arm-substitute was a large part of the reason for Maria’s exquisite
fn6
control. If she started to lose her balance, it would only happen very
slowly, because the distance between the ends of the pole means that the
same torque has less effect. Maria was concerned with falling over to one
side, but the long pole would also have made it very difficult for her to twist
from left to right. And so it was with the T. rex. The same bit of physics that
was Maria’s best defence against falling 50 metres to certain death in the
churning water had also, 70 million years earlier, made it impossible for a T.
rex to change direction quickly.
Gravity pulling on solid objects is a familiar concept, mostly because
we’re solid objects that are pulled on. But around the solid objects in our
world, fluids are flowing – air and water are shifting around in response to
the forces acting on them. I think it’s a great tragedy that we can’t usually
see fluids shifting as clearly as we see leaves falling or bridges rising.
Liquids experience the same forces, but they aren’t limited to holding the
same shape, and so the world of fluid dynamics is beautiful: sweeping,
whirling, meandering, surprising and everywhere.
The lovely thing about bubbles is that they are everywhere. I think of
them as the unsung heroes of the physical world, forming and popping in
kettles and cakes, bioreactors and baths, doing all sorts of useful things
while often only fleetingly in existence. They’re such a familiar part of our
background that we often don’t really see them. A few years ago, I was
asking groups of five- to eight-year-old children where you might find
bubbles, and they all happily told me about fizzy drinks, baths and
aquariums. But the last group of the day were tired, and my cheerful
encouragement was met with a grumpy silence and blank stares. After a
long pause, and a lot of shuffling, one unimpressed six-year-old stuck his
hand up. ‘So,’ I said brightly, ‘where might you find bubbles?’ The boy
stared at me with a do-I-really-have-to glare, and announced loudly:
‘Cheese … and snot.’ I couldn’t fault his logic, although I had never
thought about either. It seemed likely that his experience of bubbly snot
outranked mine, anyway. But for at least one animal, bubbly snot is the key
to a whole lifestyle. Meet the purple marine snail, janthina janthina.
The snails that live in the sea generally scoot about on the ocean floor or
on rocks. If you were to prise one off its rock, carry it a little way up into
the water and let go, it would sink. The ancient Greek polymath
Archimedes (he of ‘Eureka’ fame) was the first to work out the principle
that determines when something is going to float and when it’s going to
sink. He was probably much more interested in ships, but the same
principle applies to snails and whales and anything else which is submerged
or semi-submerged in any fluid. Archimedes worked out that there is
effectively a competition between the submerged object (our snail) and the
water that would be there if the snail wasn’t. Both the snail and the water
around it are being pulled downwards towards the centre of the Earth.
Because the water is a fluid, things can move around in it very easily. The
gravitational pull on an object is directly proportional to its mass – double
the mass of your snail and you double the pull on it. But the water around it
is also being pulled downwards, and if the water is pulled down more, the
snail will have to float upwards so there’s more room for water underneath.
Archimedes’ Principle, stated for our hapless mollusc, is that there is an
upward push on the snail equal to the downward gravitational pull of the
water that should have been in that space. This is the buoyancy force, and
every submerged object experiences it. In practical terms, it means that if
the snail has a greater mass than the water that would fill a snail-shaped
hole, it will win the gravitational battle and sink. If the snail has less mass
(and is therefore less dense), the water will win the battle to be pulled
downwards and the snail will float. Most marine snails are more dense than
sea water overall, and so they sink.
For most of their history, marine snails just sank and that’s the way it
was. But at some point in the past, a ‘normal’ marine snail had a bad day
and got an air bubble trapped in its egg cases. The clever bit about
buoyancy is that it’s only the average density of the overall object that
counts. You don’t have to change the mass of the object. You can just
change the space it takes up – and air bubbles take up lots of space. One
day, a bigger air bubble was trapped, the balance tipped the wrong way, and
the first marine snail took flight through the water and drifted up towards
the sunlight. The door to the vast larder at the sea surface had been opened
… but only for a snail that could puff itself up; and so evolution got to
work.
Today, Janthina janthina, the descendant of the first snails that got lost in
space, is common in the warmer oceans of the world. Now bright purple,
the snails secrete mucus, the same sort of slime you see on garden stones in
the early morning, and use their muscular foot to fold the mucus and trap air
from the atmosphere. They build themselves a large bubble raft, often
bigger than themselves, to ensure that their total density is always less than
the sea water they’re in. So they always float, upside down (bubble raft up,
shell beneath), preying on passing jellyfish. If you see a purple snail shell
on a beach, it’s probably from one of these.
Buoyancy can be a very useful and quick indicator of what’s inside a
sealed object. For example, if you take identically sized cans of a fizzy
drink, one diet version and one with a full load of sugar, you’ll see that the
diet can floats in fresh water and the other one sinks. The cans have exactly
the same volume, so the difference is all inside, and it’s all that dense sugar.
A standard 330 ml can of pop has 35–50 g of sugar inside it, and that extra
mass counts, making the can overall more dense than water. That means it
beats the water in the battle with gravity, and so it sinks. The mass of
sweetener in diet pop is minuscule, so that can is basically just filled with
water and air, and it floats. A slightly more useful example is a raw egg.
Fresh eggs are more dense than water, so they sink and lie flat in cold water.
But if they’ve been sitting in your fridge for a few days, they’ll have been
gradually drying out, and as the water sneaks out of the shell, air molecules
sneak into an air sac at the rounded end to fill the gap. An egg that’s about a
week old will sink but stand up on the pointy end (so that the additional air
is closer to the surface). And if the egg floats completely, it’s been around
for a bit too long – have something else for breakfast!
Of course, if you can control the amount of air you’re carrying with you,
and how much space it takes up, you can choose whether to float or sink.
When I first started studying bubbles, I remember finding a paper written in
1962 that stated authoritatively: ‘Bubbles are created, not only by breaking
waves, but also by decaying matter, fish belchings and methane from the
seafloor.’ Fish belchings? It seemed clear to me that this had been written
from the blinkered comfort of a large leather armchair, probably in the
depths of a London club and much closer to the port decanter than the real
world. I thought it was a very funny misconception, and said so. Three
years later, while working underwater in Curaçao, I turned around to see a
massive tarpon (about 1.5 metres long), swimming just over my shoulder
and belching copiously from its gills. That was me told … In fact, many
bony fish do have an air pocket known as a swim bladder to help them
control their buoyancy. If you can keep your density exactly the same as
your surroundings, you’re in balance and you’ll stay put. The tarpon’s swim
bladders are unusual (tarpon are a rare example of a fish that can breathe air
directly as well as extracting oxygen via its gills), but I had to admit that
fish do belch. I still maintain that it’s not a significant contributor to the
fn7
number of ocean bubbles, though.
The consequences of gravity depend on what is being pulled on. Tower
Bridge is a solid object, and so gravity can change the position of the bridge
but not its shape. The snail is also a solid object, and it’s moving through
ocean water that can flow around it to adjust. But gases can flow too (their
ability to flow is why both liquids and gases are called fluids). Solid objects
can also move through gases as they follow the pull of gravity: a helium
party balloon and a Zeppelin rise for the same reason the bubbly snotty
snail does. They are fighting the battle of gravity with the fluid around them
and losing.
So the presence of a constant gravitational force can make things
unstable, which generally means that there are unbalanced forces and things
will shuffle around until balance is restored. If a solid object becomes
unstable it flips over or falls down, and any liquid or gas surrounding it will
just flow around to make room for the movement. But what happens when
the thing that is unstable isn’t a single solid object like a balloon, but the
fluid itself?
Strike a match, light the wick of a candle, and a fountain of bright, hot
gas is switched on. Candle flames have cast a warm glow over scribes,
conspirators, schoolchildren and lovers for centuries. Wax is a soft,
unassuming fuel, and that makes its transformation all the more surprising.
But each one of these familiar yellow flames is a compact and powerful
furnace, fierce enough to smash apart molecules and forge tiny diamonds.
And each one is sculpted by gravity.
As you light the wick, the heat of the match melts both the wax in the
wick and also the wax close to it, and the first transformation is to liquid.
Paraffin waxes are hydrocarbons, long chain molecules with a carbon
backbone that’s twenty to thirty atoms long. The heat doesn’t just give them
the energy to slither over each other like a pile of snakes (which is what
liquid wax would look like if you could see the molecules). Some will get
enough energy to escape completely, drifting out and away from the wick.
A column of hot gaseous fuel forms – so hot that it pushes hard on the
surrounding air, taking up a huge amount of space for a relatively small
number of molecules. The molecules are the same, so the gravitational pull
on them is the same in total. But now they’re taking up more space, so the
gravitational pull for every cubic centimetre has gone down.
Just like a bubbly snotty snail in the ocean, this hot gas must rise because
there is cool dense air trying to slide underneath it. The hot air is pushed up
an invisible chimney, mixing with oxygen along the way. Even before
you’ve moved the match away from the candle, the fuel is breaking apart
and burning in the oxygen, making the gas even hotter. These are the blue
parts of the flame, and they reach a staggering 1,400°C. The fountain that
you’ve started intensifies, as the hot air is pushed upwards ever more
quickly. It’s fed from beneath because the wick is just a long thin sponge,
soaking up other wax molecules that have been melted by the furnace.
But the fuel doesn’t burn perfectly. If it did, the flame would stay blue
and candles would be useless as light sources. As the long chain molecules
are snapped and bullied by the heat, some of the detritus remains unburned
because there isn’t enough oxygen to go round. Soot, tiny specks of carbon,
is carried upwards by the flow and heated. This is the source of the
comforting yellow light that glows as the soot reaches 1,000°C. The light of
a candle is only a byproduct of the fierce heat, and this light is just the glow
of a miniature hot coal in a fire. These tiny carbon particles are so hot that
spare energy in the form of light is pouring off them, and out into the
surroundings. It’s been discovered that the maelstrom of a candle doesn’t
just produce soot in the form of graphite (the stuff we think of as black
carbon). It also produces tiny amounts of the more exotic structures that can
be formed when carbon atoms join together: buckyballs, carbon nanotubes,
and specks of diamond. It’s been estimated that the average candle flame
produces 1.5 million nanodiamonds each second.
A candle is the perfect example of what happens when a fluid needs to
rearrange itself to satisfy the pull of gravity. Hot burning fuel rises very
quickly as cool air pushes underneath, forming a continuous convection
current. If you blow out the candle, the column of gaseous fuel will keep
streaming upwards above the candle for a few seconds, and if you lower a
match down from above, you’ll see the flame jump to the wick as the
fn8
column is re-lit.
Convection currents like this help move energy around and share it out
wherever a fluid is heated from below. They are why fish-tank heaters,
underfloor heating and saucepans on the stove are all so effective – none of
those things would work nearly as well without gravity. When we say ‘heat
rises’, that’s not quite true. It’s more that ‘cooler fluid sinks as it wins the
gravitational battle’. But no one thanks you for pointing that out.
Buoyancy doesn’t just matter for hot air balloons, snails and romantic
candlelit dinners. The oceans, the vast engines of our planet, take their
marching orders from gravity just like everything else. The depths are not
still. Water that hasn’t seen sunlight for centuries is flowing across and
around the planet, on a long, slow journey back to the daylight. But before
looking down into the depths, look up. Next time you see a tiny moving
glint high in the sky on a clear day, a passenger jet at cruising altitude,
allow yourself to appreciate just how high up it is: about 10 km. Then
imagine yourself standing at the deepest part of the ocean floor, the bottom
of the Marianas Trench. The ocean surface would be the same distance
fn9
above you as that plane is. Even the average depth of the oceans is 4 km, a
little under half the distance to that plane. The ocean covers 70 per cent of
the surface of the Earth. There is a lot of water out there.
And hidden in those dark depths is a familiar pattern. The same
mechanism causing raisins to dance in lemonade is also driving the vast
oceans of the Earth on their slow journey around the planet. The scale is
different and the consequences are more important, but the principle is
exactly the same. The blue of our blue planet is in motion.
But why would it move? The oceans have had millions of years to adjust
to their situation. Surely they’d have reached wherever they’re going by
now? Two things keep stirring the pot: heat and salinity. They matter
because they affect density, and a fluid with areas of different density will
flow to adjust as the battle of gravity plays out. We all know that the ocean
is salty, but I am still staggered every time I really think about just how
much salt is out there. To make a standard full household bath as salty as
the ocean, you need to add about 10 kg of salt, a large bucketful. A whole
bucket, just for one bath! It’s not the same everywhere in the ocean – the
salinity ranges from about 3.1 per cent to about 3.8 per cent, and although
that difference sounds tiny, it matters. Just as putting sugar in a fizzy drink
makes it more dense, the huge amount of salt makes sea water more dense
than fresh water. Colder water is more dense than warm water, and the
oceans range from about 0°C close to the poles to 30°C close to the equator.
So cold, salty water will sink and warmer, fresher water will rise. And that
simple principle takes sea water on a continual journey around the planet. It
may be thousands of years before one bit of water returns to the same part
of the ocean again.
fn10
In the North Atlantic, water is cooling as the wind steals heat away.
Where the sea surface freezes to form sea ice, the new ice is mostly just
water; the salt gets left behind. Together, those processes make the sea
water colder, saltier and denser, and so it starts to sink, pushing the less
dense water out of its way as it answers the call of gravity and finds its way
to the bottom of the sea. As it slithers slowly along the sea floor, it’s
channelled by valleys and blocked by ridges, just like a river. From the
North Atlantic, it flows southwards along the bottom of the ocean at a few
centimetres per second, and after a thousand years it reaches its first
obstacle: Antarctica. Unable to creep further south, it turns to the east as it
meets the Southern Ocean. This ocean, the great watery roundabout at the
bottom of the planet, links all the sea water on the planet because on its way
round the white continent, it merges with the lower edge of the Atlantic, the
Indian and the Pacific oceans. The vast, slow flow of water from the North
Atlantic creeps around Antarctica until it turns northwards again,
journeying far into either the Indian or the Pacific Ocean. Gradual mixing
with the water around it reduces its density, and it eventually finds its way
back to the surface, after perhaps 1,600 years without a single sunbeam
passing through it. There rain, runoff from rivers and ice-melt dilute the salt
again, while wind-driven currents push it along on the rest of its journey
until it finds its way back into the North Atlantic, perhaps about to repeat
the cycle. It’s called the thermohaline circulation: ‘thermo’ for heat, and
‘haline’ because of the salt. This ocean overturning is also sometimes
referred to as the Ocean Conveyor Belt, and although this conveys a slightly
simplistic picture, these flows do girdle the planet and they are driven by
gravity. The wind-driven surface currents have carried explorers and traders
for centuries. But the ocean conveyor system as a whole carries a cargo of
at least equal importance to our civilization: heat.
More heat from the Sun is absorbed at the equator than in any other area
on the planet, both because the Sun is higher in the sky there, and because
the planet is widest there and so there’s a large area for absorption. Heating
up water even a tiny bit takes a lot of energy, so warm oceans are like a
giant battery for solar energy. The shifting ocean is redistributing that
energy around the planet, and the thermohaline circulation is the hidden
mechanism behind our weather patterns. Much of our thin, fickle
atmosphere whooshes about on top of a steady heat reservoir that constantly
provides energy and moderates extremes.
The atmosphere gets all the glory, but the oceans are the power behind
the throne. Next time you look at a globe, or a satellite picture of Earth,
don’t think of the oceans as the empty blue bits between all the interesting
continents. Imagine the tug of gravity on those giant, slow currents, and see
the blue bits for what they are: the biggest engine on the planet.
3
Small is beautiful
Surface tension and viscosity
Before you can worry about things that are too small to see, you have to
know that they’re there. Humanity faced a catch-22 here – if you don’t
know there’s anything there, why would you go looking for it? But all of
that changed in 1665 with the publication of one book, the first scientific
bestseller: Robert Hooke’s Micrographia.
Robert Hooke was the Curator of Experiments at the Royal Society, and
so he was a generalist, free to roam among the scientific toys of the day.
Micrographia was a showcase for the microscope, designed to impress the
reader with the potential of this novel device. The timing was perfect. This
was an era of great experimentation and rapid advances in scientific
understanding. Lenses had been pootling around the edges of human
civilization for a few centuries, mostly unappreciated and seen as novelties
rather than serious tools for science. But with Micrographia, their moment
had arrived.
The wonderful thing about this book is that although it wears the robes of
respectability and authority, as befits a publication of the Royal Society, it’s
unashamedly the product of a scientist at play. It’s full of detailed
descriptions and beautiful illustrations, expensively produced and carefully
presented. But underneath all that, Robert Hooke was basically doing what
every child does when given a microscope for the first time. He just went
round looking at everything. There are stunningly detailed pictures of
razorblades and nettle stings, grains of sand and burnt vegetables, hair and
sparks and fish and bookworms and silk. The level of detail revealed in this
tiny world was shocking. Who knew that a fly’s eye was so beautiful? In
spite of the careful observations, Hooke didn’t make any claims to in-depth
study. In the section on ‘gravel in urine’ (the crystals commonly observed
on the insides of urinals), he speculates on a way of curing this painful
affliction and then happily leaves the hard work of actually solving the
problem to someone else:
It may therefore, perhaps, be worthy some Physicians enquiry, whether there may not be
something mixt with the Urine in which the Gravel or Stone lies, which may again make it
dissolve it, the first of which seems by it’s regular Figures to have been sometimes Crystalliz’d
out of it … But leaving these inquiries to Physicians or Chymists, to whom it does more
properly belong, I shall proceed.
And proceed he does, dancing through mould and feathers and seaweed, the
teeth of a snail and the sting of a bee. On the way, he coins the word ‘cell’
as a description of the units that made up cork bark, marking the start of
biology as a distinct discipline.
Hooke hadn’t just shown the way to the world of the small; he’d thrown
open the doors and invited everyone in for a party. Micrographia inspired
some of the most famous microscopists of the following centuries, and also
whetted the scientific appetite of fashionable London. And the fascination
came from the fact that this fabulous bounty had been there all along. The
annoying black speck buzzing around rotting meat was now revealed as a
minute monster with hairy legs, bulbous eyes, bristles and shiny armour. It
was a shocking discovery. By then, great voyages had crossed the world,
new lands and new people had been discovered, and there was great
excitement about what was to be found in far-away places. It hadn’t really
occurred to anyone that navel-gazing might have been severely underrated,
and that even belly-button fluff might have much to say about the world.
And once you’d got over the shock of the flea’s hairy legs, you could see
how they worked. The world down there was mechanical, it was
comprehensible, and the microscope made sense of things that humans had
noticed for years but hadn’t been able to explain.
But even that was just the start of the voyage into the world of the small.
Over two centuries more would pass before the existence of atoms was
confirmed, each one so tiny that you’d need 100,000 of them to make a line
as long as a single one of the cork cells. As the famous physicist Richard
Feynman was to point out many years later, there’s plenty of room at the
bottom. We humans are just lumbering about in the middle of the size
scales, oblivious to the minuscule structures that our world is built of and
built on. But 350 years after the publication of Hooke’s Micrographia,
things are changing. We can do more than just peer into that world like a
child peering into a protective glass museum case, forbidden to touch. Now
we’re learning to manipulate atoms and molecules on that scale; the glass is
off the case, and we can join in. ‘Nano’ is in fashion.
A major part of what makes the world of the tiny both fascinating and
extremely useful is that things work differently at that level. Something
that’s impossible for a human might be an essential life skill for a flea. All
the same laws of physics apply; the flea exists in the same physical universe
fn2
as you and me. But different forces take priority. Up here in our world,
there are two dominant influences. The first is gravity, pulling us all
downwards. The second is inertia; because we’re so big, it takes a lot of
force to get us moving or to slow us down. But as you get smaller,
gravitational pull and inertia also get smaller. And then they find
themselves in competition with the other weaker forces that were there all
along, but insignificant. There’s surface tension, the force shifting the
coffee granules about as the coffee puddle dries. And then there’s viscosity.
Viscosity in the world of the small is why we don’t get a nice layer of cream
on top of the milk any more.
It was always the gold- and silver-topped milk bottles they went for. If
you were early enough, and careful when you opened the front door, you’d
catch them at it. Bright-eyed perky little birds, perched on the top of the
bottle, snatching hasty sips of cream from the hole they’d pecked in the thin
aluminium bottle top while keeping a beady eye on the world around them.
As soon as they knew they were rumbled, they were off, probably to try
their luck at the neighbour’s doorstep. For about fifty years in this country,
blue tits were masters at stealing cream. They learned from each other that
right below the flimsy lid there was a rich fatty treasure, and that
knowledge spread throughout the UK blue-tit population. Other bird species
didn’t seem to cotton on to this trick, but the blue tits were waiting for the
milkman every morning. And then the game came to an end quite suddenly,
not just because of plastic milk bottles but because of something more
fundamental. For as long as humans have been milking cows, the cream has
risen to the top. But these days, it doesn’t.
The bottle that the hungry blue tit was hopping about on contained a
mixture of all sorts of goodies. Most of milk (nearly 90 per cent) is water,
but floating around in that are sugars (that’s the lactose that some people
can’t tolerate), protein molecules assembled into minuscule round cages,
and bigger globules of fat. All of this is jumbled up together, but if you
leave it to sit for a while, a pattern emerges. The fat globules in milk are
tiny – between 1 and 10 microns in size, which means you could fit
somewhere between 100 and 1,000 of them in a line between the millimetre
markers on a ruler. And those tiny blobs are less dense than the water
around them. There’s less ‘stuff’ in the same volume of space. So as they’re
being jostled about with everything else, there’s a tiny difference in where
they go. Gravity is pulling the water around them downwards a tiny bit
harder than it’s pulling the fat globules, and the fat is very gently squeezed
upwards. That means that the fat is ever so slightly buoyant, and will very
slowly rise up through the milk.
The question is: how fast will it rise? And here’s where the viscosity of
the water starts to matter. Viscosity is just a measure of how hard it is for
one layer of a fluid to slide over another layer. Imagine stirring a cup of tea.
As the spoon goes round, the liquid around the spoon has to move, flowing
past other liquid next to it. Water isn’t very viscous, so it’s very easy for
those layers to move past each other. But then think about stirring a cup of
syrup. Each sugar molecule is holding on to the ones around it very firmly.
To move these molecules past each other, you’ve got to break those bonds
before the molecules can move on. So it’s hard work to shunt the fluid
about, and we say that the syrup is viscous.
In the milk, the fat globules are pushed upwards because they’re buoyant.
But if they want actually to move upwards, they have to shove the liquid
around them out of the way. As part of that pushing process, the nearby
liquid has to slide over itself, so its viscosity matters. The more viscous it is,
the more resistance there is to the fat globules rising.
Right under the blue tit’s feet, this battle is going on. Each fat globule is
being pushed upwards by its buoyancy, but it experiences a drag force
because of the liquid around it having to move to let it pass. And the same
forces acting on the same sort of fat globule come to a different
compromise for different globule sizes. The drag has a much greater effect
when you’re small, because you have a large surface area relative to your
mass. You’ve only got a small buoyancy to use to shove quite a lot of the
surrounding stuff out of the way. So even though the smaller fat globule is
in exactly the same liquid, it rises more slowly than a bigger one. In the
world of the small, viscosity generally trumps gravity. Things move slowly.
And your exact size matters a lot.
In the milk, the larger fat globules rise faster, bump into some smaller,
slower ones and stick to them, forming clusters. These clusters experience
less drag for their buoyancy because they’re even bigger than individual
globules, so they rise even faster. The blue tit just has to sit and wait at the
top, and breakfast will arrive at its feet.
fn3
And then came homogenization. Milk manufacturers worked out that if
they squeezed the milk at very high pressure through very tiny tubes, they
could break up the fat globules and reduce their diameter by about a factor
of five. That reduces the mass of each one by a factor of 125. Now the
weedy upward buoyant push on each globule provided by gravity is
completely overwhelmed by viscous forces. The homogenized fat globules
fn4
rise so slowly that they might as well not bother. Just making them smaller
shifts the battle into different terrain where viscosity can score a clear
victory. Cream won’t rise to the top any more. The blue tits had to find
another source of breakfast.
fn5
So the forces are the same, but the hierarchy is different. Both gases and
liquids have viscosity – even though gas molecules don’t stick to each other
like the ones in liquids, they jostle each other a lot, and the giant game of
bumper cars has the same effect. This is why an insect and a cannon ball
don’t fall at the same speed unless you take all the air away and drop them
in a vacuum. Air viscosity matters a lot for the insect, and hardly at all for
the cannon ball. If you take the air away, gravity is the only force that
matters in both cases. And a tiny insect trying to fly in air uses the same
techniques that we use to swim in water. Viscosity dominates their
surroundings, just as it does ours in the pool. The smallest insects are
swimming through air much more than they’re flying through it.
Homogenized milk demonstrates the principle, but its application goes
far beyond the doorstep. Next time you sneeze, you might want to think
about the size of the droplets you’re spraying round the room. What stops
cream going up also stops disease coming down.
Tuberculosis has been with humans for millennia. The earliest record of
it is in ancient Egyptian mummies from 2400 BC; Hippocrates knew it as
‘phthisis’ in 240 BC, and European royalty were called upon to cure the
‘king’s evil’ in medieval times. As the Industrial Revolution drove people
to live in towns, ‘consumption’, the disease of the urban poor, was
responsible for a quarter of all deaths in England and Wales in the 1840s.
But it wasn’t until 1882 that the culprit was found, a tiny bacterium called
Mycobacterium tuberculosis. Charles Dickens described the common sight
of consumptives coughing, but he couldn’t write about one of the most
important aspects of the malady, because he couldn’t see it. Tuberculosis is
an airborne disease. Carried out of the lungs with each cough are thousands
of fluid droplets, plumes of minuscule crusaders. Some of them will contain
the tiny rod-shaped TB bacteria, each only three-thousandths of a
millimetre long. The fluid droplets themselves start off fairly big, perhaps a
few tenths of a millimetre. These droplets are being pulled downwards by
gravity and once they hit the floor, at least they’re not going anywhere else.
But it doesn’t happen quickly, because it’s not just liquids that are viscous.
Air is too – it has to be pushed out of the way as things move through it. As
the droplets drift downwards, they are bumped and jostled by air molecules
that slow their descent. Just as the cream rises slowly through viscous milk
to the top of the bottle, these droplets are on course to slide through the
viscous air to reach the floor.
Except they don’t. Most of that droplet is water, and in the first few
seconds in the outside air, that water evaporates. What was a droplet big
enough for gravity to pull it through the viscous air now becomes a mere
speck, a shadow of its former self. If it was originally a droplet of spit with
a tuberculosis bacterium floating about in it, it’s now a tuberculosis
bacterium neatly packaged up in some leftover organic crud. The
gravitational pull on this new parcel is no match for the buffeting of the air.
Wherever the air goes, the bacterium goes. Like the miniaturized fat
droplets in today’s homogenized milk, it’s just a passenger. And if it lands
in a person with a weak immune system, it might start a new colony,
growing slowly until new bacteria are ready to be coughed out all over
again.
Tuberculosis is treatable if the right drugs are available. That’s why it has
mostly disappeared from the western world. But at the time of writing, TB
is still the second greatest killer of our species after HIV/AIDS, and it’s a
gigantic problem in the developing world. Nine million people developed
TB in 2013, and 1.5 million of them died. The bacterium has changed in
response to antibiotics, becoming resistant to so many waves of drugs that
it’s obvious it can’t be eradicated using medicine alone. The number of
multi-drug-resistant strains of TB is on the increase. Outbreaks are popping
up in hospitals and schools. So recently the focus has shifted to those tiny
droplets. Rather than cure TB once you have it, how about changing your
buildings to prevent the spread of those disease-laden plumes so it never
gets to you in the first place?
Professor Cath Noakes works in civil engineering at the University of
Leeds, and she is one of the researchers chipping away at this particular
coalface. Cath is very enthusiastic about the potential for relatively simple
solutions to emerge from a sophisticated understanding of tiny floating
particles. Engineers like her are now learning how these tiny vehicles for
disease travel, and it turns out this has very little to do with what’s in them
and how long they’ve been there. It has everything to do with the battle of
forces on the particle, and the battle lines are drawn by the particle size. It’s
been discovered that even the larger droplets can travel further than anyone
fn6
had thought, because turbulence in the air can keep them aloft. The tiniest
ones can stay in the air for days, although ultraviolet and blue light damage
them. If you know where your particle sits on the size scale, you can work
out where it’s going to go. So, if you are designing a ventilation system for
a hospital, it’s becoming possible to plan to remove or contain specific
particle sizes, and therefore control the spread of disease. Cath tells me that
each airborne disease may require a different plan of attack, depending on
how much of it you need to get sick (in the case of measles, very little) and
where in your body the disease settles (the TB bacterium has different
effects in your lungs and your windpipe). These studies are still in their
early days, but they’re advancing very quickly.
Humans have been at the mercy of TB for generations, but now we can
visualize its spread, and that gives us the chance to control it. Where our
ancestors saw only a foul room of sickness, awash with mysterious
miasmas, we now understand the subtle swirling of the air around each
patient, the sorting and shunting of disease particles, and how the
consequences take effect. The outcomes of this research will be
incorporated into the hospital designs of the future, and many lives will be
saved by engineering on the macro scale to influence particles on the micro
scale.
Viscosity matters when something small is moving through a single fluid
– fat globules rising through milk or a tiny virus falling through the air.
Surface tension, its partner in the world of the small, matters at the place
where two different fluids touch. For us, that’s usually where air touches
water, and everyone’s favourite example of air mixing with water is a
fn7
bubble. So let’s start with a bubble bath.
The sound of a bathtub filling up is distinctive and jolly. It announces the
imminent reward after a hard day, a soak to recover from a particularly
tough badminton match or just a bit of pampering. But the moment you
pour in a bit of bubble bath, the sound changes. The deep rumble gets softer
and quieter as the foam builds up, and the place where the water stops and
the air starts gets harder to identify. Pockets of air are trapped inside watery
cages, and all it took was a tiny amount of stuff from a bottle.
It was a group of European scientists in the late nineteenth century who
picked apart the puzzle of surface tension. The Victorians loved bubbles.
Soap production expanded dramatically between 1800 and 1900, and the
white suds washed the workers of the Industrial Revolution. Bubbles
provided the Victorians with good fodder for moralizing; they were the
perfect symbol of pure cleanliness and innocence. And they were also a
nice example of classical physics at work, just a few years before Special
Relativity and quantum mechanics came along and poked a sharp pin in the
ballooning idea of a neat, tidy and well-behaved universe. But even so, the
serious men with top hats and beards didn’t work out the secrets of bubble
science all by themselves. Bubbles were so universal that anyone could
have a go. Enter Agnes Pockels, often described as a mere ‘German
housewife’, but really a sharp-minded critical thinker, who used the limited
materials available to her and a decent dose of ingenuity to examine surface
tension for herself.
Born in 1862 in Venice, Agnes was of a generation very firmly convinced
that the woman’s place was in the home. So that’s where she stayed while
her brother went off to university to study. But she learned advanced
physics from the material he sent to her, carried out her own experiments at
home, and generally kept up with what was going on in the academic world.
When she heard that the famous British physicist Lord Rayleigh was
starting to take an interest in surface tension, something she had done many
experiments on, she wrote to him. He was so impressed with the letter
describing her results that he sent it to be published in the journal Nature so
that it could be seen by all the greatest scientific thinkers of the day.
Agnes had done something very simple and very clever. She had
suspended a small metal disc (something about the size of a button) on the
end of a string and let it sit on the surface of the water. Then she had
measured how much force it took to pull it away from the surface. The
mystery was that the water held on to the disc; you had to pull harder to get
it off the water surface than you would have had to pull to pick it up off the
table. That pull from the water is what we call surface tension; so in
measuring the pull Agnes was measuring surface tension. She could then
investigate the surface of the water, even though the thin layer of molecules
responsible for the pull was far too small for her to see directly. We’ll see
how in a minute; but first, back to the bath.
A bath full of pure water is a jostling swarm of water molecules playing a
very crowded game of bumper cars. But one of the things that makes water
such a special liquid is that all those molecules are very strongly attracted to
all the other water molecules around them. Each one has a larger oxygen
atom and two small hydrogen atoms (that’s the two Hs and an O in H2O).
The oxygen sits in the middle with the two hydrogens stuck to it on either
side, making a shallow V-shape. But although the oxygen is very strongly
attracted to and bonded to its own two hydrogen atoms, it’s also flirting
with any others that happen to go by. So it’s constantly tugging on the
hydrogen from other water molecules. This is what holds water together.
It’s called hydrogen bonding, and it’s very strong. In the bath, water
molecules are constantly pulling on the other water molecules around them,
tugging the whole mass of water together.
The water molecules on the surface are a bit left out. They are being
pulled by the water molecules underneath them, but there’s nothing above
them to pull back. So they’re being pulled down and sideways but not up:
and the effect of this is to make the surface behave like an elastic sheet,
pulled tight over all the water molecules below the top layer, and pulling
itself inward so that it is as small as possible. This is surface tension.
As you run the tap, air gets carried downward into the bath, making
bubbles. But when these bubbles float up to the surface, they can’t last. The
round dome of the bubble is stretching the surface and the surface tension
isn’t strong enough to haul it back. So the bubbles burst.
One of the things that Agnes did was to set up her button so that it was
being pulled upwards, but not quite hard enough to pull it off the surface.
And then she touched the surface of the water nearby with a drop of
something like detergent. After a second or so, the button would pop off the
surface. The detergent had spread across the water, and it had reduced the
surface tension. All it takes to reduce the surface tension is to provide a thin
top layer so that the water molecules don’t have to be the ones right at the
surface.
When it’s finally time to add the bubble bath, it’s time to say goodbye to
a clean, flat, minimal surface. That dollop of scented gloop gets carried
down into the water and immediately does its best to hide at the edges. Each
molecule has one end that loves water and one end that hates water. If the
end that hates water can find some air, it will stay with it, but the water-
loving end isn’t giving in either. So at any place where water touches air, a
thin layer of bubble bath sits right at that surface. It’s just one molecule
thick, and each molecule is the same way round so that the water-loving
ends are all still submerged in water, and the water-hating ends are all still
in air. With this thin coating, a large surface isn’t a problem. The bubble
bath doesn’t provide the strong pull that water does, so the elastic sheet
effect becomes really weak. It’s time for a surface party, and that’s what the
foam is. By reducing the surface tension, bubble bath makes it easier for
bubbles to last because their large surface is much more stable.
It’s probably worth noting here that we associate the white foam with
things getting cleaner, but in modern detergents the best stuff for sticking to
the water surface and making the foam is not the best stuff for pulling dirt
and grease off clothes and plates. You can make a very good cleaning
detergent which hardly makes foam at all, and in fact the foam often gets in
the way. But the purveyors of cleaning products did such a good job of
convincing people that beautiful white foam was your guarantee of a
thorough cleaning that they’ve now backed themselves into a corner.
Foaming agents are now added to make sure there are bubbles, because
otherwise consumers complain.
Like viscosity, surface tension is something that we’re aware of up here
at our size scales, although it’s usually less important than gravity and
inertia. As you get smaller, surface tension pushes its way up the hierarchy
of forces. It explains why goggles fog up and how towels work. And the
real beauty of the world of the small is that you can contain many tiny
processes inside one giant object, and their effects add up. For example, it
turns out that surface tension, which only dominates in the tiniest of
situations, also makes possible the largest living things on our planet. But to
get there, we need to look at another aspect of surface tension. What
happens when the surface separating a gas and a liquid bumps up against a
solid?
My first open-water swim turned out not to be for the faint-hearted.
Fortunately, I didn’t know that beforehand so I couldn’t worry about it.
When I was working at the Scripps Institution of Oceanography in San
Diego, the big annual event for my swim team was from La Jolla beach to
Scripps pier and back, 4.5 km across a fairly deep marine canyon. I had
only ever swum properly in swimming pools, but I’m always up for trying
something new and I’d been swimming a lot, so I turned up and hoped I
didn’t look like too much of a rookie. The mass entry to the water was a bit
of a scrum, but it got better after that. The first part of the swim was across
the top of a stunning kelp forest, and it was almost like flying. The sun
glinted through the huge stalks of bull kelp just like it does in forests on
land, and then the kelp disappeared downward into the murky depths,
reminding me that there were many creatures swimming about down there
that I couldn’t see. Once we were past the kelp, the water got choppy, and I
had to pay much more attention to where we were going. And that was
getting harder. The pier was fuzzy on the horizon, and I couldn’t see
anything at all down below. After slightly too long, I realized that the
reason everything had disappeared was that my goggles had fogged up. Oh.
Inside my plastic goggles, sweat had evaporated from the warm skin
around my eyes. The harder I worked, the more evaporated. The air trapped
between me and my goggles was now a mini-sauna, hot and humid. But the
ocean around me was nice and cool, and so my goggles were being cooled
from the outside. When water molecules in the air bumped into the nice
cool plastic, they gave up their heat and condensed, becoming liquid again.
But that wasn’t the problem. The problem was that as all those water
molecules found each other on the inside of my goggles, they stuck
together, far more attracted to each other than they were to the plastic.
Surface tension was pulling them inwards, forcing them to collect in tiny
droplets so that there was as little surface as possible. Each droplet was tiny
– perhaps 10–50 microns across. So gravity was insignificant compared
with the surface forces sticking them to the plastic and there was no point in
waiting around in the hope of them falling off.
Each little droplet acted like a lens, bending and reflecting the light that
hit it. When I lifted my head to look for the pier, the light that had been
travelling straight to my eyes was messed up by the droplets. Like a tiny
house of mirrors, they had scrambled the image so that I was just looking at
vague grey fuzz. I stopped briefly to rinse out my goggles, and for a while I
had a crystal clear view of the pier again. But the fog came back. Rinse.
Fog. Rinse. Eventually I just stuck next to my swimming partner because
she had a bright red swimming cap, and the red made it through the silly
little water droplets.
When we reached the pier, we paused to check that everyone was OK.
With a bit of time to think, I finally remembered something I’d been taught
just a week or so before by a scuba diver. Spit in your goggles, and rub it
over the inside of the plastic. At the time, I’d made a face, but now I didn’t
want to go all the way back across the canyon blind, so I spat. And the
swim back was a completely different experience. That was partly because
my swim partner had decided that she was bored and wanted it all done
with, and I had to struggle to keep up. But it was mostly because I could see
– swimmers, kelp, the beach we were aiming for, the occasional curious
fish. Human saliva acts a bit like detergent: it reduces surface tension. My
goggles were still a mini-sauna and the water was still condensing, but
surface tension wasn’t strong enough to bunch it up in droplets. So it was
just spread out in a thin film covering the entire surface. Since there were
no watery lumps and bumps and boundaries, light could travel through in a
straight line, and I could see clearly. Back at the beach, I stumbled out of
the water euphoric, partly with relief for having finished the swim, and
partly with a new appreciation of what the underwater world had to offer.
This is one way to stop things fogging up: to spread a thin layer of
surfactant on the surface. Lots of things will do that job – saliva, shampoo,
shaving cream or expensive commercial antifog. If the surfactant is ready
and waiting, any water that condenses will immediately be coated in it. By
providing that coating, you are weakening the surface tension, and swaying
the battle of forces in each fog droplet so that the water covers the plastic
evenly. The water can stick to the whole surface of the goggles, as long as
there are no stronger forces to pull it away. Surface tension is the only other
force that stands a chance of competing, so when you weaken that, the
fn8
problem vanishes.
So one solution is to reduce the surface tension. But there is another
solution: increasing the attractiveness of the goggles. A droplet on its own
will suck itself up into a ball. If you put it on plastic or glass, it will sit up
high and barely touch since the water molecules will shuffle about until as
few as possible touch the plastic. But if you put the droplet on a solid
surface that attracts water molecules nearly as strongly as other water
molecules do, the water will snuggle up to that surface. Instead of a perky
near-spherical droplet, you get a flattened drip that feels the pull of the
surface as much as it feels the pull of its neighbours. These days, I buy
goggles that have a coating on the inside that attracts water – it’s called
hydrophilic. Water still condenses, but it spreads out along that surface,
attracted to the coating. Condensation in goggles is here to stay, but fogging
fn9
up is a thing of the past.
Weakening surface tension has its uses. But that pull between individual
water molecules is really strong. And the smaller the volume of water
you’re interested in, the more it matters. So what surface tension is really
useful for is plumbing on the tiniest scales. Down there, you don’t need
pumps and siphons and huge amounts of energy to shunt water around; you
just need to make things small enough for gravity to be irrelevant and let
surface tension get on with the hard work. Mopping up is boring, but the
world would be very different without it.
I’m a messy cook, reasonably competent, but far more interested in the
cooking process itself than the trail of devastation that I tend to leave
behind me. This makes me nervous when using other people’s kitchens.
Years ago in Poland, I set out to make apple pie for the international group
fn10
of volunteers I was working with at a school. It didn’t start well. The tall,
fierce school cook bellowed ‘NO!’ with some enthusiasm when I asked
whether I could use the kitchen, and it took me a few puzzled seconds to
remember that we’d been talking in Polish and ‘no’ is their word for ‘yeah’.
My Polish wasn’t very good, and I didn’t follow all of the details that came
next, but I took away the very strong message that the kitchen was to be left
clean. Very clean. No spilling anything. Definitely immaculate. So later that
evening, after she’d gone home and I’d assembled all the ingredients, of
course the first thing I did was to knock over a large and newly opened
carton of milk.
My first reaction was just to will the milk to vanish, so that the stern cook
would never know it had existed. Milk is slippery and sticky, can’t be
picked up or swept away, and this particular batch of it was advancing
along the kitchen floor at an alarming rate. But there is a tool for gathering
a liquid together, for putting it all back in one place. It’s called a towel.
As soon as the towel touched the milk, the liquid had a new set of forces
bossing it about. Towels are made of cotton, and cotton attracts water.
Down there on a tiny scale, water molecules were attaching themselves to
cotton fibres, and slowly creeping over the surfaces of each fibre. And
water molecules are so strongly attracted to each other that the first one to
touch the towel can’t crawl upwards by itself. It can only move up if it
brings the next water molecule with it. And that one has to bring the
following one. So water creeps up the cotton fibres, bringing everything
else in the milk with it. The forces sticking the water to the towel fibres are
so strong that the measly downward pull of gravity becomes totally
irrelevant. What went down happily goes back up.
But this is only half the story. The real genius of a towel is its fluffiness.
If a towel could only coat each of its fibres in a thin layer of water, it
wouldn’t be able to collect together much liquid at all. But the fluff gives
the towel lots of air pockets and narrow channels. Once water finds its way
into a narrow channel, it’s being pulled upwards on all sides, and the water
in the middle just gets dragged along as well. The narrower the channel is,
the more surface there is for each drip of water in the middle. Fluffy towels
have loads of surface, and very narrow gaps in between, so they can suck
up a lot of water.
As I watched the puddle of milk disappear into the towel, tiny water
molecules were crowded together, jostling inside the fluff. The ones at the
bottom were just going along with the crowd, sticking to the other water
molecules next to them. The ones touching the cotton were clinging on to
both the cotton and the water molecules on the other side, holding their
position. The ones touching dry towel were latching on to the new dry
cotton and, once they were attached, pulling others up behind them, filling
the gaps in the structure. The ones at the surface were tugging on the water
molecules directly below them, trying to surround themselves with as many
other water molecules as possible, and pulling water upwards in the
process. This is capillary action. Gravity was pulling down on all of that
milk, wherever it was in the fluff. But gravity couldn’t compete with the
forces holding the entire thing up, the ones at the top, where milk just
touched dry cotton inside millions of tiny air pockets. As I turned the towel
over and moved it around, different regions of the towel filled up, storing
water in the pockets.
Water will keep creeping upwards through the gaps, bringing other water
with it, until the sum of those tiny forces from a multitude of pockets is
finally balanced by the pull of the planet. This is why when you dip the
edge of a towel in water, the liquid will spread quickly upward for a few
centimetres and then stop. At that point, the weight of the water is exactly
balanced by the upward pull of the surface tension. The narrower the
channels in the fluff, the more surface there is to contribute surface tension,
and so the higher the waterline will be. Scale really matters here – if you
made fluff that had the same shape but was a hundred times bigger, it
wouldn’t be absorbent at all. But when you shrink the shape, you shift the
hierarchy of forces and up the water goes.
The best bit of the whole thing is that if you leave the towel out to dry,
water will evaporate from those pockets and disappear into thin air. As a
means of getting rid of a problem, that’s hard to beat; the towel collects and
fn11
holds on to the liquid until it floats off of its own accord.
The spill vanquished, I finished the apple pie and left the kitchen in a
suitably immaculate state. But I had one final problem stored up that no
amount of surface science could have helped me with. The whipped cream
that I served with the apple pie was thoroughly unpleasant, as the faces of
the pie consumers made pretty clear. It wasn’t the best way to learn the
Polish word for ‘sour’, the one that preceded the word ‘cream’ on the pot.
Still, you live and learn, and I won’t make that mistake again.
The reason why towels are made of cotton is that cotton is mostly
cellulose, long chains of sugars that water molecules stick to very easily.
Cotton wool, kitchen towel, cheap paper: all of these are absorbent because
they have a fluffy structure on a tiny scale, made out of water-loving
cellulose. The question is: what are the limits of this size-dependent
physics? If you make the channels as small as physically possible, what can
you do with them? It’s not just towels that suck water up tiny channels
made of cellulose. Nature got there long before us. The mightiest example
of what the physics of the small is capable of is also the largest living
organism on our planet: the giant redwood.
The forest is quiet and humid. It feels as though it must always have been
like this, as though change is rare here. The forest floor between the tree
trunks is carpeted with moss and ferns, and the only sounds are the songs of
unseen birds and the deep, unnerving creaking as the trees shift their
weight. Up above, blue skies are visible between the spindly green
branches, and down at my feet there is water everywhere: streams, patches
of damp soil, exploratory rivulets on their way down the valley. Every so
often as I walk, my subconscious kicks me into alertness, because there’s a
looming patch of darkness in the forest, something that doesn’t fit. But it
isn’t a predator. It’s a tree: one of the real giants, a thousand-year-old
colossus lurking among the youngsters, stamping its status on the forest
with its shadow.
The coastal redwood, Sequoia sempervirens, used to cover vast swathes
of this part of northern California. These days, those huge forests have been
reduced to a few small pockets, and I’m visiting one of the most well
known, the Redwood National Park in Humboldt County. These giants are
striking because each tree trunk is completely straight and vertical, reaching
only for the sky. The tallest known tree on the planet is here, and it’s a
fn12
staggering 116 metres high. On my hike, I frequently pass trees with trunk
diameters of 2 metres or more. Possibly most astonishing of all is that just
behind the deep ridges and wrinkles in the bark, these trees are still growing
new rings. They’re alive. The tiny evergreen leaves 100 metres above me
are capturing the Sun’s energy, storing it up, making the stuff from which
new tree is built.
But life demands water, and the water is down here, where I am. So all
around me in the forest, water is flowing upwards. And this flow has never
been interrupted, not once since each tree sprouted from its seed. Some of
these trees have been here since the Roman Empire fell. They were sitting
in the California fog when gunpowder was invented, when the Domesday
Book was written, when Genghis Khan was rampaging across Asia, when
Robert Hooke published Micrographia and when the Japanese bombed
Pearl Harbor. And not once, in all that time, has the water stopped flowing.
The reason we can be sure of this is that the whole mechanism relies on the
flow never stopping. There is no way to restart it. But this is very clever
plumbing, and the fabulous piece of living architecture that keeps it all
going only works because it’s just a few nanometres across.
The water travels in the xylem, a system of tiny cellulose pipes that reach
through the tree, stretching from the roots to the leaves. This is mostly what
‘wood’ is, although the innermost wood stops helping with the plumbing as
the tree gets bigger. Capillary action, the mechanism that made my towel
absorbent, is only strong enough to suck water upwards for a few metres in
the tree’s plumbing. That’s no use for a tall tree. The tree roots can also
generate their own pressure to push water up the pipes, but that too is only
enough to push the water a few more metres upwards. Most of the work
isn’t being done by pushing. The water is being pulled. The same system
operates in all trees, but the redwoods are the kings of it.
I sit on a fallen tree trunk, just next to one of the giants, and look up. A
hundred metres above my head, tiny leaves flutter in the breeze. To
photosynthesize, they need sunlight, carbon dioxide and water. The carbon
dioxide comes from the air, and it enters the tree through tiny pockets on the
underside of each leaf, the stomata. Part of the inner wall of each of those
pockets is a network of cellulose fibres, and in between the fibres are water-
filled channels. These are the top of the water pipes, after those pipes have
branched and branched again, reducing in size each time until they reach
the stomata. Here, where the water pipes finally touch the air, each one is
fn13
approximately 10 nanometres across. The water molecules stick firmly to
the cellulose sides of each channel, and the water surface curves down into
a nanoscale bowl-shape in between. Sunlight heats up the leaf and the air
inside it, and sometimes it gives enough energy to one of those surface
water molecules to pull it away from the mob below it. That evaporated
water molecule drifts out of the leaf into the air. But now the nanoscale
bowl is out of shape – it’s too deep. Surface tension is pulling it inwards,
pulling the water molecules closer together to reduce the surface area.
There are lots of new molecules that could fill the gap, but they are all
further back in the channel. So the water in the channel is pulled forwards
to replace the lost molecule. And then the water further back in the channel
has to shuffle along to replace that, and so on down the tree. Because the
channel is so tiny, the surface tension can exert an enormous pull on all the
water below it, enough (when you include the contribution of a million
other leaves) to pull the entire column of water up the tree. It’s a staggering
thought. Gravity is pulling the entire tree’s worth of water downwards, but
fn14
the combination of many tiny forces is winning the battle. And it’s not
just a battle against gravity; the upward forces are also defeating the friction
from the tube walls as water is squeezed through the tiny channels.
Pushing up from the forest floor around me are the real babies – trees that
are just a year old. Their water columns are just beginning to take shape. As
the new tree grows, the pipe system stretches but never breaks, and so the
top of the water column is always wetting the inside of the stomata. Water
is just pulled up towards the air as it keeps growing. The tree can’t refill the
pipe if it empties, but it can keep the pipe full as it grows. However tall the
tree gets, this water column must never be broken. The reason why the
tallest redwoods are near the coast is that the coastal fog helps their leaves
fn15
stay moist. Less water needs to reach the top from the roots, so the system
can be slower and the trees taller.
This process of water evaporating from the leaves of trees is called
transpiration, and it’s happening whenever you look at a tree in the sunlight.
These sleepy giant redwoods are actually massive water conduits, sucking it
from the forest floor, rerouting some of it for photosynthesis and then
letting the rest escape into the sky. It’s the same for every tree. Trees are a
vital part of the Earth’s ecosystems, and they wouldn’t be able to climb up
into the sky unless they could take water with them. And the beauty of it is
that they don’t need an engine or an active pump to do that. They just shrink
the problem, solve it using the rules of the small, and then repeat that
process so many million times that it becomes the physics of giants.
The tiny world where surface tension, capillary forces and viscosity
dominate over gravity and inertia has always been a part of our everyday
life. The mechanisms may be invisible, but the consequences are not. And
these days we’re not just spectators admiring the elegance and exoticism of
what happens down there. We’re beginning to be engineers, working within
it. There’s a word for the rapidly developing field of Lilliputian plumbing,
the manipulation and control of fluids flowing through narrow channels:
‘microfluidics’. It’s not a familiar word to most of us now, but it’s going to
have a huge impact on our lives in the future, especially when it comes to
medicine.
Today, people with diabetes can monitor their blood sugar using a simple
electronic device and a test strip. A tiny drop of blood touched to the test
strip will immediately whoosh into the absorbent material due to capillary
action. Tucked away in the tiny pores of the strip is an enzyme, glucose
oxidase, and when this reacts with blood sugar it produces an electrical
signal. The hand-held device measures that signal, and voilà! – an accurate
measure of blood sugar appears on the screen. It’s easy to see this as a
description of the obvious – paper soaks up a fluid so that it can be
measured. So what? But this is just a crude demonstration of the principle.
It gets a lot more sophisticated than that.
If you can move a fluid through tiny tubes and filters, gather it in
reservoirs, mix it with other chemicals along the way and see the results,
you have all the components of a chemistry lab. No need for glass test
tubes, hand-held pipettes and microscopes. This is the premise of the
growing ‘lab-on-a-chip’ industry, the development of tiny devices to carry
out medical tests. Nobody likes having a whole vial of blood extracted from
them, but a single drop isn’t too hard to part with. Smaller diagnostic
devices are often cheaper to make and easier to distribute. And you don’t
even need to make them from fancy modern materials like polymers or
semiconductors. Paper might do just fine.
A group of researchers at Harvard, led by Professor George Whitesides,
is on the case. They have engineered diagnostic test kits about the size of a
postage stamp, made of paper, but containing a maze of water-loving paper
channels with waxed water-hating walls. When you touch a drop of blood
or urine to the correct part of the paper, capillary action drags it through the
main channel, splits it up and reroutes it to lots of different test zones. Each
one contains the ingredients to do a different biological test, and each
fn16
reservoir will change colour depending on the test results. The researchers
suggest that someone a long way from a doctor could do the test locally,
take a picture of the result with their phone and e-mail it to a distant expert
who could interpret it. As ideas go, it’s brilliant. Paper is cheap, the device
doesn’t require power, it’s lightweight and a flame is all you need to dispose
of it safely. As with all these devices, it’s got a lot of checks and balances to
face before we know whether the simple-sounding idea can deal with the
real world. But it’s hard not to be convinced that one way or another,
devices like this will be a big part of medicine in the future.
The genius of all this is that when we look at a problem, we may be able
to choose to do the engineering on the size scale that makes the problem
easiest to solve. It’s like being able to choose which laws of physics you
want working for you. Small really is beautiful.
4
A moment in time
The march to equilibrium
One sunny day in Cambridge, I finally had to admit that I had been defeated
by a snail.
It’s not traditional to take up gardening during your final year as an
undergraduate, but the house I was sharing with three friends had a garden,
and the temptation was too much. In the odd spare hours between work and
sport that year, I enthusiastically chopped down the huge forest of nettles
that had taken over, and discovered buried treasure in the form of rhubarb
plants and rose bushes. My Dad laughed at me for planting potatoes
(‘typical Pole’, he said), but they were only part of my new vegetable patch.
Most exciting of all, there was a grimy greenhouse, filled with rubble and a
grapevine. Seedlings (leeks and beetroot, I think) could grow in shelter
before joining the vegetable patch in the spring. In late February I sowed
seeds in trays, and waited for new plants to grow.
After a while, it was noticeable that there weren’t any seedlings, but there
were a lot of snails. I’d arrive with my watering can to find a smug mollusc
parked in the middle of each tray, surrounded by bare soil and the
occasional green hint of a chewed-off shoot. Not to be defeated, I chucked
the snails outside, resowed the seeds, and placed the trays on top of bricks
to make it harder for the snails to crawl in. Two weeks later, the nascent
seedlings were gone, and there were more snails than ever. I tried a few
different approaches, none successful, until I had only one idea left. This
time, I took pairs of empty flower pots and balanced upside-down teatrays
on top, so they were like giant mushrooms with two stalks. I greased the
edges of each one, and put the seedling trays on top of the teatray
mushrooms. After replacing the compost, I sowed the last of the seeds,
crossed my fingers and went back to studying condensed matter physics.
The seedlings grew undisturbed for about three weeks. And then the
inevitable day came when I found one fat, happy snail where the seedlings
should have been. I remember standing in the greenhouse, analysing in
forensic detail the possible routes that this creature could have taken. There
were only two. Option one: it could have crawled up the inside walls of the
greenhouse and out across the underside of the roof, and then somehow
have dropped off at exactly the right location to land on a seed tray. This
seemed unlikely. Option two: it had crawled along the bench and up the
flower-pot sides, slimed its way upside down to the outer edge of the
teatray, crawled around the edge without falling off, and then trekked along
the top of the teatray to the seedlings. In either case, I had to admit that it
fn2
had probably earned the bounty. How could a snail do that? Both cases
involved it crawling upside down, while glued to the surface only by its
own mucus. If you watch a snail move, it’s different from a caterpillar – it
doesn’t lift itself off the surface as it goes. It’s just stuck to its slime, and yet
somehow it manages to shift itself along. But that slime is the snail’s secret
weapon, because it behaves just like ketchup.
If you watch a snail moving, you won’t see very much because the outer
rim of its foot is just moving at a constant slow speed. Everything on the
edges is happening slowly, so the mucus is like the stationary ketchup: thick
and gloopy and hard to move. But underneath, in the middle, muscular
waves are travelling from the back of the snail to the front. Each wave is
pushing forward on the mucus really hard, and it’s forcing the mucus to
shift very quickly. And just like the ketchup, the mucus is shear-thinning, so
if you’re shunting it very fast it suddenly flows very easily. The snail is
sailing over the top of this liquid mucus on those muscular waves, taking
advantage of the lower resistance. It needs the thick slime as well, so that it
has something to push against. The only reason why snails (and slugs) can
move is that the same mucus can behave either like a solid or like a liquid,
depending on how fast they force it to move. The huge advantage of this
method is that they don’t fall off the underside of things because they never
lift themselves away from the surface.
How does the slime manage this trick? It’s a gel of very long molecules
called glycoproteins, all mixed up together. When it’s just sitting still,
chemical links form between the chains, so it behaves like a solid. But when
you push hard enough, the links suddenly break, and all the long molecules
can slither over each other like strands of spaghetti. Let it sit still again, and
the links will re-form; and after only a second, you’ll have a gel again.
If I had known all this, could I have protected the seedlings? Not by
choosing a surface that they couldn’t stick to or climb over, it turns out. The
mucus can stick to pretty much anything you’d find around the house –
including non-stick pan liners. Experiments have shown that snails can even
stick to super-hydrophobic surfaces, the ones that water can hardly touch.
It’s a pretty amazing achievement, but probably one best appreciated by
those who don’t have precious seedlings to protect.
The same mechanism also explains non-drip paint. When this paint sits
still, it’s thick and gooey. But when you push on it with a brush, it becomes
much less viscous, and it’s easy to spread a thin, even layer of it over a wall.
As soon as you take the brush away, the paint goes back to being very
viscous and so it doesn’t run off down the wall before it’s dried.
Ketchup and snails are small, but this same bit of physics can have serious
consequences on a much larger scale. Christchurch in New Zealand was a
charming and peaceful city when I visited it in 2002. The land there is made
up of sediment, layer upon layer of tiny particles deposited by the Avon
river over successive millennia. It’s a beautiful location, but the city was
sitting on a time bomb. At 12.51 p.m. on 22 February 2011, a magnitude 6.3
earthquake struck just 6 miles away from the city centre. The earthquake
itself was bad enough, throwing people into the air and tearing buildings
apart. But the sediment that the town was built on was only strong and solid
if it was stationary. Just like the ketchup, powerful shaking turned it into a
liquid. The small-scale details are a tiny bit different – instead of the bonds
between long molecular chains breaking, water sneaks in between sand
grains and pushes them apart, allowing them to flow. But the overall
physics is the same: when it’s agitated quickly, the solid ground starts to
flow like a liquid.
A car is a heavy object, so gravity makes it push down hard on the
ground it’s sitting on. Cars don’t sink through the ground because the
ground is solid enough to resist that push. But for a few minutes in
Christchurch, that general rule was broken. Many cars that day were parked
on sandy roadsides, resting on packed soil that hadn’t moved for decades.
As the earthquake shook the ground, the layers of sand were forced to slide
over each other from side to side extremely quickly. If this had happened
slowly, the cars would have been safe. But it happened so quickly that water
crept in between the sand grains and the sand grains didn’t have time to
settle back into place before they were forced back in a different direction.
So instead of sand resting on sand, the ground was suddenly made of a
mixture of sand and water that had no fixed structure. Any car sitting on top
of this mixture would sink downwards into the mush as the shaking
continued. But as soon as the shaking stopped, it only took a second or so
for the sand grains to settle slightly, so that they were supported by other
sand grains instead of water. The ground had resolidified, but by now the
car was half-buried.
This process was responsible for a lot of the damage in Christchurch.
Cars sank into the silt and buildings fell because the ground couldn’t hold
them up. It’s known as ‘liquefaction’, and it takes something as powerful as
an earthquake to move the sediment fast enough to cause it. But if you
move soft, sandy ground fast enough, its strength will vanish. This is also
why flailing about in quicksand is such a terrible idea. If you fight and
struggle, the quicksand becomes liquid-like and you’ll just sink in. Move
slowly, and you stand a chance of controlling where you are. Time matters.
When you change the timescale of what you’re doing, you often change the
outcome.
We like to say that something was so fast ‘it happened in the blink of an
eye’. A blink takes about a third of a second, and the average reaction time
of a human is around a quarter of a second. That sounds pretty fast, but just
think about what has to happen in that time if you’re taking a standard
reaction test. When light rays hit your retina, specialized detection
molecules twist around, and this starts a chain of chemical reactions that
cause a small electric current. This signal travels through the optic nerve
into the brain, stimulating brain cells to send signals to each other as they
work out that this is something that requires a reaction. Then electrical
signals travel out to the muscles, slowed down when they are ferried across
the gaps between nerve cells by chemical diffusion. Once the order to
contract has been received, molecules in the muscle fibre ratchet over each
other until your hand hits the button. All that, just for you to do the fastest
thing you can possibly do.
Our fabulous complexity comes at the cost of speed. I think of humans as
pretty slow beasts, lumbering through the physical world, because so many
different stages are involved in everything we do. While we plod through
all that, many simpler physical systems are just getting on with things, lots
of things. But those simple, quick processes are too fast for us to see. You
can get a hint of this world if you drop a single drip of milk into your coffee
from quite high up. You might just see the drop bounce right back out
before falling back into the drink. It’s right on the edge of the fastest we can
see. My PhD supervisor used to say that if you were quick, you could
change your mind about having the milk and catch it on its way out, but I’m
pretty sure you’d need the help of something smaller and faster than a
human to do it.
The thought of how much we’re missing because we’re slow is what
inspired my PhD. I was fascinated by the idea of a world that could be
doing things right in front of my eyes, things that were too small and too
fast to see. So I chose a PhD that let me play with high-speed photography,
technology that let me see the parts of the world that are normally invisible
because they are so fast. But cameras like that are only available to humans.
What do you do if you have the same problem, but you’re a pigeon?
In 1977, an enterprising scientist named Barrie Frost persuaded a pigeon
to walk on a treadmill. This is one of those experiments that would
probably win an IgNobel prize these days, as the perfect example of a piece
of science that makes you laugh and then makes you think. As the treadmill
belt slowly moved backwards, the bird had to walk forward to stay in the
same place. The pigeon apparently got the hang of it quite quickly, but
something was missing as it plodded along. If you’ve ever sat in a town
square and watched pigeons strut around in search of food, you’ll have
noticed that their heads bob backwards and forwards as they walk. I’ve
always thought it looks like a really uncomfortable thing to do, and it seems
odd to put in all that extra effort. But the pigeon on the treadmill wasn’t
bobbing its head, and that told Barrie something very important about the
bobbing. The bird obviously didn’t need to do it in order to walk, so it
wasn’t anything to do with the physics of locomotion. The head-bobbing
was about what it could see. On the treadmill, even though the pigeon was
walking, the surroundings stayed in the same place. If the pigeon held its
head still, it saw exactly the same view all the time. That made the
surroundings nice and easy to see. But when a pigeon is walking on land,
the scenery is constantly changing as it goes past. It turns out that these
birds can’t see ‘fast’ enough to catch the changing scene. So they’re not
really bobbing their heads forwards and backwards at all. They thrust their
head forward, and then take a step that lets their body catch up, and then
thrust their head forwards again. The head stays in the same position
throughout the step, so the pigeon has more time to analyse this scene
before moving on to the next one. They get one snapshot of their
surroundings, and then they jerk their head forwards to get the next
snapshot. If you spend a while watching a pigeon, you can convince
yourself of this (although it takes a bit of patience, because they are usually
fn3
quite quick). No one seems to know exactly why some birds are so slow at
gathering visual information that they need to bob their heads, and others
aren’t. But the slower ones can’t keep up with their world without breaking
it down into a freeze-frame movie.
Our eyes can keep up with our walking pace, but if you need to examine
something close to you while you’re walking or running, you usually feel
an overwhelming urge to stop for a bit to have a good look. Your eyes can’t
collect information fast enough to get all the detail while you’re moving.
Humans actually play exactly the same game as the pigeon (without the
head-bobbing), and our brain stitches things together so that we’d never
know. Our eyes dart rapidly from place to place, adding information to our
mental image on each stop. If you look at yourself in a mirror and look
directly at the reflection of one of your eyes and then the other, you will
notice that you never see your eyes move, even though someone standing
next to you will see them flick from one side to the other. Your brain has
stitched your perception of the scene together in such a way that you’d
never know there was a jump; but those jumps are happening all the time.
The point is that we’re only a tiny bit faster than the pigeon, and this
highlights how much there must be that’s faster than us. We are used to life
at a limited range of timescales – we can follow things that last from about
a second to a few years – but that’s not all there is. Without science to help
us, we are blind to anything happening over a few milliseconds or over a
few millennia. We can only perceive our bit in the middle. That’s why
computers can do so much and part of why they seem so mysterious. They
can do what they need to do in tiny amounts of time, so they can get on
with it and finish amazingly complex tasks before we perceive any time
passing. Computers continue to get faster, but we can’t perceive why,
because a millionth and a billionth of a second are the same to us: both too
fast to notice. But that doesn’t mean the distinction isn’t significant.
What you see depends on the timescale on which you are looking. To
grasp the contrast, let’s compare the speedy and the ponderous: a raindrop
and a mountain.
A large raindrop takes one second to fall 6 metres, the height of a two-
storey building. What happens to it during that second? This raindrop is a
jostling cluster of water molecules, each one held firmly in the grip of the
group, but constantly shifting its allegiances within that group. A water
molecule, as we saw in the last chapter, consists of an oxygen atom
accompanied by two hydrogen atoms, one on either side, the trio forming a
V-shape. The whole molecule can bend and stretch as it hops through the
loose network formed by billions of identical others. In that one second, this
molecule may hop 200 billion times. If our molecule reaches the edge of the
multitude it will find that there’s nothing outside the droplet that can
compete with the huge attraction of the masses, so it’s always pulled back
to the centre. The cartoon raindrop shape is a fiction: raindrops have lots of
shapes but none of them have sharp points. Any pointed edges will be
rapidly smoothed away, because individual molecules can’t resist the pull of
the mob. But despite the strength of that pull, the perfect shape is never
reached. There is constant readjustment in response to the buffeting of the
air. A drop may be squashed flat, but will then pull itself back together,
overshoot, become stretched into a rugby ball shape and then back again,
170 times in this one second. The globule is constantly wobbling and re-
inventing itself, a battleground between the external forces trying to tear it
apart and the fierce pull of the mob keeping it together. Sometimes a
raindrop flattens into a pancake, then stretches into a thin umbrella, and
then explodes into an army of tiny droplets. All of this happens in less than
a second. We can’t see any of it, but that droplet has transformed itself a
billion times in the blink of an eye. Then the droplet splats down on to bare
rock, and the timescales shift.
This rock is granite. It has not moved or changed in human memory. But
four hundred million years ago there was a giant volcano in the southern
hemisphere, and magma from below squeezed into the gaps in the volcanic
rock. Over the following millennia the magma cooled, separating slowly
into crystals of different types, and became hard unyielding granite. As
more time has passed, the rocky leviathan has been ground down by ice
ages, chipped away by plants and ice, polished by rain. While the volcano
was wearing away, it was also travelling. Since the giant explosion that
finished it, this chunk of continent has been creeping north. On top of it,
species and geological eras came and went as the machinery of the planet
shunted the ill-fitting jigsaw pieces of its surface together and apart. Today,
a tenth of the total lifetime of our planet later, all that is left of the original
dramatic volcano is the sorry remains of its exposed guts. We call it Ben
Nevis, the highest mountain in the British Isles.
When you and I look at either the mountain or the raindrop, we notice
very little change. But that’s just because of our own perception of time, not
because of what we’re looking at.
We live in the middle of the timescales, and sometimes it’s hard to take
the rest of time seriously. It’s not just the difference between now and then,
it’s the vertigo you get when you think about what ‘now’ actually is. It
could be a millionth of a second, or a year. Your perspective is completely
different when you’re looking at incredibly fast events or glacially slow
ones. But the difference hasn’t got anything to do with how things are
changing; it’s just a question of how long they take to get there. And where
is ‘there’? It is equilibrium, a state of balance. Left to itself, nothing will
ever shift from this final position because it has no reason to do so. At the
end, there are no forces to move anything, because they’re all balanced. The
physical world, all of it, only ever has one destination: equilibrium.
Imagine a lock gate in a canal. Locks were invented for the most
ingenious of reasons: to allow boats on a canal to climb hills. They work
because boats can propel themselves forward against water flow, but only if
that water flow is really slow. No canal boat can power up a waterfall, but
with the help of a lock, a boat can still climb a hill. A lock consists of two
sets of gates which form a complete bottleneck in a canal, trapping an
isolated pool of water between them. On one side of the lock the water is
higher; on the other, it’s lower. Anything wanting to go up or down the
canal has to go through the lock. Let’s say there’s a boat waiting at the
bottom. The water in between the gates is initially at the same height as the
canal at the bottom. The lower gates open, our boat chugs into the lock, and
the lower gates close. Now the top gate is opened, just slightly, and water
flows into the lock. This is the important bit. When the top gates were
closed, the water above the lock had no reason to go anywhere. It was
sitting in the lowest place it could be, in equilibrium. There was nowhere
better for it to be, and it would stay put there indefinitely. But as soon as a
gap is opened that connects it with the pool of water between the gates, this
changes. Suddenly, there’s a route to somewhere better. Gravity is always
pulling the water downwards, and we’ve just opened the door for the water
to respond to gravity’s pull and move itself even further downwards. So it
flows in to join the boat, and it keeps filling the lock up until the water
height inside it is the same as the water height above the lock. No one had
to do anything other than provide the route to a new equilibrium. But now
the boat is at the same height as the top part of the canal, and once the gates
are fully opened it can chug along on its way upstream, against the very
slow canal flow. Behind it, once the gates are closed again, everything is in
equilibrium. The water between the locks will stay there indefinitely
because it has nowhere better to be. All the forces are balanced. Then at
some point a boat will enter the lock from upstream, someone opens the
lower gate and the water is allowed to flow out into the downstream canal,
where it will continue on its way to a new equilibrium.
The lesson of all this is that you can get a lot done in the world by
controlling where the equilibrium position is. Left to themselves, things
shuffle around until everything is balanced and then they stay there. The
way to get things done is to be in control of where equilibrium is. If you can
move the goalposts on demand, you can make sure that things flow in the
direction you want them to go in, and only when you say so.
The idea that the physical world will always move towards balance – that
hot and cold liquids will mix until everything is the same temperature, or
that a balloon will expand until the pressure is equal inside and out – is
related to the concept that time only flows one way. The world can’t run
backwards. Water is never going to flow by itself through a lock from the
lower level to the higher level. That means that you can tell which way is
forwards by looking for systems moving towards equilibrium. While
moving things by brute force will cost you a lot of energy, influencing the
speed of the slide to equilibrium often costs very little. It is also often
extremely useful.
The Hoover Dam is one of the biggest civil engineering achievements of
the last century. Driving towards it from Las Vegas, you weave through a
red rocky landscape where it seems impossible that anything large could be
hidden. The only clues that there might be something unusual nearby come
from occasional glimpses of sparkling blue water completely out of place in
the middle of a desert. And then you turn a corner and there it is, all 7.5
million tons of it: a giant concrete bung lodged in the middle of this rugged
American landscape.
A hundred years ago, the Colorado river ran unfettered through its
narrow canyon. Rain from high up in the Rocky Mountains and the vast
plains to the east was funnelled downhill through a series of valleys and out
to the Gulf of California. The problem for the farmers and city-dwellers
downstream wasn’t the amount of water – there was plenty of it – but the
timing of its arrival. In the spring, huge floods could wash away fields, but
by the autumn only a feeble trickle was left, not enough for the growing
population. The water was always going to start in the same mountains and
plains, and end up in the same bit of ocean. But what farmers and
fn4
townspeople alike really needed was to control when it got there, and
especially to stop it all arriving at once. And so the bung was built.
A drop of water that’s made its way off the Rockies and all the way down
through the Grand Canyon now finds itself in Lake Mead, the giant
reservoir that has built up behind the dam. It hasn’t got anywhere else to go,
at least not for a while. The key thing here is that the droplet is held where
it is – high up – because it can’t go down any further. In 1930, a droplet
leaving the Grand Canyon would have trickled 150 metres downwards
before it came to rest. But after 1935, when the dam was completed, that
same drop could reach this point and still be 150 metres above the valley
floor. The amazing thing is that it doesn’t take any energy to keep it there,
just a carefully placed obstacle to stop it going anywhere else. It’s in
human-created equilibrium and it’s staying put.
Until, of course, humans decide that they want it to go somewhere else.
Humans can control the flow through the dam, rationing the water that
feeds the rest of the Colorado river. There are no more floods downstream,
and the river never stops flowing completely. And there’s another benefit.
As the marshalled water flows past the dam, the huge pressure that has built
up turns turbines that produce hydroelectric power. The consequences of
this water shunting are that hundreds of thousands of people can live and
work in the arid deserts of the American south-west.
The Hoover Dam was built to control the timing of water flow, but the
principle it demonstrates goes far beyond water use. When it comes to
harvesting energy, all we are ever really doing is providing a few obstacles
to energy that was already on its way from somewhere to somewhere else.
The physical world will always move towards equilibrium, but sometimes
we can control where the nearest equilibrium is and how quickly something
in the world can get there. By controlling that flow, we also control the
timing of energy release. Then we make sure that as the energy flows
through our artificial obstacles on its way towards equilibrium, it does
something useful for us. We don’t create energy and we don’t destroy it. We
just move the goalposts and divert it.
Like many civilizations before ours, we face the problem of limited
resources. Fossil fuels are made of up plants that built themselves using
energy from the Sun, diverting that energy from its alternative outlet: gentle
warmth, which is the equivalent of the bottom of the river when it comes to
usefulness. Fossil fuels are the energy equivalent of dams, a form that stores
energy in a temporary equilibrium. When we come along, dig them up and
provide the right kick, we’re choosing the timing of energy release by
providing a route to another accessible equilibrium, via a flame and
chemical decomposition to carbon dioxide and water. The problem we have
is that there are only so many ‘upstream’ resources in the form of fossil
fuels, and in a few human lifetimes we have released energy that took
millions of years to accumulate. The fossil-fuel reservoirs are being
emptied, and they will not be refilled for millions of years more. Renewable
energy, like the hydroelectricity from the Hoover Dam and many others,
diverts the waterfall of solar energy that is flowing through our world now.
The game facing our civilization remains the same: how do we stop and
start the energy flow efficiently, so that we can do what we want without
changing our world too much?
Next time you turn on something that is battery-powered, you’re
choosing the time of energy release from the battery by opening an
electrical gate, and guiding the energy through the circuits of the device to
help you do something useful. After that, it’ll end up as heat, which it
would have done anyway. This is what the switches in our world are, all of
them. They’re the gatekeepers controlling the timing of a flow, and the flow
is only ever going one way: towards equilibrium. If we let the flow whoosh
through all at once we get one result; if we slow it down, letting it trickle
through at times that suit us, we get an entirely different result. Time
matters here because it’s only ever going in one direction: by choosing
when the flow towards equilibrium happens, and the speed of that flow, we
give ourselves enormous control over the world. But it’s not always the case
that things reach equilibrium and then stop. If they’re going really fast as
they approach the balance point, they may well just keep going and fly right
through. This opens the door to a whole new set of phenomena, including
some problems.
Mid-afternoon teabreak is an essential part of my working day. But I
noticed recently that even acquiring a mug of tea forces me to slow down,
and it’s not just about the time taken to boil the kettle. My office at
University College London is at one end of a long corridor, and the tearoom
is at the other end. The journey back to my office, accompanied by a full
mug of tea, happens at the slowest pace of my entire day (my normal
walking speed at work is somewhere between ‘brisk’ and ‘race pace’). It’s
not that there’s too much tea in the mug; the problem is the sloshing. Every
step makes it worse. Any sensible person would accept that slowing down
is a reasonable solution. But any physicist would do some experiments first,
just to see whether that’s the only solution. You never know what you might
find. And I wasn’t going to give in to the obvious without a fight.
If you put water in a mug, sit the mug on a flat surface and give it a bit of
a push, the water will start to slosh from side to side. What’s happening is
that as you shove it, the mug moves but the water initially gets left behind,
so it piles up against the side of the mug you’ve pushed. Then you have a
mug that has higher water on one side than the other, so gravity pulls the
higher water down, and the water on the other side is pushed upwards. For
an instant the surface is flat again, but the water has no reason to stop
moving. It just carries on going up the other side. Gravity is tugging on it as
it goes, but it takes a while to stop the water completely. By the time it’s
stopped, the water level is higher on the second side than the first, and then
the cycle starts all over again. If the mug is sitting on a flat surface, the
sloshing from one side to the other will gradually die away, and equilibrium
will be reached. But if you’re walking, things are different.
The cycle is where the problem lies. If you try the shoving test with mugs
of a few different sizes, you’ll see that the sloshing happens in the same
way for them all, but it happens more quickly in a narrow mug and more
slowly in a wider mug. A mostly full mug always sloshes the same number
of times each second, however big the initial push was. But that number
depends on the mug, and the thing that matters most is the mug radius.
There’s a conflict between the downward force of gravity, which is
pulling everything back to equilibrium, and the momentum of the fluid,
which is greatest just as it passes through the equilibrium position. In a
bigger mug, there’s more fluid and it has further to go, so the cycle takes
longer to turn around. The special frequency that each mug has is known as
its natural frequency, the rate it will slosh at if pushed and then left to get
back to equilibrium by itself.
I spent a while playing with the mugs in my office. I have one tiny one
with a picture of Newton on it that is only 4 cm across. Water in this one
sloshes about five times each second. The biggest one is about 10 cm
across, and it sloshes about three times each second. This large mug is old
and cheap and ugly and I’ve never really liked it, but I still have it because
sometimes you just need a lot of tea.
When I come out of the tearoom with my full mug and take a couple of
brisk steps down the corridor, I start the sloshing. If I want to get back to
my office without spilling the tea, I have to prevent this sloshing from
growing. This is the crux of the problem. As I walk, I can’t help rocking the
mug slightly. If the pace of that rocking matches the natural frequency of
the sloshing, the sloshing will grow. When you push a child on a swing, you
push in a regular rhythm that matches the rate of the swing, and so the
swing gets bigger. The same happens with the tea. This is called resonance.
The closer the external push is to the natural frequency of the sloshing, the
more likely it is that tea will be spilled. The problem for all thirsty humans
is that it just so happens that most people walk at a pace that is very close to
the natural sloshing frequency of the typical mug. The faster you walk, the
closer to it you get. It’s almost as if the system was designed to slow me
down, but it’s just an inconvenient coincidence.
So it turns out there isn’t really a satisfactory solution. If I use the tiny
mug, it sloshes too fast for my walking pace to make the sloshing worse and
the tea won’t spill. But I want more than a thimbleful of tea. If I use the
larger mug, my brisk walking is very close to its natural frequency, and
disaster is just three steps down the corridor. The only solution is to slow
down, so that the rocking from the walking is much slower than the
fn5
sloshing frequency. I feel better for having tried, but the lesson here is that
I can’t beat the time-dependency of the physics.
Anything that swings – oscillates – will have a natural frequency. It’s
fixed by the situation, and the relationship between how hard the pull to
equilibrium is and how fast things are going when they get there. The child
on the swing is just one example, along with a pendulum, a metronome, a
rocking chair and a tuning fork. When you’re carrying a shopping bag and
it seems to be swinging at a rate which doesn’t match your steps, that’s
because it’s just swinging at its natural frequency. Big bells have deep notes
because their size means that they take a long time to squish and stretch and
squish again, so they ring with a low frequency. We get a huge amount of
information about the size of objects by listening to them, and it’s because
we can hear how long they take to vibrate.
These special timescales are really important for us, because we can use
them to control the world. If we don’t want the oscillation to grow, we have
to make sure that the system isn’t pushed at its natural frequency. That’s the
game with the tea. But if we want an oscillation to continue without much
effort, we choose to nudge it along at its natural frequency. And it’s not just
people who use this. Dogs do, too.
Inca is poised and ready, focused on the tennis ball like a sprinter waiting
for the starting gun. As I lift the plastic arm holding the ball, she tenses, and
then the ball sails over her head and she’s off, a slim bundle of enthusiasm
and seemingly endless energy. Her owner Campbell and I chat while Inca
rushes happily across the scrubby grass. She doesn’t bring the thrown ball
back to us, because she’s already got a second tennis ball in her mouth
(apparently this is a ‘spaniel thing’), but when she reaches it she stands
guard until we catch her up and lob the first ball further ahead. After half an
hour of non-stop chasing, she finally sits down, tail cheerily swishing the
grass, and looks up at us, panting.
I kneel down and stroke her back. All that running around has made her
hot. She isn’t sweaty because dogs don’t sweat, but she still has to get rid of
all that excess heat. The panting looks like hard work, presumably using
lots of energy and generating even more heat. It seems like a bit of a
paradox. Inca is untroubled by my ponderings but quite happy to be
stroked, and a strand of saliva drips from her wide-open mouth. After I’ve
been out running, my breathing rate comes down back to normal quite
gradually, but when Inca stops panting it happens very suddenly. Big brown
eyes look up at me, and I wonder how much longer she needs to recover
before it’s time for more tennis balls.
By far the most efficient way to lose heat is to evaporate water. That’s
why we sweat. Turning liquid water into a gas takes a huge amount of
energy, and conveniently the gas then floats away, taking that energy with
it. Since dogs don’t sweat, they don’t produce water on their skin that can
evaporate, but they have plenty of water in their nasal passages. Panting is
all about pushing as much air as possible over the wet insides of their noses,
to get rid of heat quickly. As if to demonstrate the point, Inca starts panting
again. I reckon she’s taking about three breaths each second, which seems
like a lot of hard work. But the really clever bit is that it isn’t. Her lungs act
as an oscillator. This is the most efficient rate for her to breathe at because
it’s the natural frequency of her lungs. As she breathes in, she’s stretching
the elastic walls of the lungs, and after a while, the elastic pushes back
strongly enough to turn the cycle around. Just as the lungs get back to their
unstretched size, she puts in a tiny bit of energy to send them off on the
cycle again. The downside is that when she’s breathing this fast, she’s not
really replacing the air deep in her lungs, so she isn’t actually taking on
board much extra oxygen while all this is going on. That’s why she doesn’t
breathe like this all the time. But just at the moment the need to lose heat
trumps her need for oxygen, and by pushing her lungs at exactly the right
frequency, she’s getting as much air as possible through her nose for as little
effort as possible. So the panting is generating a tiny amount of heat
compared to what she’s losing. She’s breathing in through her nose, but
she’s got her mouth wide open because the dribbling is also cooling her.
When the saliva evaporates, that helps lose a bit of heat energy too. The
panting stops again, and Inca eyes up the abandoned tennis ball. One
inquiring look at Campbell is enough (he’s well trained) and the game
begins again.
The natural frequency of something depends on its shape and what it’s
made of, but the biggest factor is its size. This is why smaller dogs pant
faster. They’ve got tiny lungs, which naturally inflate and deflate many
more times each second. Panting is a very efficient way of losing heat if
you’re small. But it gets less efficient as you get bigger, and that may be
why larger animals sweat instead (especially hairless ones like us).
Every object has a natural frequency, and often more than one if there are
different possible patterns of vibration. As the objects get bigger, those
frequencies generally get lower. It can take quite a push to make a really
massive object move, but even a building can vibrate, very very slowly. A
building can in fact behave a bit like a metronome, a sort of upside-down
pendulum – the base is fixed while the top moves. Higher up, the wind is
faster than at ground level, and this is enough to give tall narrow buildings
the sort of shove that will start them swaying at their natural frequency. If
you’ve ever been in a tall building on a very windy day, you’ve probably
felt this. A single cycle can take a few seconds. It’s disconcerting for
humans inside, so the architects of these buildings spend a lot of time
working out how to reduce the swaying. They can’t remove it completely,
but they can change the frequency and flexibility to make it less noticeable.
If you feel it happening, don’t worry – the building will have been designed
to bend, and it isn’t going to fall.
The wind may be gusty, but it doesn’t push in a regular rhythm that could
match the building’s natural frequency, so there’s a limit to how bad the
swaying can get. But the jolt of an earthquake sends out ripples in the
ground, huge waves travelling out from the epicentre, slowly tipping the
earth from side to side. What happens when a tall building meets an
earthquake?
On the morning of 19 September 1985, Mexico City started to move.
Tectonic plates underneath the edge of the Pacific Ocean, 217 miles away,
ground over each other to generate an earthquake of magnitude 8.0 on the
Richter scale. In Mexico City, the shaking lasted for three to four minutes,
and it tore the city to pieces. It’s estimated that ten thousand people lost
their lives, and massive damage was inflicted on the city’s infrastructure.
Recovery took years. The US National Bureau of Standards and the US
Geological Survey dispatched a team of four engineers and one
seismologist to assess the damage. Their detailed report showed that a
horrid coincidence of frequencies was responsible for a lot of the worst
damage.
First of all, Mexico City sits on top of lake-bed sediments that fill a hard
rock basin. The earthquake monitoring devices showed beautiful regular
waves with a single frequency, even though normally earthquake signals are
much more complex than that. It turns out that the geology of the lake
sediments gave them a natural frequency of oscillation, and so they had
amplified any waves that lasted about two seconds. The whole basin had
temporarily become a tabletop shaking at almost exactly one single
frequency.
The amplification was bad enough. But when they looked at specific
damage, the engineers discovered that most of the buildings that had
collapsed or were badly damaged were between five and twenty storeys
high. Taller or shorter buildings (and there were plenty of both) had
survived almost untouched. They worked out that the natural frequency of
the shaking closely matched the natural frequency of the mid-sized
buildings. With a long-lasting regular push at exactly the right frequency,
these buildings had been twanged like tuning forks, and they didn’t stand a
chance.
These days, controlling the natural frequency of buildings is taken very
seriously by architects. Management of shaking is even sometimes
celebrated. In the Taipei 101, a 509-metre monster in Taiwan that from
2004 to 2010 was the largest building in the world, the place to visit is the
viewing galleries on the 87th–92nd floors. This section of the building is
hollow, and suspended inside it is a 660-ton spherical pendulum, painted
gold. It’s beautiful and weird – and practical. It’s there not just as an
aesthetic quirk, but to make the building more earthquake-resistant. The
technical name for it is a tuned mass damper, and the idea is that when
there’s an earthquake (a common occurrence in Taiwan), the building and
the sphere swing independently. When an earthquake starts, the building
sways one way and pulls the spherical pendulum sideways too. But by the
time the sphere has moved in that direction, the building has swayed back
the other way, and is now tugging the sphere back. So the sphere is always
pulling in the opposite direction to the movement of the building, reducing
its sway. The sphere can move 1.5 metres in any direction and it reduces the
fn6
overall oscillation of the whole building by about 40 per cent. The humans
inside would be far more comfortable if the building never moved. But
earthquakes shove the building out of equilibrium so that it has to move.
The architects can’t stop that happening, but they can tweak what happens
on the return journey. The occupants of the building have no choice but to
sit tight as the huge tower sways past the equilibrium position and back
again, until the energy is lost and serene stasis is restored.
WHEN YOU GO to the beach, it’s almost impossible to stand for any length of
time with your back to the sea. It feels wrong, both because you’re missing
out on the grandeur of the sight and also because facing the other way stops
you keeping an eye on what the ocean might be up to. And it’s oddly
reassuring to watch the boundary between sea and land as it constantly
renews and remodels itself. When I lived in La Jolla, California, my reward
after a long day was to wander down to the ocean, sit on a rock, and watch
the waves as the sun went down. Just a hundred metres off shore, the waves
were long and low, difficult to see. As they rolled towards the shore they’d
get steeper and more obvious until they finally broke on the beach. I could
sit and watch the endless supply of new waves for hours.
A wave is something that we all recognize, but they can be hard to
describe. The ones at the seashore are processions of ridges, a wiggly shape
in the water surface that is travelling from over there to over here. We can
measure them by looking at the distance between successive wave peaks
and the height of the peaks themselves. A water wave can be as tiny as the
ripples you make when you blow on your tea to cool it, or bigger than a
ship.
But waves have one quite weird feature, and in La Jolla it was the
pelicans that made it obvious. Brown pelicans live all along that coast, and
they look so ancient that you wonder whether they’ve just flown through a
wormhole from a few million years ago. They’ve got ridiculously long
beaks that usually stay folded up against their bodies, and small groups of
these curious birds are often seen gliding solemnly just above the waves
parallel to the coast. Once in a while, they’d plonk themselves down
unceremoniously onto the ocean surface. And this was the interesting bit.
The waves that the birds were sitting on rolled endlessly towards the shore,
but the pelicans didn’t go anywhere.
Next time you stand on the shore and watch waves rolling towards you,
fn1
watch the seabirds sitting on the surface. They’ll be parked quite happily,
passengers being carried up and down as the waves go past, but they’re not
fn2
going anywhere. What this tells you is that the water isn’t going anywhere
either. The waves move, but the thing that is ‘waving’ – the water – doesn’t.
The wave can’t be static; the whole thing only works if the shape is moving.
So waves are always moving. They carry energy (because it takes energy to
shift the water into the wave shape and back again), but they don’t carry
‘stuff’. A wave is a regular moving shape that transports energy. I think this
is partly why I found sitting on the beach and looking out to sea so
therapeutic. I could see how energy was continually carried towards the
shore by the waves, and I could see that the water itself never changed.
Waves come in many different types, but there are some basic principles
that apply to them all. The sound waves made by a dolphin, the water
waves made by a pebble and the light waves from a distant star have a lot in
common. And these days, we don’t just respond to the waves that nature
provides for us. We also make our own, very sophisticated, contribution to
the flood, and it connects the scattered elements of our civilization. But
humans consciously using waves to cement cultural bonds isn’t new. This
story began centuries ago, in the middle of a gigantic ocean.
A king surfing the ocean waves probably sounds like a snapshot from a
particularly weird dream. But 250 years ago in Hawaii every king, queen,
chief and chiefess owned a surfboard, and royal prowess at the national
sport was a source of considerable pride. Special long narrow ‘Olo’ boards
were reserved for the elite, while the commoners used the shorter and more
manoeuvrable ‘Alaia’ boards. Contests were common, and provided the
fn3
central drama for many Hawaiian stories and legends. When you live on a
stunning tropical island surrounded by deep blue ocean, building a culture
around playing in the sea sounds perfectly sensible. But the Hawaiian surf
pioneers had something else going for them: the right sort of waves. Their
small island nation in the middle of a vast ocean was perfectly placed.
Hawaiian geography and physics filtered the complexity of the ocean, and
kings and queens surfed on the consequences.
While the Hawaiians were chanting to urge the flat, windless sea to rise
into ready-to-surf swell, the ocean thousands of miles away could have
looked very different. The winds in massive storms shove on the ocean
surface, dumping energy by forcing the water up into waves. But the waves
in storms are confused mixtures of short and long waves travelling in
different directions, breaking and rebuilding and clashing. Winter storms
are common at a latitude of about 45°, so the storms would be to the north
of Hawaii in the northern hemisphere winter, and to the south of Hawaii in
the southern hemisphere winter. But waves have to travel. Even as the
storm winds were dying down, the patch of ruffled ocean would have been
expanding out past the edges of the storm and into undisturbed water. Out
here, a sorting process could take place. The true nature of the confused
mess would be revealed – not jumbled chaos, but a crowd of different wave
types all sitting right on top of each other. Water waves that have a longer
wavelength (that’s the distance between peaks) travel faster than those with
a shorter wavelength. So the first waves to escape would be the longest,
racing outwards ahead of their shorter cousins. But there is a price to pay as
a water wave travels. Energy will gradually be stolen by the surroundings,
and the price per mile is higher for the shortest waves. Not only are they
losing the race, they’re losing their power as well, and it doesn’t take too
long for them to vanish. Thousands of miles from the storm and days later,
all that remains are the longest waves, a smooth regular swell, radiating out
across the planet.
So Hawaii’s first advantage is being in a spot far enough away from the
massive storms to experience them only in the form of that residual smooth,
tidy, long-wavelength swell. Its second advantage is that the Pacific Ocean
is very deep and the islands’ volcanic sides are steep. Waves travel across
the ocean surface undisturbed until they suddenly meet a steep slope. Then
all the energy that was spread over a huge depth has to become more
concentrated in the shallows, so the height of the waves must increase. And
very close into shore, the Hawaiians were waiting for the last gasp of these
slow monsters, as the waves became so steep that they had to break over the
perfect beaches of the islands. And as they broke, the kings and queens
were ready with their surfboards.
Water waves are probably the first waves that most people are aware of.
Something that a duck can bob about on is easy to imagine and to
understand. But waves come in lots of different types, and many of the
same principles apply to them all. All waves have a wavelength, a
measurable distance between one peak and the next. Because they’re
moving, all of them also have a frequency, the number of times they go
through a cycle (peak to trough and back to peak again) in one second. All
waves have a speed, too, but some of them (like the water waves) travel at
different speeds depending on their wavelength. The problem with most
waves is that we can’t see what’s doing the waving. Sound waves travel
through air, and they’re compression waves; instead of a moving shape,
what’s passed along is a push. The hardest waves to imagine are the most
common of all: light waves, which move through electric and magnetic
fields. But even though we can’t see electricity, we can see the effects of
fn4
light being a wave all around us.
One of the main reasons that waves are interesting and useful is that the
environment they’re passing through often changes them. By the time a
wave is seen or heard or detected, it’s a treasure trove of information
because it carries the signature of where it’s been. But that signature is only
stamped in relatively simple ways. There are three main things that can
happen to a wave: it can be reflected, it can be refracted, or it can be
absorbed.
If you wander past the fish counter at a supermarket and look at what’s on
offer, what you see is mostly silver. The exceptions to the rule are tropical
fish like red mullet and red snapper, and the bottom-dwelling fish such as
sole and flounder. But mostly, you’re looking at fish that swim in the open
ocean in big schools, like herring, sardines and mackerel. Silver is
interesting because it isn’t really a colour. It’s just our word for something
that acts as a trampoline for light, bouncing it back out into the world. All
waves can be reflected, and almost all materials reflect some light. What’s
special about silver is that it sends everything back indiscriminately. Every
colour is treated in the same way, no exceptions. Polished metal is really
good at this trick, and it’s useful because the angle at which the light arrives
is the angle at which it leaves. If you take an image of the world and use a
mirror to bounce it in a different direction, the relative angles of all those
light rays stay the same. It’s difficult to polish metal perfectly enough to get
a perfect image, and mirrors have been very highly prized in human history.
And yet we take silver fish for granted. The fish can’t even use metal; in
order to be silvery, they’ve got to build structures that do the same job out
of organic molecules. That’s complicated, and therefore expensive in
evolutionary terms. If you’re a herring, why do you bother?
Herring roam the seas in schools, feeding on small shrimp-like creatures
and hoping to avoid the big carnivores: dolphins, tuna, cod, whales and sea
lions. But the oceans are huge, empty places with nowhere to hide. The
only solution is invisibility, or the closest that nature can come to it:
camouflage. So should fish be blue, to match the watery background? The
problem with that is that the exact hue depends on the time of day and
what’s in the water, so it changes all the time. But the herring absolutely
must look like the water behind them, in order to survive. So they turn
themselves into swimming mirrors, because the empty ocean behind them
looks exactly like the empty ocean in front of them. They can reflect 90 per
cent of all the light that falls on them, similar to a high-quality aluminium
mirror. By bouncing light waves back out into the eyes of potential
predators, a herring can swim about behind a shield made of light.
Reflection isn’t always this perfect. Quite often, only some of the light is
reflected by an object. But that’s fantastically useful if two objects are
sitting next to each other and we want to tell them apart. The one reflecting
blue light is my tea mug, and the one reflecting red is my sister’s. So
reflection matters when a wave hits a surface. But it’s not the only thing that
can happen when a wave meets a boundary. Refraction can shunt waves
about in a more subtle way, altering how they travel.
When a Hawaiian queen stood on a cliff overlooking the coast, watching
the surf build, she would have noticed that even though the swell out on the
open ocean was approaching from a different direction each day, at the
point the water waves reach the shore, they are always parallel with the
beach. Waves don’t ever come in sideways, whatever direction the coast
faces. That’s because the speed of water waves depends on the depth of the
water, and waves in deeper water will travel faster. Imagine a long, straight
beach with swell coming in from a direction that’s slightly to the left of
straight-on. The part of the wave crest that’s on the right, further away from
the shore, is in deeper water. So it travels faster, catching up the closer part
of the wave, and the whole wave crest turns clockwise as it moves towards
the shore, lining up with the beach. By the time the wave breaks, the wave
crest is parallel to the shore. So you can change the direction that a wave is
travelling in by changing the speed of some parts of the wave crest relative
to others. This is called refraction.
It’s easy to imagine changing the speed of a water wave, but what about
light? Physicists are always talking about ‘the speed of light’. It’s an
unimaginably gigantic speed, and a crucially important fixture in Einstein’s
most famous legacies: the Theories of Special and General Relativity. The
discovery that there is a constant ‘speed of light’ was controversial and
difficult to accept and brilliant. So it feels a bit like spoiling the party to tell
you that you have never in your life detected a light wave that was
travelling at the speed of light. Even water will slow light down, and you
can confirm this for yourself with a coin and a mug.
Put the coin flat on the bottom of the mug so that it’s touching the side
closest to you. Now bend down until the edge of the mug just hides the coin
from you. Light travels in straight lines, and at this point there is no straight
line that can get from the coin to your eyes. Now, without moving your
head or the mug, fill the mug up with water. The coin will appear. It hasn’t
moved, but the light from it changed direction as it left the water and now it
can reach your eyes. It’s an indirect demonstration that water slows down
light. As the light meets the air, it speeds up again and so the wave is bent
through an angle as it crosses the boundary. We call this refraction. And it’s
not just water that does this; everything light passes through slows it down,
but by different amounts. The ‘speed of light’ means its speed in a vacuum,
when light is travelling through nothingness. Water slows light down to 75
per cent of that speed, glass to 66 per cent, and light in diamond is dawdling
along at 41 per cent of its maximum speed. The more it’s slowed down, the
bigger the bend at the boundary with the air. This is why diamonds are so
much more sparkly than most gems – they slow light down much more than
fn5
the others. And that bending is the only reason that you can actually see
glass, water or diamonds. The material itself is transparent, so we don’t see
it directly. What we see is that something is messing about with light
coming from behind it, and we interpret that something as a transparent
object.
It’s nice that we can see diamonds (and will come as a relief to anyone
who has shelled out for one), but refraction isn’t just about aesthetics.
Refraction gives us lenses. And lenses opened the doors to a huge chunk of
science: microscopy to discover germs and the cells that we’re made of,
telescopes to explore the cosmos, and cameras to record the details
permanently. If light waves always travelled at the speed of light, we would
have none of those things. We live in a bath of light waves, and those waves
are constantly being reflected and refracted, slowed down and sped up as
they travel. Just like the chaos of the stormy ocean surface, overlapping
light waves of different sizes are travelling in every possible direction
around us. But by selecting and refracting, keeping some waves out and
slowing others down, our eyes marshal a tiny fraction of that light so that
we can make sense of it. The Hawaiian queen standing on the cliff was
watching water waves by using light waves, and the same physics applies to
both.
That’s all fine if some waves have arrived for you to see after being
reflected or refracted. But what if they never reach you at all?
One of life’s little oddities is that if you give a child some crayons and
tell them to draw water coming out of a tap, the water in their picture is
blue. But no one has ever seen blue water coming out of a tap. Tap water
has no colour (if yours has, I suggest you seek advice from a plumber). If
you did see blue water coming out of a tap, you certainly wouldn’t drink it.
But the water in pictures is always blue.
On satellite pictures of the Earth, the oceans are definitely blue. It’s not
because of the salt – there are ponds of salt-free meltwater on top of
glaciers, and they are also a stunningly deep, spectacular blue. They almost
look like someone has filled pockets in the ice with blue food colouring.
But where water is trickling over the ice to join the rest of the meltwater, it
has no colour. What matters for the colour isn’t what’s in the water, but how
much water you have.
Light waves hitting the water surface are either reflected back up into the
sky or pass through and travel down into the depths. But sometimes, a tiny
particle or even the water itself acts as an obstacle, sending the wave off in
a new direction. This redirection may happen to the same light wave
enough times that it eventually makes its way back out to the air. And on
that long journey, the water has filtered the light. The light waves coming
from the Sun are a mixture of lots of different wavelengths, all the colours
of the rainbow. But the water can absorb light, and it absorbs some colours
more than others. The first to go is the red light – a few metres of water is
enough to get rid of most of that. And then the yellows and greens follow
after a few tens of metres. But blue light is hardly absorbed at all – it can
travel for huge distances. And so by the time the light is on its way out of
the ocean, most of what’s left is blue. The reason tap water is colourless is
that there isn’t enough of it to make a difference. Tap water does have a
colour, the same colour as all the other water in the world. But that colour is
so faint that you need a huge amount of water all together to actually see the
fn6
effect that the water is having on the waves going through it. When you do
see it, it’s spectacular, and bright blue crayon really is the right choice. But
you’d never learn that from a tap.
So as waves travel, they can be absorbed by whatever they’re passing
through. It’s a very slow process of attrition, sneaking away wave energy
tiny bit by tiny bit. The amount that’s lost depends on what type of wave it
is and also its wavelength. All this variability means there’s a huge richness
in what waves are doing and what they can tell us. We can see and hear
some of the contrasts in one of my favourite atmospheric phenomena: the
thunderstorm.
A thunderstorm is a magnificent spectacle, a dramatic reminder that air is
far more than an invisible filler for the sky. Our atmosphere is host to vast
quantities of water and energy, and usually these hefty commodities are
shunted around slowly and peacefully. The thundercloud, the mighty
cumulonimbus, develops in order to rebalance the atmosphere when
peaceful shunting is no longer enough. The system starts when buoyant,
warm, moist air near the ground shoves upwards into the cooler air above,
taking huge amounts of energy with it. In the centre of the vast cloud, hot,
humid air rises rapidly, churning the atmosphere above it and liberating
huge raindrops. Most dramatic of all, the churning causes electrical charges
to be separated and redistributed to different parts of the clouds. The
charges accumulate until nearby clouds or the Earth itself are stabbed by
giant pulses of electrical current, carrying the excess electrical charge away.
Each lightning bolt lasts for less than a millisecond, but the thunder echoes
across the landscape for far longer. I love thunder and lightning, both for the
theatrical spectacle and for the glimpse it gives us into the atmospheric
engine. Thunderstorms produce such unlikely opposites: the sharp,
shocking flash of lightning contrasting with the deep, drawn-out rumble of
thunder. But both are beautiful examples of the versatility of waves.
The lightning bolt is temporary. The electrical connection is a
superheated tube of atmosphere, stretching from the thundercloud to the
Earth or perhaps to another cloud. It’s a corridor full of molecules that have
been blown apart by the energy rushing past them. For a brief instant, the
temperature in that tube may reach 50,000°C, and so it blazes blue-white. A
giant pulse of light waves whooshes outward from the tube, filling the
landscape, but they rush away at such an enormous speed that they’re gone
in an instant. As the superheated tube carrying the electrical current heats
up, it expands sideways, thumping into the air around it. This gigantic
pressure pulse ripples outward through the air, following the light, but much
more slowly. These are sound waves, and this is the thunder. We know that
lightning bolts exist because they make both light and sound waves.
The most important thing about a wave is that it’s a way of letting energy
move, but without also having to move air, water or ‘stuff’ of any kind.
This means that waves can billow through our world very easily, disturbing
things enough to be interesting and useful, but not so much that they’re
shoving our world about and causing major disruption. A lightning strike
liberates a lot of energy, and light and sound waves can carry some of that
energy out into the rest of the world, sharing it out. Even though the air
doesn’t go anywhere overall as the sound ripples past, huge amounts of
energy are transferred onwards. Light and sound are different types of
wave, but the same basic principles apply to both. For example, both light
and sound can be changed by the environment that they pass through. In the
case of thunder, we can directly hear what’s happening to the waves.
My favourite place to be is about a mile from the lightning strike. Once
the flash has signalled that the sound is on its way, I like to imagine that
giant pressure ripple spreading out towards me. As I look out across the
landscape, I can see right through the ripple, but it takes a few seconds to
reach me with the first whipcrack of thunder. These sound waves are
travelling at about 340 metres every second or 767 mph, which means
they’re taking 4.7 seconds to cover a mile. That sharp crack is similar to the
original sound made as the lightning bolt expanded right at the ground. But
here’s what makes the sound of thunder so distinctive: what I hear just after
the initial crack is the sound from slightly higher up the lightning bolt. It
started as the same sound, but it took longer to reach me because it had to
travel a sloping, and therefore longer, path. And then as the thunder rumbles
on, I’m hearing the sound from higher and higher up that same lightning
bolt. If it takes five seconds for the first crack to reach me, it’ll take two
more seconds before the sound from one mile up hits me, and another four
seconds before the sound from two miles up arrives. All these sound waves
started off more or less the same, just in different places. And that means
that as I listen, I can hear how the atmosphere is changing these waves. As
time goes on, the only difference is that they’ve travelled further. So the
highest-pitched sounds, that first sharp crack, disappear very quickly as the
high-frequency waves are absorbed by the atmosphere, but the lower-
frequency waves rumble on. As time goes on, and the waves have travelled
further and further, the overall pitch gets lower and lower, because the
highest notes are consumed by the air, but the lowest notes just keep going.
If you’re far enough away, the air takes it all and the sound never reaches
you. But the lightning has a greater reach – these light waves are different
and they don’t depend on the air for assistance as they travel. They don’t
get absorbed by air as easily, but they can be altered in other ways as they
whoosh through the world.
In a sense, waves are very simple. Once they’ve been made, they are
always on their way to somewhere else. And whether they’re sound waves
or ocean waves or light waves, they can be reflected or refracted or
absorbed by their environment. We live our lives in the middle of this
complex flood of waves, sensing the patterns in those that give us clues
about our surroundings. Our eyes and ears tune into the vibrations all
around us, and those vibrations carry two very important commodities:
energy and information.
On a grim, grey, cold winter day, toast is the perfect comfort food. The only
problem is that the gratification is not instant. I usually put the kettle on for
tea, then put bread in the toaster, and then pace the kitchen impatiently
while I wait for my treat to be ready. After I’ve washed a mug or two and
tidied the work surface, I often find myself staring into the toaster, checking
on what it’s up to. The nice thing about toasters is that you can see they’re
up to something, because the heating elements glow red. They’re not only
heating up the air that touches them, they’re radiating light energy too. And
this glow is a built-in thermometer. You can tell how hot the element is just
from its colour. This bright red tells me that the innards of my toaster have
reached 1,000°C. That’s horrifically hot – enough to melt aluminium or
silver. But if it’s glowing that bright cherry red, then 1,000°C is how hot it
is. It’s a rule that comes from the way our universe works. Everything that
is this temperature will glow the same colour of red, and other colours
indicate other temperatures. If you look into a coal fire and see the
innermost coals glowing bright yellow, you know that they are around
2,700°C. Something that is white-hot is 4,000°C or above. But when you
think about it, that’s odd. Why should colour have anything to do with
temperature?
While I’m staring into the toaster, I’m watching energy transform from
heat to light. One of the most elegant things about the way the universe
works is that anything that has a temperature above absolute zero is
constantly converting some of its energy to light waves. And light must
travel, so the energy whizzes out into the surroundings. The red-hot heating
element is converting some of its energy into red light waves, at the long-
wavelength end of the rainbow. But most of the energy it’s emitting has
even longer wavelengths than that, and we call these waves infrared.
Infrared is just like the light we can see, except that each wave is longer. We
can only detect it indirectly, by feeling the warmth where it’s been
absorbed. Even though we can’t see them, infrared waves are essential for a
toaster – they are what heats the toast up.
Hot objects send out more light at some wavelengths than others. At any
temperature, there’s a peak wavelength which accounts for most of the
light, and the radiated light dies away on either side of that peak. The
toaster is sending out a big bulge in the infrared, and the tail of the bulge is
visible red. So I see red. I can’t see the light that’s heating my toast, but I
can see the tail of longer wavelengths.
If I had some kind of super-toaster that could get even hotter, perhaps to
2,500°C, the heating elements would look yellow. That’s because the hotter
object would send out light with shorter wavelengths, so the visible tail
would include more of the rainbow: red, orange, yellow, and a little bit of
green. When we see both red and green light together, we interpret that as
yellow. Only something that has this temperature would send out this exact
segment of the rainbow. And if the temperature increased even more – if I
had a hyper-toaster that could get to 4,000°C – the light sent out would
include the whole rainbow, all the way to blue. And when we see all the
rainbow colours at once, we see white. So something that is white-hot is
actually sending out a rainbow, but all the colours are mixed up. The
disadvantage of the hyper-toaster is that it would melt pretty much whatever
you made it out of. But it would brown your toast very quickly. And
possibly your kitchen as well.
So a toaster is just a way of making waves. The red light waves that you
see are just some of the waves that it has made because of its temperature.
The infrared waves that you can’t see heat up your toast. This is why toast
only browns at the surface in a toaster; it’s only the bits that the light
touches that can absorb infrared and heat up. The reason I’m quite happy to
stare at the toaster while I’m waiting is that I’m imagining all the light it’s
giving out that I can’t see. I know it’s there, because the red glow is a
giveaway.
But of course, there’s a catch. The problem with this method of
generating light waves is that you always get the same set of waves
together. There’s no way to choose some of them but not others. An orange-
hot coal and molten steel and anything else that’s 1,500°C must emit the
same collection of colours all together. So you can measure the temperature
of something by its colour, when it’s hot enough for you to see the colours.
The surface temperature of the Sun is about 5,500°C – that’s why it gives
out white light. In fact, this is the only reason we can see stars in the night
sky; they’re so hot that light must pour out from their surface and across the
universe, light with a specific colour that gives their temperature away.
And we – you and I – we also have a colour because of our temperature.
It’s not a colour that we can see, but it’s visible to special cameras adapted
for the right sort of infrared. We’re much cooler than the toaster, but we’re
still glowing. We emit light waves with wavelengths that are mostly 10–20
times longer than visible light. Each of us is a lightbulb in the infrared, just
because of our body temperature. And so are dogs and cats and kangaroos
and hippos – all warm-blooded mammals. Anything and everything that is
above absolute zero (the scarily cold temperature of −273°C) is a lightbulb
like this, with the colour crossing from the infrared to even longer
wavelengths (the microwave range) as the temperatures get colder.
So we live our lives bathed in waves, and not just the ones we can see,
the ones that might catch our eye if we look in the right direction. The Sun,
our own bodies, the world around us, and also the technology we create are
constantly making light waves. And the same goes for sound waves – high
notes, low notes, the ultrasound that bats use to hunt and the infrasound that
elephants use to follow the weather. The amazing thing is that all of these
waves can be travelling through the same room, and none of them will
interfere with any of the others. The sound waves are the same whether a
room is completely dark or full of disco lights. The light waves aren’t
affected by piano concertos or screaming babies. All of this is what we tap
into when we open our eyes and use our ears. We’re just siphoning off some
of the useful bits from the flood, selecting the waves that send us the most
useful information.
But which ones do you choose? The answer will be different for the
newest self-driving cars and for an animal that needs to survive in a forest.
There’s a huge richness of information out there, and you can pick and
choose which of the waves will help you most. That is why blue whales and
bottle-nosed dolphins can hardly hear each other, and also why neither of
them gives a hoot about the colour of your wetsuit.
The transition from gas to liquid and back again is happening all the time
around us. But we don’t see the transition from liquid to solid and back
nearly as often. For most metals and plastics, melting happens a long way
above everyday temperatures. For smaller molecules like oxygen, methane
and alcohol, melting happens at fantastically low temperatures, the sort of
temperatures that require very specialized freezers. Water is an unusual
molecule, since it both melts and evaporates at temperatures that occur
around us fairly regularly. But when we think of frozen water, we think
most often of the North and South Poles of the Earth. They’re cold, white,
and forever associated with the great expeditions of the twentieth century
that took humans into some of the most inhospitable environments on the
planet. Freezing water caused them a lot of problems. But sometimes it also
offered unusual solutions.
The transition from gas to liquid is all about molecules getting close
enough to each other to touch, while still moving freely enough to flow
over each other. The transition from liquid to solid is about the moment
those molecules get locked into place. The freezing of water is the most
common example of this, but water freezes like almost nothing else. There’s
nowhere this weirdness is more visible than the frozen north – the Arctic
Ocean.
If you travel to the northernmost part of Norway, stand on the coast and
look still further north, you see the sea. During the ice-free summer months,
the 24-hour daylight nourishes vast mobile forests of tiny ocean plants, a
seasonal smorgasbord that attracts fish, whales and seals. Then, towards the
end of the summer, the light starts to disappear. The surface water
temperature, which only reached 6°C even at the height of the summer,
starts to drop. The water molecules, slipping and sliding over each other,
slow down. The water is so salty here that it can get down to −1.8°C and
stay liquid; but one clear dark night, the ice starts to form. Perhaps a flake
of ice is blown on to the water, and if the slowest water molecules bump
into it, they will stick. But they can’t stick just anywhere. Each new
molecule rests at a fixed place relative to the others, and the jumble of
bustling molecules is replaced by a crystal, in which well-ordered water
molecules are marshalled into a hexagonal lattice. And as the temperature
drops further, the ice crystal grows.
The utterly weird thing about water crystals is that the rigorously aligned
molecules take up more space now than when they were dashing around in
the warm. With almost any other substance, parking molecules on a regular
grid would make them sit closer together than when they’re allowed to
roam free. But water’s not like that. Our growing crystal is less dense than
the water around it, and so it floats. Water expands as it freezes. If it didn’t,
the newly frozen ice would sink, and the polar oceans would look very
different. But as it is, the temperature drops further, the freezing ice
expands, and the ocean grows itself a coat of solid white water.
There are lots of things in the frozen Arctic to get excited about: polar
bears and ice and the Northern Lights. But there’s one particular piece of
Arctic history that I absolutely love, a story that’s all about the peculiarities
of ice freezing, and of working with nature rather than against it. It’s about
a bulbous, stout little ship which survived one of the most extraordinary
voyages in the history of polar exploration. She’s called the Fram.
Explorers in the late 1800s were drawn to the North Pole. It wasn’t that
far away from western civilization. The northern parts of Canada,
Greenland, Norway and Russia had all been visited and at least roughly
mapped. But the North Pole itself was a big mystery. Was it land? Sea? No
one had ever reached the pole, so no one knew. The voyage defeated
explorers again and again because the sea ice grew and shrank and shifted.
As weather conditions changed, sea ice could pile up on top of itself,
making ridges and ice quakes. The grip of this ice could grind ships to
pieces. The USS Jeanette suffered a typical fate in 1881, becoming trapped
in sea ice for months just off the northern coast of Siberia. As the weather
cooled, and molecules of sea water locked on to the bottom of the ice lattice
at the sea surface, the expanding ice gripped the hull. After months of ice
growing and shrinking, squeezing then releasing the ship, the USS Jeanette
succumbed and was crushed. Explorers who made it off their vessels on to
solid ice faced different perils: the ice could melt and open up huge canals,
impassable except with a boat. From any of the countries around the Arctic
Circle it was hundreds of miles to the pole, and the shifting ice was a
formidable obstacle.
Three years after she sank, unmistakable wreckage from the USS
Jeanette washed up near Greenland. It was an astonishing find because the
wreckage had crossed the entire Arctic, right from one side to the other.
Oceanographers wondered whether there was a current that left the coast of
Siberia, travelled across the North Pole, and carried on to Greenland. And a
young Norwegian scientist called Fridtjof Nansen had a wild idea. If he
could make a ship that would withstand the ice, he could take it to Siberia
and freeze it into the ice where the USS Jeanette had sunk, and maybe three
years later he’d pop out in Greenland. But crucially, on the way, he might
pass over the top of the North Pole. No trekking, no sailing … just let the
ice and the wind do the work for you. The only problem would be the wait.
Nansen’s reward for this idea was to be both hailed as a genius and derided
as a madman. But he was going anyway. He raised the money and
employed one of the best naval architects of the age, because the ship itself
would have to be like no other ship ever floated on the ocean. And so the
Fram was made.
The difficulty was that as water freezes, the water molecules must take
their places in their rigid lattice. If the temperature sinks low enough, they
will stick. And if there isn’t enough room to sit in their proper place, they’ll
push outward on everything nearby in order to make space. Any ship frozen
into the ice suffered because growing ice took up more and more space,
forcing itself outwards. No known ship could resist that pressure, and no
one knew how thick the ice would get in the middle of the Arctic. The
Fram solved the problem in a brilliantly simple way. She was made to be
chubby and round, only 39 metres long and 11 metres wide. She had a
smooth curvy hull, almost no keel, and engines and rudder that could be
lifted right out of the water. When the ice came, the Fram became a floating
bowl. And if you squeeze a curved shape like a bowl or a cylinder from
below, it will pop upwards. If the squeeze from the ice got too much, the
Fram would just be pushed upwards to sit on top of it – or so went the
theory. She was made from wood that was over a metre thick in places, and
insulated to keep the crew warm. And in June 1893, she left Norway with
tremendous public support and a crew of thirteen, rolling her way around
the northern coast of Russia until she reached the place where the Jeanette
had sunk. In September, she saw ice close to 78°N, and not long after that
she was surrounded. As the ice first trapped her, she creaked and groaned,
but as it expanded around her she rose, shifting upwards exactly as
expected. Frozen in, she was on her way.
For the next three years, the Fram floated with the sea ice, drifting
northward at an agonizingly slow 1 mile a day. Sometimes she went
backwards or round in circles. The fickle freezing ice squeezed and released
her, and she rose and fell in response. Nansen kept his crew occupied with
scientific measurements, but got increasingly frustrated with the slow
progress. When the Fram reached 84°N, it was apparent that she was not
going to get to the pole, 410 nautical miles away. Nansen took a companion
and left the ship, skiing over the ice in an attempt to go where his ship
couldn’t. He set a new record for the furthest north anyone had been, but his
best was still 4° short of the pole. He carried on across the Arctic towards
Norway, meeting a fellow explorer on Franz Josef Land in 1896. The Fram
and her remaining eleven crew stayed the course, carried by the ice to
85.5°N, only a few miles south of Nansen’s new record. On 13 June 1896,
she popped out of the ice just north of Spitsbergen, exactly as originally
planned.
Even though the Fram never reached the pole, the scientific
measurements taken during her journey were invaluable. Now we knew for
sure that the Arctic was an ocean and not a land, that the North Pole was
hidden beneath ever-shifting sea ice, and that there really was a current that
crossed the Arctic between Russia and Greenland. The Fram went on to
carry men on two other great trips. The first was a four-year mapping
expedition to the Canadian Arctic. And then in 1910 she carried Amundsen
and his men to Antarctica, where they would beat Captain Scott to the
South Pole. Today, she sits in her own museum in Oslo, lauded as the
greatest symbol of Norwegian polar exploration. Instead of fighting the
inexorable expansion of the ice, she had used it to ride across the top of the
world.
The expansion of ice as it freezes is so familiar to us that we don’t really
notice it. Put an ice cube in your drink and it floats – that’s just the way
things work. But there’s an easy way to see that the frozen water really is
the same stuff, just taking up more space. If you put some water in a
transparent glass and add some largish lumps of ice, the ice floats so that
most of it is below the surface but about 10 per cent sticks up above the
liquid level. You can mark the liquid level on the outside of the glass with a
marker pen. The question is this: as the ice melts, will the water level go up
or down? Once it’s melted, all those water molecules that are now sticking
up above the water level will have to join the rest of the drink. Does this
mean that the water level will rise? This is proper cocktail party physics, if
you’re patient enough (or bored enough) at a party to spend time watching
ice melt.
The answer is straightforward, and you should test it for yourself if you
don’t believe me. The water level will stay in exactly the same place. Once
the molecules in the ice become liquid again, they can fit together more
closely. This means that they’ll fit perfectly in the hole that the submerged
part of the ice was taking up. That bit of the ice cube that’s sticking up
above the water line is exactly the size of the extra volume that the ice cube
has because it expanded as it froze. You can’t see the atoms themselves in
their lattice, but you can directly see the extra space they need when
fn2
frozen.
Water transforms from liquid to solid in a particular way – the atoms in
the solid each have a fixed location in a lattice. This is called a crystal even
when it isn’t the gleaming centrepiece of a tiara. A crystalline material is
just one that has a fixed repeating structure when it’s solid, like salt or
sugar. But there is another sort of solid, one without this strict positioning.
These solids have a structure more like that of a liquid frozen on its way
somewhere else. Even though the atomic positioning is all happening on a
minuscule scale, and is far too small for us to see, we can still sometimes
see the effect that it has on the object that we pick up. The most obvious
example of this is glass.
I remember seeing glassblowers for the first time on a family trip to the
Isle of Wight when I was about eight. I was spellbound by the smooth
globules of molten glass, glowing and ballooning, constantly shifting from
one beautifully bulbous shape to another. I had to be dragged away because
I would happily have spent all day gazing at this wizardry, the magic of
blobs flowing until they were vases. It was many years before I got around
to what I’d really wanted to do: having a go at it myself. But one chilly
morning this year, my cousin and I arrived at a small stone barn where they
would apparently pull back the curtain and show us how the magic was
done.
It started with the pool of molten glass that sat in a small furnace,
glowing bright orange because it was at the horrific temperature of 1,080°C.
Protected with Kevlar gloves, we obediently poked long iron rods into the
pool and twisted them, so that glass with the consistency of honey wound
on to the iron as we twirled it. That was the easy bit. The hard bit was
everything else. Glassblowing is about controlled coaxing, and there were
three main forms of persuasion that we could apply. Heating the glass up
makes it softer. Holding it still lets gravity conveniently pull it downwards
without you having to touch it. And if the iron rod is hollow, you can blow
bubbles in the molten blob.
We took turns at practising all three, and the astonishing thing about glass
is how quickly its nature changes. As the molten blob comes out of the
furnace, you have to keep spinning the iron because it really is liquid; stop
twirling and it will just drip onto the floor. A couple of minutes after that,
we could roll the blob along a metal workbench and it felt as though it had
the consistency of plasticine. Only three minutes later, you could tap it on
the bench and hear it go ‘ting’, just as you’d expect a solid glass object to
behave. The fun of glass is that you’re manipulating a liquid, playing with
the smoothness and curviness that liquids offer. A solid, cold bit of glass is
just a liquid that was interrupted, frozen in time like a fairytale character.
Glass gets its character from the way its atoms move around each other.
The most common form of glass (and what we were practising with) is
soda-lime glass. It’s mostly silica (silicon dioxide, SiO2, which makes up
the majority of sand), but it’s also got sprinkles of sodium, calcium and
aluminium in it. What makes a glass distinctive is that instead of the atoms
having specific places in a regular lattice, they’re all jumbled up. Each atom
will be linked to the ones around it, and there won’t be too much free space,
but it’s all quite disorderly. As the glass is heated up, the atoms jiggle about
more, moving apart ever so slightly, and since they weren’t in strictly
regimented positions to begin with, it’s quite easy for them to slip past each
other. The molten glass that we took from the furnace was made of atoms
with loads of heat energy, and they would easily slither over each other as
gravity pulled them downward. But as it cooled in the air, the atoms would
move a bit less, settling slightly closer to each other, and the liquid became
more viscous.
The clever bit about glass is that as it cools down, there isn’t enough time
for the atoms to move into an eggbox-like regular pattern. So they don’t.
Glass becomes solid when the atoms are just too sluggish to move over
each other any more. It’s quite hard to say exactly where the line between
liquid and solid really is.
The first task was to make a bauble each, which turned out to be a posh
description for blowing a bubble of glass and then watching the teacher
attach a loop of molten glass to the top. Blowing the bubble was hard work;
my cheeks hurt afterwards as though I’d just inflated a particularly stubborn
balloon. The most delicate part of the process is right at the end, when the
final piece of glass needs to be separated from the iron rod. You pull and
shape the glass so that there’s a thin neck where you want it to break. Then
you file that neck to introduce tiny cracks. And then you take it over to
what was entertainingly called the ‘knock-off bench’, tap on the iron ever
so gently, and the glass bubble breaks off. It all worked perfectly – until we
were on the last one, when the newly introduced cracks weren’t going to
wait. The final bauble dropped off the end of the rod just as it was being
finished, hit the concrete floor, and bounced. Twice. The teacher quickly
scooped it up, and it was fine. But this delicate membrane of glass had
bounced. And apparently if it had fallen just a minute or so later when it
was just a little bit cooler, it would have shattered.
This is the lesson of glass. The way its atoms behave depends on its
temperature. When it’s hot, the atoms can flow freely over each other. Cool
it just enough not to be sticky, and the atoms can press together and rebound
so that the glass can bounce. A little bit cooler, and the atoms really are
frozen in place. Any atom that’s pushed slightly out of place opens up a
crack in a fragile brittle solid, and the glass can be smashed into sharp
smithereens.
Glass is intensely satisfying because it captures the curvy beauty of a
liquid without you having to worry about where the liquid is going. It has
the atomic structure of a liquid – a fairly disorganized mob – but it’s
definitely a solid. The bouncing is a giveaway: elasticity is something that
solids have and liquids don’t. And you can see the consequences of that
structure in how the material behaves as the temperature changes.
This might be the time for a bit of myth-busting about old glass windows.
It has sometimes been said that the reason that 300-year-old windows are
thicker at the bottom than at the top is because the glass has flowed
downward over time. This isn’t true; window glass isn’t a liquid and it isn’t
flowing anywhere. It’s because these window panes were made using an
incredibly ingenious method. A molten glass blob was stuck on an iron rod,
and the rod was spun very quickly until the glass flowed outwards into a
fn3
flat disc. This disc was cooled, and cut up to make window panes. The
downside of this method is that the disc will always be thicker closer to its
centre. So the diamond-shaped window pane pieces were cut with the
thicker bit at one end, and when it was put into the window, the thicker end
was often placed at the bottom to help rain run off. The glass didn’t move
itself downwards, it was put there.
Our glass blobs were not allowed to cool down straight away. They were
put in an oven overnight, one that would bring the temperature down slowly
over the entire night until it matched room temperature in the morning. The
reason for that is that even once the glass is solid, the atoms aren’t
absolutely stuck in place. If you heat something up, the atomic arrangement
changes slightly even if the temperature change isn’t enough to turn the
solid into a liquid. The same thing happened as the glass blobs cooled
down: the atoms shifted slightly. The reason for the oven is to allow this
slight rearrangement to happen slowly and evenly over the whole structure.
If it happened unevenly, unbalanced internal forces could shatter the glass.
Once again, those extra internal stresses are the result of a very simple
principle: the positions of the atoms may be fixed, but the distance between
neighbouring atoms isn’t. If you heat something up, it almost always
expands.
*
The world of digital measuring devices has many advantages, but it has one
definite downside: we’re disconnected from what the measurement really
means. One of the saddest losses is the glass thermometer, an essential tool
in science labs and homes for two and a half centuries. You can still buy
them and I still use them in my lab, but in lots of places they’ve been
superseded by digital alternatives. The shiny mercury thread that I
remember from my childhood has been replaced by coloured alcohol, but
the modern version is essentially the same as the device that Fahrenheit
invented in 1709. There’s a narrow glass rod with a skinny tube running all
the way through the middle of it. At the lower end, the tube widens out into
a bulb, a reservoir of liquid. Put this end of the thermometer in anything –
bath water, your armpit, the sea – and what happens is both elegant and
simple. The temperature of something is directly related to the amount of
thermal energy that it has. In liquids and solids, thermal energy is expressed
as the jiggling about of atoms and molecules. If you put your thermometer
in the bath, you’re surrounding the cold glass with hot water. The molecules
in the water are moving more quickly, so they’ll jostle the atoms in the
glass, giving them the energy to jiggle faster as well. This is heat travelling
by conduction. So when you put the thermometer in the bath water, thermal
energy flows into the glass. The atoms in the glass don’t go anywhere –
they just fidget on the spot, vibrating from side to side. The temperature of
the glass is a measure of this fidgeting, and now the glass is hotter than it
was. Then the atoms in the glass bump into the liquid alcohol until it starts
jiggling faster too. That’s the first part – the bulb of the thermometer is
heated up until it’s at the same temperature as its surroundings.
When atoms in a solid vibrate faster because of extra heat, they push the
nearby atoms away, just a tiny bit. The glass takes up more space when it’s
hot, just because its fidgeting atoms require it. This is why things expand
when they get hot. But the alcohol molecules space out much more as they
speed up; so alcohol expands about thirty times as much as glass for the
same temperature change. Now the alcohol in the thermometer bulb is
taking up more space than it was, but the only extra space is up the tube. So
as the molecules in the alcohol vibrate and push each other apart, the liquid
creeps up the tube. The distance it travels is directly related to the thermal
energy of its molecules, and so the marks on the thermometer correspond to
the amount of thermal energy in the liquid. It’s beautifully simple. When the
liquid in the bulb cools, the alcohol takes up less space as its molecules
slow down. When the liquid heats up, it takes up more space as its
molecules vibrate more energetically. So a reading from a glass
thermometer is a direct measurement of atoms jostling each other.
Different materials expand by different amounts when they’re heated.
That’s why running hot water over a stuck jam jar lid can be helpful: both
the glass and the metal lid will expand, but the metal expands far more than
the glass. After it’s expanded, it’s easier to remove; even though the
difference in its size is far too tiny for you to see, you can feel the result.
Generally, solids expand less than liquids as they’re heated up. The
expansion is only a tiny fraction of the overall volume, but it’s enough to
make a difference. Next time you cross a road bridge on foot, keep an eye
out for a metal strip at either end of the bridge, running across the road. It’s
likely to be made out of two interlocking comb-shaped plates. This is an
expansion joint, and once you know what to look for, they’re pretty
common. The idea is that as the temperature rises and falls, the combs allow
the bridge materials to expand and contract without buckling or cracking. If
the bridge sections expand, the fingers of the comb are pushed further into
each other; if the bridge contracts, the fingers pull back but without creating
a serious gap in the road.
Thermal expansion may be elegant and useful in a thermometer, but it
can have serious consequences on larger scales. One of the problems caused
by our emission of greenhouse gases is that sea level is rising steadily. The
current global average sea-level rise is about 3 mm per year, and it’s rising
more quickly as time goes on. As glaciers and ice sheets melt, water that
was locked up on land is flowing back into the sea, so there’s more water in
the global ocean. But that accounts for only approximately half of the
current rise. The other half comes from thermal expansion. As the oceans
warm, they take up more space. The current best estimate is that 90 per cent
of all the extra heat energy that the Earth has because of global warming has
ended up in the oceans, and the extra sea-level rise is the consequence.
August on the East Antarctic Plateau is still and silent. While the northern
hemisphere basks in summer, Antarctica spins in the darkness on the
bottom of the world. On the high mountain range that stretches right across
the plateau, it’s almost the end of a night that has lasted four months. Very
little snow falls here, but the surface ice is still 600 metres thick. The
weather is calm. Heat energy is constantly leaching away into the starry
night, and there is no sunlight to replace it. This deficit means that along the
high mountain range, the winter temperature is regularly −80°C. On 10
August 2010, one mountainside sank to −93.2°C, the coldest temperature
ever recorded on Earth.
In the ice crystals making up the snow, heat energy is stored as
movement energy as the atoms jiggle about their designated location in the
solid ice. So the answer to the question ‘How cold can it get?’ sounds
straightforward: the coldest possible temperature must be the point when
the atoms stop moving completely. But even at the coldest place on our
planet, where there is no life and no light, there is still movement. The
whole plateau is made up of atoms that are quivering, and they’ve got about
half of the movement energy that they would have just before the ice melts
at 0°C. If you could take away every last bit of that energy, they would be
as cold as it’s possible to get. We have a name for this temperature: absolute
zero, defined as −273.15°C. It’s the same for every type of atom and every
situation, and it means that there is no heat energy at all. Compared to that,
even Antarctica in winter, the coldest place on our planet, looks pretty
warm. Fortunately, perhaps, it’s very difficult to slow atoms down to a
complete stop. It takes a lot of ingenuity to make sure that nothing nearby is
going to give away any of its energy to your sample and spoil it all. But
there are scientists who are devoting their lives to inventing extremely
clever methods of removing heat energy from matter. This is the field of
cryogenics, and it’s opening the door to devices that are useful even in the
nice warm world where we live, especially improved magnets and medical
imaging technology. Most of us, however, find it very uncomfortable even
to think about being really cold. So watching ducks waddle about barefoot
on ice can be very puzzling.
Winchester is a pretty little place in the south of England, with an ancient
cathedral and a colony of very English teashops serving hefty scones on
dainty plates. It can be spectacular in summer, with colourful flowers and a
bright blue sky making the whole place look picture-postcard perfect. But
one year I took a friend there on a snowy winter day, and it looked even
better. Bundled up in scarves and thick coats, we stomped all the way down
to the end of the high street, until we reached the modest river and the soft
blanket of undisturbed snow on its banks. My favourite thing about
Winchester has nothing to do with stone buildings or King Arthur or
scones. What I had dragged my friend all the way through town on a
freezing cold day to see was far more prosaic: ducks. We crunched our way
through the snow for a short distance along the river path, and there they
were.
Just as we arrived, one duck on the bank waddled across the last bit of ice
and hopped into the water. And then it did exactly what all the others
around it were doing: it faced the stream, started paddling like mad, and
reached down to dabble in the water in front of it in search of food. The
river is very shallow at this point, but the water flows very quickly. There
are plants growing on the bottom, just within reach, but the ducks have to
paddle furiously to stay in one place in order to forage. The river in
Winchester is a treadmill for ducks, and I find it endlessly entertaining.
They paddle continuously, all facing the same way, and they never seem to
stop.
A small child next to us looked down at her snow-covered boots, then
pointed to a duck that stood on the ice on the bank and asked her mother an
extremely good question: ‘Why don’t his feet get cold?’ Her mother didn’t
answer because at that moment, the real comedy show began. One of the
paddling ducks had got a bit too close to one of the others, provoking a
burst of quacking and flapping and pecking. The funny bit was that as soon
as the scrap started, both of them forgot to paddle, and so they both
whooshed off downstream with the flow, quacking as they went. After a
few seconds, they realized how fast they were moving, forgot about each
other, and started trying to paddle back upstream to where they had started.
It took a while.
The water was close to freezing, and yet the ducks probably weren’t
feeling the cold. Hidden beneath the water surface, a duck has an extremely
ingenious way of preventing heat loss through its feet. The problem is one
of heat transfer. If you put something hot next to something cold, the faster,
more energetic molecules in the hot object will bump into the molecules in
the cold object, transferring heat energy from the hot object to the colder
one. This is why heat flow always has to go that way – molecules vibrating
slowly can’t give energy to those vibrating faster, but it’s easy in the other
direction. So heat energy is generally shared out until everything is the
same temperature and equilibrium is reached. The real problem for the
ducks is the blood flow in their feet. It comes from their heart, in the nice
warm centre of the duck, so it’s at 40°C. If that blood gets anywhere near
the freezing water, there is a big difference in temperature, so it will lose its
heat to the water very quickly. Then when it gets back to the body of the
duck, the warm duck will give its heat away to the cold blood and the whole
duck will cool down. Ducks can restrict blood flow to their feet a bit, so less
blood is at risk of getting cold, but that doesn’t solve the problem
completely. They use a much simpler principle. It’s this: the bigger the
temperature difference between two objects when they touch, the faster heat
will flow from one to the other. Another way of putting it is to say that the
closer the temperatures of the two objects are, the more slowly heat will
flow from one to the other. And that’s what really helps the ducks.
As all that frantic paddling was going on, warm blood was flowing down
the arteries of each duck’s legs. But those arteries were right next to the
veins carrying blood back from the feet. The blood in the veins was cool.
So the molecules in the warm blood jostled the blood vessel walls, which
then jostled the cooler blood. The warm blood going to the feet got a bit
cooler, and the blood going back into the body was warmed up a bit.
Slightly further down the duck’s leg, the arteries and the veins are both
cooler overall, but the arteries are still warmer. So heat flows across from
the arteries to the veins. All the way down the duck’s legs, heat that came
from the duck’s body is being transferred to the blood that’s going back the
other way, without going near the duck’s feet. But the blood itself goes all
the way around. By the time the duck’s blood reaches its webbed feet, it’s
pretty much the same temperature as the water. Because its feet aren’t much
hotter than the water, they lose very little heat. And then as the blood travels
back up towards the middle of the duck, it gets heated up by the blood
coming down. This is called a countercurrent heat exchanger, and it’s a
fantastically ingenious way of avoiding heat loss. If the duck can make sure
that the heat doesn’t get to its feet, it has almost eliminated the possibility of
losing energy that way. So ducks can happily stand on the ice precisely
because their feet are cold. And they don’t care.
This strategy has evolved many times separately in the animal kingdom.
Dolphins and turtles have a similar layout of blood vessels in their tails and
flippers so that when they swim into colder water, they can maintain their
internal temperature. It’s also seen in Arctic foxes – their paws have to be in
direct contact with ice and snow, but they can still keep their vital organs
warm. It’s very simple, but very effective.
Since my friend and I didn’t have any way of playing the same trick, we
lasted a limited amount of time out in the snow. After watching a few more
high-speed duck squabbles, and expressing suitable admiration for what
must be the fittest ducks anywhere in the world, we went in search of giant
scones.
ONE OF THE nice things about bubbles is that you know where to look for
them: at the top. They’re either on their way there, wobbling upward
through fish tanks or swimming pools, or nestled in with the crowd on top
of champagne or beer. Bubbles reliably find their way to the highest point
of the liquid they’re in. But next time you stir a mug of tea or coffee, have a
look at what’s going on at the surface. The first odd thing that happens is
that as you move the spoon around in circles, the surface of the tea develops
a hole. As the liquid whirls around, the middle of the tea sinks and the
edges rise up. And the second odd thing is that the bubbles in the tea are
spinning quietly at the bottom of the hole. They’re not at the highest point
of the tea, at the edges. They’re hiding at the lowest point on the surface,
and they stay there. If you push them away, they find their way back. If you
make new bubbles at the edges, they spiral into the centre. Odd.
When I start to stir my tea, I’m pushing on the liquid with the spoon. I
push it forwards, but there’s only so far that it can go before it meets the
side of the mug. If I did the same thing with a spoon in a swimming pool,
the water in front of the spoon would move forward, and it would keep
going forward until it mixed in with the rest of the pool. But in the mug,
there’s no room for that to happen. Even though the side of the mug isn’t
going anywhere, it can still push back on any liquid that bumps up against
it. It’s a wall, and tea can’t pass through it. Since the tea can’t go in a
straight line, it starts to move around the cup in a circle. But as that’s going
on it’s piling up against the walls because only the side of the mug can push
back. The tea will keep trying to go in a straight line, and it only moves
around in a circle because it’s being forced to curve.
This is the first lesson about spinning things. If you suddenly freed them
of their restriction, they would just keep moving in the same direction they
were going in at the moment of release. Imagine a discus thrower, spinning
around while holding on to the discus. After a few rotations, the discus is
going incredibly quickly, but it stays on its circle because it’s being tightly
held. The athlete must continually pull it back towards the centre of the
spin, and that pull is along the line of his or her arm. The second that they
let go, the discus travels forward in a straight line, with exactly the direction
and speed it had before the release.
As I’m stirring my tea, the hole develops because each bit of tea tries to
move in a straight line, but that makes it push up against the cup sides and
so there’s less left in the middle. When I stop stirring, the hole remains
because the liquid is still spinning. As the whirling slows down, it takes less
of a push to keep the tea going in a circle, and so there’s less of a pile-up at
the sides. You can see all this in a liquid because it’s free to move, so it can
alter its shape.
And at the centre of the circles, the bubbles are spinning away. What
their presence in the middle tells you is that it’s the least favourable place to
be. When a glass of beer is sitting on a table, the bubbles move to the top
because the beer is winning the competition to get closest to the bottom.
And it’s the same for the mug of tea. The bubbles are in the middle because
the tea is winning the competition to move out to the sides. The liquid is
more dense than the gas, so the gas drifts into the space left over.
Our civilization is full of things that spin – clothes dryers, discus
throwers, flipped pancakes and gyroscopes. The Earth itself spins as it
circles the Sun. Spinning is important because it lets you do lots of
interesting things, sometimes involving enormous forces and oodles of
energy, all without actually going anywhere. The worst that can happen is
that you end up back where you started. The bubbles in the tea are just the
start. The same principle also explains why you wouldn’t launch a rocket
from Antarctica and how doctors measure whether you have enough red
blood cells. Spinning could also play an important role in our energy grid in
the future. All of those possibilities come from a restriction: the one thing
you can’t do when you’re spinning is to travel in a straight line.
If you’re zooming around in a circle, there must be something either
pulling or pushing you inwards, forcing you to change direction constantly.
That applies to anything that’s spinning, whatever the situation. If that extra
force is taken away, you just continue in a straight line. So if you want to
travel in a circle, you must have something to provide an extra, inward
push. The faster you’re going, the stronger that push has to be, because the
faster you need to curve, the more force it takes. Spectator sports love a
good racetrack; they have the same benefit as anything else that spins. You
can achieve huge speeds without really going anywhere, and certainly not
anywhere the paying audience can’t see you. To make sure that they get
enough of an inward push to stay on the track, some sports have taken
racetrack building to fairly extreme lengths. Indoor track cycling is the
prime example of this. But it wasn’t the lengths that terrified me when I
tried it … it was the steepness.
I’ve been an enthusiastic cyclist all my life, but this was a very different
kettle of fish. The inside of London’s Olympic velodrome is shiny and vast
and oddly quiet. You pop up into the middle of the stillness, and they issue
you with a lean, mean-looking bike with a single gear, no brakes and the
most uncomfortable saddle I’ve ever had to sit on. When the group for the
beginners’ session was assembled, we clopped outwards to the track and
held on to the rail while we clipped into the pedals. The track seemed
enormous. There are two longer straight sides, and then the sweeping
banked sections at each end that towered over us. They are so steep (43° in
places) that it seems as though the designer had really wanted to build a
wall. Cycling looked like entirely the wrong thing to be doing here. But it
was too late for our little group now. The track was waiting for us.
First of all, we were sent around the flat oval that sits inside the main
track. The surface was beautifully smooth, and the bikes made a lot more
sense. Then we were instructed to venture outwards on to the light blue
strip with the first slight gradient. And then, feeling slightly like baby birds
being pushed out of their nest to learn to fly, we were told to face the main
track.
There was a nasty surprise straight away. I had thought the banking
would be gradual, but it isn’t. The gradient at the bottom is pretty similar to
the gradient at the top. As soon as you stray outwards on to the race surface,
you’re cycling across a pretty significant slope. Pedalling faster seemed like
a good idea, but that was only because I was forcing my brain to let logic
make the decisions while it busily pretended that instinct didn’t exist. I
forgot how stupidly uncomfortable the saddle was after the first three laps.
Round and round we went, like demented hamsters on a gigantic wheel,
pausing occasionally so that the instructors could check on us. Twenty-five
minutes in, I was still terrified, but I was learning.
The game here is that you want the bike to lean inward so much that it’s
perpendicular to the track. The only way to do that without slipping down
the slope is to be going very quickly, because then you’re like the swirling
tea. The bike wants to continue in a straight horizontal line, but it can’t
because the curved track is in the way. That push-back from the track is
providing the inward force that keeps you going in a circle. The bike is
pushing into the track so hard that when you add that push to gravity, it’s as
if gravity has changed direction. Now you’re being pulled into the track,
instead of downward to the centre of the Earth. The faster you cycle, the
more you change the direction of the effective gravity. It still feels as
though you’re cycling on a wall, but at least it’s one that you’re glued to by
something that feels familiar.
I understood the theory, but the practice was a bit different. For a start,
there’s no resting. You can’t stop pedalling because you can’t freewheel. If
the wheels are going round, your legs are going round and that’s just the
way it is. On a few occasions, I instinctively paused as I would if I needed a
couple of seconds’ rest on the road, and I was rewarded by a gigantic kick
of adrenaline as the bike bucked me up out of the saddle. You can’t
freewheel on these bikes at all. You have to keep going, however much your
legs are burning. If you slow down, you slide down the bank. I felt some
new respect for the athletes who do this on a regular basis. And then there
are all the other people on the track to contend with. If you move up to
overtake someone, you are taking a longer path, so you have to increase
your speed a lot even to have a chance. I was quite happy not to do too
much overtaking.
The lesson from all this was that if you’re doing things right, the steeper
slopes will give you a stronger inward push. And the reason that you need
that push on the ends but not along the sides is that the semicircular ends
are where you’re changing direction. The more quickly you change
direction, the bigger the push you need to make it happen. If you tried to
cycle this quickly on a flat track of the same shape, you’d skid out to the
sides – tyre friction alone can’t supply that inward push. The velodrome is
what happened when the cycling world refused to allow its need for indoor
speed to be limited by friction.
If you’ve ever wanted to know what it feels like to be a penny rolling
down a whirlpool-shaped charity donation funnel, this is the way to do it.
At the end of an hour I was properly fired up on adrenaline, and really glad
fn1
that it was time to stop. The scary thing about the effective gravity pulling
me into the track was the knowledge that if I slowed down suddenly, it
would change back. And gravity pointing downwards is a pretty
uncomfortable thought when you’re cycling on a 43-degree wall.
The cyclist is being pushed inwards by the track in the same way that the
ground is pushing up on us all the time. If the ground underneath you
suddenly vanished, you would fall because gravity is pulling you
downwards. So the ground itself is pushing back on us, to counter gravity’s
downward tug. Cyclists feel the track both pushing them upwards and
pushing them inwards. Overall, that will feel as though gravity is pulling
them down and outwards.
There’s a track cycling event appropriately called the ‘flying 200-metre
time trial’. I reckon it must feel like flying, even though it’s called that
because they’re already up to speed before the clock starts. The world
record at the time of writing was set by François Pervis and it’s 9.347
seconds. That’s 21 metres every second, or nearly 48 mph. For him to spin
around the end of the track at that speed, the track has to be pushing him
inwards almost as hard as the floor is pushing him upwards. François was
glued to the track by a force almost twice as strong as normal gravity.
As we saw in chapter 2, a constant background force like gravity is
useful for all sorts of things, although some of them (like separating out
cream) take ages. But spinning offers us an alternative. You don’t need to
move to a new planet to reap the benefits of increased gravity. Cyclists can
almost double their effective gravity at the top of a track, but even the best
track cyclists in the world can ‘only’ get up to about 50 mph. In theory, you
could just keep spinning faster and the forces on you would keep getting
stronger.
Remember gravity helping the droplets of cream to separate from the rest
of the milk and rise to the top of the bottle in chapter 2? If the force pulling
the milk downward is only as strong as gravity, it takes a few hours for the
fat droplets to separate out. But if you put the milk in a long, spinning tube
and whizz it around very quickly, the outward pull is so strong that the
cream droplets will separate out in just a few seconds. This is how all our
cream is separated from the milk these days – they don’t just let it sit and
wait for it to sort itself out. Modern food production hasn’t got the time for
that. Spinning something around generates a pull that can be as strong as
you like, as long as you can spin fast enough. This is what a centrifuge is: a
spinning arm that can hold on to something, pulling it inwards to make it
spin, and making the object feel as though it’s being squashed against the
outer side by a very strong force.
You can make those internal spinning forces so strong that things that
would never separate out under gravity alone can be teased apart. For
example, if you ever have a blood test for anaemia, laboratory technicians
will put a sample of your blood in a centrifuge and spin it so fast that it
experiences an outward force perhaps twenty thousand times as great as
gravity. Red blood cells are far too small to separate out under gravity in
any normal circumstances, but they can’t resist the forces generated by the
centrifuge. Under those conditions, it takes only five minutes for almost all
the red blood cells to be pulled outwards from the centre of the centrifuge,
towards the bottom of the tube. They’re more dense than the liquid they’re
in, so they win the race to the bottom. Once they’re all there, the tube can
be lifted out and it’s possible to directly measure the percentage of your
blood that is made up of red blood cells, just by measuring the thickness of
the lowest layer. This is a simple test that can indicate a range of health
problems, and it’s also used to check for blood doping in athletes. If it
wasn’t for the forces generated by spinning, this measurement would be
much harder and much more expensive to carry out. And these forces can
be applied to much larger things than blood samples. One of the biggest
centrifuges in the world is designed to spin an entire human being.
Many people envy astronauts for their adventures: their amazing views of
our home planet, all the technical toys they get to play with, the
accumulation of fabulous stories to tell, and the accolades of having one of
the rarest and most hard-earned job descriptions in the world. But ask most
people what they envy the most and you almost always get the same
answer: weightlessness. All that floating about without ‘up’ or ‘down’
being a problem sounds both highly exciting and very relaxing. So it might
seem slightly strange that astronauts in training need to be just as well
prepared for the opposite problem: forces that far exceed gravity. The only
current way to get to space is to sit on top of a rocket that’s accelerating
pretty quickly. And it’s even worse on the way back down: re-entry into the
Earth’s atmosphere can generate forces four to eight times as strong as
gravity, the sort of forces that a fighter pilot doing tight turns at high speed
might have to deal with. If you feel slightly queasy as a lift accelerates, this
might not be for you. Depending on the direction of the additional g-forces,
extra blood will be pushed towards or away from your brain, possibly even
bursting the tiny capillary blood vessels in your skin. The details aren’t
necessarily pleasant. But humans can not only survive these forces, they can
also work while subjected to them (as you have to if you’re piloting a
spacecraft back to Earth), and they do both better if they’re used to it. So a
way has been found to train them.
All current astronauts and cosmonauts will spend considerable periods of
time at the Yuri Gagarin Cosmonaut Training Centre in Star City, just to the
north-east of Moscow. Among the lecture halls, medical facilities and
spacecraft mock-ups sits the TsF-18 centrifuge. From the centre of a huge
circular room, the arm of the centrifuge stretches 18 metres outwards. The
capsule at the end can be changed depending on what’s needed on any
given day. The tests that any budding astronaut has to pass involve sitting in
the capsule while the arm rotates once every two or four seconds, which
doesn’t sound like much until you calculate that the capsule itself must
therefore be travelling around at either 120 or 60 mph. Once they’ve shown
that they’ve got the right stuff, astronauts can practise working in these
conditions, and are constantly monitored to check on how their bodies are
responding. And it’s not just astronauts – test pilots and fighter pilots can
also train here. The centre even offers the experience to members of the
public who can afford to pay for it. Be warned, though: the only thing about
it everyone seems to agree on is that it’s very uncomfortable. But if you
want to experience a consistent very high force, taking a spin is the way to
do it.
The centrifuge is one way of exploiting the forces generated when
something spins: by taking advantage of the ability to generate a very
strong force in a single direction, and treating it like artificial gravity. But
there is also a second way of employing forces from spinning. The tea and
the cyclist and the astronaut were all confined – they were all forced to
move in a circle because there was a solid barrier pushing back on them,
preventing them from moving outwards. But what if you’re spinning and
there’s nothing external to trap you on a fixed circular path? This is a pretty
common scenario. Rugby balls, spinning tops and frisbees all spin without
anything external pushing them inwards. But the best way of seeing what’s
going on is far more fun, and also edible: pizza.
To my mind, the perfect pizza should have a thin crispy base, the vital but
understated foundation that lets the toppings shine. Raw pizza dough starts
out as a rotund blob, a living lump that needs to be kneaded and nurtured to
bring out the best in it. Transforming the blob into a delicate sheet without
breaking it is an essential skill for a pizza-maker, and some go a step
further, taking the basic skill and turning it into theatre. The chefs who toss
pizzas have mastered the trick of letting the spinning do the work for them.
Why push and prod each part of the dough with your fingers when you can
just let physics sort out all those messy details? Especially when the flying
disc gives you the mysterious aura of a dough wizard.
Tossing pizza dough has evolved into a proper spectator sport of its own;
there’s now a world championship every year. There are even those who
call themselves ‘pizza acrobats’, whose party piece is keeping a constantly
spinning pizza base (or two) flying and somersaulting around their body for
several minutes at a time. No one seems to eat pizza made from such well-
travelled dough, but it definitely looks impressive. However, there are
plenty of pizza chefs out there who spin their pizza dough briefly without
making a cabaret act of it, and who have every intention of turning it into
someone’s dinner. What is the spinning actually doing?
Some pizza-mad friends of mine recently took me to a very friendly
restaurant with an open kitchen, and I asked whether I could watch
someone spinning pizza dough. The young Italian chefs giggled a bit, but
then gathered around the one who was brave enough to volunteer. Half
embarrassed and half proud to show off, he patted a ball of dough to flatten
it slightly, then picked it up and with a slight flick of the wrist, sent it
twirling into the air.
What happened next happened very quickly. As the circle of dough left
his hand, it was suddenly free of anything external pulling and pushing on
it. It’s helpful to think about a single point on the edge. It’s travelling
around in a circle, but only because the rest of the dough is stuck to it,
pulling it inwards. That inward pull is always necessary for something to
rotate. In the case of the cyclist, the track is constantly pushing the bike
from the outside, so that the cyclist must curve inwards towards the centre
instead of continuing on a straight line. For the pizza dough, it’s a pull from
the middle that makes the edge of the dough curve around towards the
centre. Either way, there has to be a force directed towards the middle of the
spin. But dough is soft and elastic. If you pull on it, it stretches. The middle
of the dough is pulling the edge inwards, but that means there’s a pulling
force across the dough. And so the dough has to stretch. When any solid
object spins, the spinning generates forces inside it that you can’t see. The
internal pull that’s keeping the pizza together is also stretching the dough,
and the edge is getting further and further from the centre. The brilliance of
this for a pizza chef is that the internal pull is smooth and symmetrical. All
of the pizza is spinning, so all of it stretches out away from the centre.
You can feel these internal pulling forces yourself sometimes. If you hold
a bag containing a reasonably heavy object out horizontally, and spin
yourself around, you’ll feel a pull trying to stretch your arm. This is the
inward pull that’s keeping the bag spinning in a circle. Fortunately for you,
your arm is much less stretchy than pizza dough, so it stays the same length.
But the longer your arm and the faster you spin, the more of a pull you’ll
feel.
So as the pizza dough was spinning in the air, the same pull that was
keeping the edge moving around in a circle was gradually stretching the
dough outwards. I reckon the dough was in the air for less than a second,
but it was a very thick pancake when it went up, and a beautiful thin smooth
circle when it came down. The chef kept it spinning and sent it up again,
but this time the internal pulling forces were so strong that the dough tore
itself apart in the middle, and what came down was sad and raggedy. The
chef grinned sheepishly. ‘That’s why we don’t normally do it,’ he said. ‘The
dough that makes the best pizza is too soft to spin, so we have to stretch it
fn2
by hand on the board.’ It turns out that the dough used in the acrobatic
competitions is made using a special recipe so that it’s stretchy and strong,
but doesn’t necessarily produce the cooked pizza with the best texture.
Right at the edge of a pizza, the internal pulling force could be five to ten
times as strong as gravity, which is why the pizza base stretches much more
quickly when you spin it than it does if you just hold it up and let it droop
under its own weight.
A spinning pizza base is lovely to watch because it changes shape in
response to forces that are entirely hidden away inside itself. Spinning
anything generates a pulling force from the centre to the edge – the same is
true in a spinning rugby ball or frisbee – but you’d never know about it in
these solid objects because they’re strong enough to resist being stretched.
Or at least, they’re stretching so little that you can’t tell. But everything will
stretch a tiny bit. Even the Earth itself.
Our planet is constantly spinning as it travels around the Sun. And, just like
the pizza dough, it’s stretched by the forces that are pulling each bit of it
back inwards, keeping each bit of rock travelling in a circle. Fortunately for
all of us, gravity is strong enough to prevent any consequences as extreme
as they are for the dough, and the Earth stays fairly spherical. But it does
have what is helpfully called an ‘equatorial bulge’, which sounds like a
euphemism for having eaten too much cake. If you stand at the equator,
you’re 21 km further away from the centre of the planet than someone at
the North Pole is. Our planet is held together by gravity but shaped by its
spin. And so even though Mount Everest is the tallest mountain on Earth,
the top of Everest is not the furthest point from the centre of the Earth. That
accolade goes to Chimborazo, a volcano in Ecuador. Its summit is only
6,268 metres above sea level (Everest is 8,848 metres tall by the same
measure), but it’s sitting right on top of the equatorial bulge. So, when
you’re standing on top of Chimborazo, you’re a little over 2 km further
from the centre of the Earth than anyone who has just struggled to the top of
Everest. Pointing that out when you both get home probably won’t make
you popular, though.
Overall, the forces generated by spinning can be useful in two ways. The
pizza is one – spinning something without confining it generates a pull
inside the object as it tries to hold itself together while it spins. The cyclist
is the other – if you do put a wall in the way, confining whatever is spinning
with something that pushes back, you can generate a strong consistent
gravity-like force on that object. But the common theme is that an inward
pull or push has to come from somewhere. If that inward force ever goes
away, the object can’t stay on its circular path.
Only a solid object can hold itself together like the pizza dough. Liquids
fn3
and gases aren’t stuck together like that. This distinction is fantastically
useful if you have both solid objects and liquids mixed together, because
now you can separate them out. The brilliance of a spin dryer for clothes is
that the clothes are trapped inside the drum, and the drum is pushing them
inwards so they have to keep going round in circles. But the water tucked
away in the clothes isn’t held in position. Since it’s free to move, it can keep
moving outwards through gaps in the material. It will only travel in a circle
if it gets an inward push from something solid. Otherwise, it will gradually
wriggle its way away from the centre, and when it meets a hole in the drum,
it’ll go flying out sideways, free of the circle completely.
When you spin something around and then let go, you start by pulling on
it with exactly the right inward force to keep it going round in a circle, and
then you suddenly take that force away. When there’s no inward pull,
there’s no reason for the object to keep going round in a circle. So it just
sails off in a straight line. This principle revolutionized medieval warfare in
Europe and the eastern Mediterranean, enabling engineers to build giant
siege engines that could batter down stone fortresses. And I have used it to
launch wellies, but not quite as effectively.
At the end of my PhD viva, just after they’d told me that I’d passed, the
external examiner smiled across the table and asked what I was going to do
with the rest of the afternoon. He was clearly expecting parties and pubs
and inebriation to be high on the agenda. He was not expecting me to say
that I was about to go cycling out into the Cambridgeshire countryside to
see whether I could find a farmer who might lend me an old tractor tyre or
two. I explained that I was making a contraption to throw wellies with, that
I had to make it out of scrap material, and that I had to finish it within the
next week. The examiner’s brow wrinkled, his eyebrows waggled
uncertainly, and then he smoothly pretended he hadn’t heard and asked me
what plans I had for a job. But it was true. I had agreed to be part of a rare
all-female team taking part in a Scrapheap Challenge roadshow, and the
challenge was to build something that could compete at throwing
wellington boots, to be put through its paces at the Dorset Steam Fair. There
were three of us, we had no money and very little time, and as far as I could
see the only option available was to use an ancient and very effective
technology: the trebuchet.
A trebuchet is an extremely ingenious device that was developed over
many centuries with the input of several civilizations: the early Chinese, the
Byzantine and Islamic empires, and finally western Europe. When it came
of age in the eleventh and twelfth centuries, it proved itself a monstrous
lumbering brute, capable of demolishing castles that had previously been
thought impregnable. A trebuchet could hurl 100-kg rocks over hundreds of
metres. Siege engines like this contributed to the disappearance of motte-
and-bailey castles (strategically useful but made only of wood and earth).
Solid stone was the only defence, and so stone fortresses became the norm.
The benefits of the trebuchet were the same for me and my team as for
the medieval warmongers: it’s mechanically simple and extremely effective.
We borrowed scaffolding poles from a local building site, dug around in the
college skip for stuff to make a sling from, persuaded the technicians in the
Cavendish Laboratory to let me have a 5-metre-long metal beam, gathered
all this stuff together at the top of the college playing fields and set to work.
Churchill College in Cambridge had been my home for nearly eight years
by then, and the college staff were used both to me and to the sudden
appearance of novel contraptions. Looking back, I’m still astonished at (and
extremely grateful for) the cheerful acceptance that met any of the students
whenever we had a new daft idea. At the opposite end of the playing field
that week, someone else was testing a stratospheric balloon to send a teddy
bear into space.
The basic structure of a trebuchet is very simple. You build a frame that
gives you a pivot point which is perhaps 2 or 3 metres off the ground. Then
you attach one long beam to that, like a giant see-saw, but you position the
pivot point so that there’s much more of the beam on one side than the
other. Now you’ve got an A-frame that looks as though it’s got a long stick
laid across the top. The long end is the one that starts touching the ground.
You attach a sling to the long end, and lay the sling on the ground
underneath the frame. The first time we assembled it all, it was a beautiful
sunny day, perfect for launching anything.
Then we hit a problem. The beautiful thing about a trebuchet (unless
you’re the one that’s about to get a rock thrown at you) is that it uses
gravity to spin the see-saw and the sling. You attach a heavy weight to the
short end of the see-saw, and then as you drop the weight, it pulls the see-
saw on your side down very quickly. The whole beam spins around the
pivot, tracing out a vertical circle, and the sling also spins around the other
end of the beam. So you’ve got lots of very fast rotation, and the projectile
in the sling is spinning around the pivot because it’s being pulled inwards
by the sling. So far, so good. The first task was to get to this point, but we
couldn’t find a weight that was heavy enough to move everything. I offered
to swing from the beam myself as a human weight, but even I wasn’t quite
heavy enough. We were stumped. That night, I spent a while pouring out
my frustration to another set of friends, batting away their suggestions that I
should just eat more cake. Then one of them offered me his scuba-diving
weights. So the next day, I rigged myself up with a belt carrying 10 kg of
diving weights, and we tried again. This time, it worked perfectly. I swung
under the pivot, the seesaw swung over the top, and the sling swung over
the top of that. Everything was spinning. Now it was time for the next step.
The sling is only held in place by a small loop, and the trick to it all is
that when the sling is almost at its highest point, the loop slips off. The sling
is effectively broken. That means that the force that was pulling the
projectile inwards and keeping it on the circle has vanished. Now the
situation has changed. At this moment, the projectile in the sling is
travelling forwards and upwards very fast. As soon as it’s free of the inward
force, it just keeps going in a straight line. Since it was travelling forwards
and upwards before, it keeps going forwards and upwards. But it doesn’t go
directly outward from the centre of the spin. It carries on going sideways, as
if following a line that sat on top of the spin circle. That was the theory. We
put a shoe in the sling and lined everything up. I faced away from the
playing field and swung down on the see-saw. The other end of the see-saw
swung up, dragging the sling up, around and over the pivot. At exactly the
right moment (first time!) the sling released, and the shoe went flying over
the top of my head, out on to the playing field. I wouldn’t ever want to do
that with a rock, but the shoe proved the point perfectly. Our contraption
could at least throw a wellington boot, and in the time we had, that was the
best we could do. After a bit more practice, we took our frame apart ready
to transport it to the competition the next day.
Arriving at the Dorset Steam Fair poked a large pin in our buoyant
bubble of confidence. Every other team consisted of middle-aged men who
had spent months in garages building gloriously decorated contraptions for
throwing wellies. Our small pile of scaffolding poles and discarded carpet,
collected over just a few days, looked sparse and unloved. But we put a
brave face on it and put it all back together. The competition officials (also
middle-aged men) came round to look at it. ‘It’s silly to swing from it,’ one
said. ‘You should do what the medieval warriors did, and just pull the lever
down with a piece of rope. That’ll work far better.’ My protests that the
counterweight was the invention that had led to the success of this device
went unheard. The reason it never made it as a powerful siege weapon
before the eleventh century is precisely because people tried to do it by
manpower. But the officials stuck their hands in their pockets, opined that
pulling on a rope was a far better idea, hinted that we enthusiastic but
inexperienced females should be grateful that we were getting extra help
from them, and didn’t leave until my team-mates had given in and agreed
with them. There was no time to argue. The time of the competition had
come.
The first challenge was to throw as many wellies as possible over a line
about 25 metres away in two minutes. The top five teams would then carry
on to the next stage of the competition, which was to see who could throw
the furthest overall. The clock started. The three of us heaved on the rope,
turning the see-saw and flinging the sling. But the first welly barely cleared
our heads. We couldn’t pull downwards fast enough to make the see-saw
rotate properly. We tried again. And again. After about a minute, I
convinced my team-mates that this wasn’t working, and we set ourselves up
for the original idea. I put my diving weights on, jumped off the small filing
cabinet we were using as a platform, swung underneath the pivot, and
wheeeeee … the first welly went soaring over my head and over the line.
Next one. Welly in the sling, up on filing cabinet, swing down, whoosh!
Next – but the whistle had gone. Our time was up. Two wellies over the line
wasn’t enough. We wouldn’t get through to the next stage. The middle-aged
men commiserated with us. Better luck next time. I hid from the official
who had suggested the rope, because I was so cross with him. It worked!
Our simple structure of scaffolding and carpet and elegant physics worked,
and it had worked the way I’d said it would. We could have competed
against the lovingly painted, intricately made garage beauties! But we had
fn4
been scuppered by the last-minute change of plan. Most of the other
competition entries were based on far less efficient methods. They might
have been colourful, but we had physical efficiency and simplicity on our
side.
So my personal success with trebuchets is a bit limited, but eight hundred
years ago this elegant idea revolutionized warfare. Being able to throw
heavy rocks with high accuracy meant that you could thump the same bit of
the castle wall again and again, until it yielded. For two centuries or so
trebuchets got bigger and better, and were given names like ‘God’s stone
thrower’ and ‘Warwolf’. Each one took vast amounts of timber to construct,
but being able to chuck another 150-kg rock at your enemies every few
minutes was worth it. Spinning the rock and the sling around an axis lets
you build up to a very high speed in a very short space of time. You don’t
want the spinning to continue – you’re just using it as a way of reaching a
high speed. Once the projectile is going fast enough, you remove the inward
force at the moment when its direction is perfect. And off it goes, soaring
out in exactly the direction it was released in. Until gunpowder became
reliable enough for the cannon to be more of a weapon than a liability, in
terms of destructive efficiency a trebuchet was as good as it got.
Lots of things are spinning. For example, right now, you and I are spinning.
We’re going round the axis of the Earth once a day, although we can’t feel it
because the Earth is so big that we’re only changing direction slowly. If we
were at the equator, our sideways speed would be 1,040 mph. In London,
where I’m writing this, I’m speeding sideways at 650 mph because we’re
closer to the spin axis. But if we all live on a massive spinning planet, and if
a loose object at the surface of a spinning thing will whizz off in a straight
line when you let go of it, why are we all still down here? The answer is
that the inward pull of gravity is strong enough to prevent the planet letting
go of us. In fact, even when you’re in orbit, the planet hasn’t actually let go
of you. And when you’re on your way up there, that extra speed you’ve got
because of the Earth’s spin can be extremely useful.
On 4 October 1957, a diminutive metal sphere named Sputnik chirped
out the first sounds of the Space Age, and the world listened open-mouthed.
Earth’s first artificial satellite was a huge technological achievement.
Sputnik orbited its home planet once every ninety-six minutes, and each
time it went past anyone with a short-wave radio could hear its distinctive
‘peep … peep … peep …’ America had woken up that morning happily
complacent in the knowledge that it was the greatest nation on Earth. It
went to bed shocked that maybe it wasn’t. Within a year, the Soviets had
sent up Sputnik II, a bigger satellite carrying a dog named Laika. The
panicked Americans hadn’t sent anything into space, but they had launched
NASA, the National Aeronautics and Space Administration. The Space
Race had begun in earnest.
But what was the real achievement of Sputnik? It wasn’t just about going
up; anything close to something as big as a planet has to live up to the
maxim ‘what goes up must come down’. The trick of putting satellites into
orbit starts with what goes up, but the real skill is in delaying their coming
down for as long as possible. Sputnik hadn’t escaped the Earth’s gravity.
That wasn’t the point. Douglas Adams summed it up perfectly, and
accurately, with the minor caveat that he was talking about flying and not
orbital space flight: ‘The knack lies in learning to throw yourself at the
ground and miss.’ Sputnik was permanently falling towards the Earth. It
just kept missing it.
Sputnik was launched from the deserts of Kazakhstan, a place that is now
the site of the Baikonur Cosmodrome, a vast space launch facility. The
rocket that carried Sputnik powered upwards through the thickest part of the
atmosphere and then turned sideways, accelerating horizontally around the
curve of the Earth. By the time the last parts of the rocket fell away, Sputnik
was whizzing around the planet at about 8.1 kilometres every second, or
18,000 mph. This is where the effort goes when you’re getting into orbit –
it’s mostly about going not upwards, but sideways.
The little metal sphere hadn’t escaped gravity at all. In fact, it needed
gravity to be there, to make sure that it stayed in orbit and didn’t just carry
on and leave the Earth behind. As it zoomed along at this fantastic speed,
the Earth was pulling it downwards with almost as much gravity as there is
fn5
on the ground. But because Sputnik had such a huge sideways speed, by
the time it had fallen a little way down towards the Earth, it had gone so far
forward that the Earth had curved away beneath it. And as it kept falling, so
the Earth’s surface kept curving away. This is the beautiful balance of being
in orbit. You’re going sideways so quickly that you fall towards the ground
and miss. And because there’s almost no air resistance, you can just keep
falling and missing, as you go round and round.
To get into orbit, you have to be going sideways fast enough to make this
balance work. And Kazakhstan already has a pretty significant sideways
speed because it’s whizzing around the Earth’s axis once a day. The further
you are from the spin axis, the faster your sideways speed is. So by
launching from somewhere close to the equator, you’re giving yourself a
pretty significant head start. A sideways speed of about 8 kilometres per
second is needed to make low-Earth orbit work. Kazakhstan is whizzing
sideways at about 400 metres per second (894 mph). So if you launch
eastward, with the spin of the Earth, just starting in Kazakhstan instead of at
the North Pole means that 5 per cent of the work is done for you.
In the spin dryer, the outside of the drum pushes the clothes inwards, so
they can’t escape. In the velodrome, it was the alarmingly sloped track that
was pushing me inwards. And for Sputnik, the tiny peeping herald of
humanity’s first adventure into space, gravity was doing that job.
Everything that spins needs something to be pulling or pushing it towards
the centre of spin all the time. For the clothes in the spin dryer just as for
Sputnik, if that force had vanished, they would have kept going in a straight
line.
So gravity definitely still matters a few hundred kilometres above our
heads. But surely the whole point of being in space is that you get to be
weightless. What about all those astronauts drifting about in zero gravity,
trying desperately not to spill anything because it’ll float around for days?
Today, the International Space Station is orbiting above our heads. The
astronauts who live on board this huge scientific facility proudly state that
they are flying on particular missions, and I don’t begrudge them that. It
sounds a lot less exciting to say that you’re going to spend six months
falling. But they’re not flying and they are falling. Just as Sputnik was
falling towards the ground and missing, so are the astronauts and the space
station.
While you’re in free fall, you can’t feel gravity because there’s nothing
pushing back. Since the astronauts can’t feel anything pushing back, they
can’t tell that gravity is there. It’s just like the moment when an elevator
starts to descend, and you briefly feel lighter – the floor isn’t pushing back
on you as hard as it was. If the elevator were to fall as fast as it could,
through a very deep elevator shaft, you’d feel weightless, too. In orbit, you
haven’t escaped gravity. You’ve just found a way to ignore it. But while you
can’t feel it, it’s still there, its inward pull keeping you spinning around our
planet.
Rotation is useful in all sorts of ways, but there are times when it’s just a
nuisance. For example, why is it that toast falls butter-side down? You’ve
just whisked the hot toast out of the toaster and applied a layer of butter
which is now starting to melt. All it takes is a moment of distraction while
you reach for your tea, and you knock the toast towards the edge of the
table. It teeters on the edge, and the next thing you know it’s on the ground,
face down. The lovely melting butter is now decorating your floor. It’s a
nuisance to clean up, made worse by the feeling that this must be the
universe getting back at you for something. Why does it have to happen in
the messiest possible way? Why does it turn over like that?
This is a real phenomenon. Various people have conducted experiments
where they have patiently pushed toast off tables many times, and it really
does fall butter-side down far more often than it falls butter-side up. It
depends a bit on how the fall starts, but generally, this is the way the world
works and we’re stuck with it. And it’s got nothing to do with the additional
weight of the butter. Most of the butter soaks into the middle of the toast,
and even if it didn’t, it’s only adding a tiny amount to the total mass of the
bread.
The first question is, why does it turn over at all? It all happens so
quickly that it’s hard to see (and anyway, if you’d been looking at the toast,
you probably wouldn’t have knocked it off the table to start with). You can
fn6
watch it happening if you’re happy to sacrifice a bit of toast or even a
placemat or book that’s about the same size. Put your sacrificial slice of
toast flat on the table right next to the edge and nudge it towards the
precipice. Just as the halfway point of the toast is over the edge of the table,
two things happen. One is that the toast starts to pivot around the table edge
like a see-saw. The other is that the slice starts to slide outwards without
any more nudging. The toast will now take care of itself. Slip, spin, splat.
So the rotation starts when the toast is just over halfway off the table. The
key to it all is that at this moment, for the first time, less of the toast is
supported by the table than is hanging off the edge. Gravity is pulling down
on all of the toast. The table pushes back up, but the air can’t. It’s all about
balance, just like a see-saw. The halfway point is the time when the gravity
pulling down on the overhanging side is just barely enough to lift up the
toast that’s still on the table. Physicists call the position of that halfway
point the ‘centre of mass’, which just means that a see-saw pivoted at that
point would be perfectly balanced.
The moment you realize the toast is falling is the moment it’s too late to
do anything about it. Once the toast has slipped off the table, it’s going to
take a fixed amount of time to fall. If your table is about 75 cm high, it will
take the toast just less than half a second to hit the floor. But once the
rotation has started there’s no reason for it to stop, and the toast keeps
fn7
turning as it falls. Since gravity is always the same, and tables are pretty
much always the same height, the toast always has the same spinning speed.
In 0.4 seconds, it will turn 180 degrees. Since the butter started on the top, it
ends up on the bottom. The physics is pretty much the same every time, so
the outcome is pretty much the same every time. Toast falls butter-side
down.
Interestingly, there is really only one thing you can do that might change
fn8
the outcome, but it does bear a significant risk of unintended
consequences. Just as you realize you’ve knocked the toast, just as it starts
to teeter on the edge, the physics suggests that giving it another good
sideways swipe will actually help. The toast will end up on the other side of
the room, but because it spends less time pivoting on the edge, it won’t be
spinning nearly as fast as it falls and may not rotate enough to turn over
before it lands. So it stands a decent chance of landing buttered side up –
but there is also a decent chance that it will end up under the sofa or stuck
to the dog.
Toast starts to spin because it has two things going for it: a point to pivot
around, and a force that’s pulling the toast around the pivot. It doesn’t
matter that the force only points straight down and doesn’t keep pulling the
toast around the circle. What matters is that the force is sufficient to move
the toast (which it is, if the centre of mass is over air and not table), and that
it pulls the toast around the pivot for a little while at least. Once the spin has
started, it’ll keep going until something stops it.
This is the principle behind the spinning eggs in the introduction. If you
think about the many things that spin freely – frisbees, tossed coins, rugby
balls, spinning tops – you’ll notice that they just keep going. It would be
really odd if you flicked a coin so that it span up into the air, and then
fn9
somehow stopped spinning before you caught it. Anything that spins has
angular momentum, which is a measure of its amount of spin. Unless
something (such as friction or air resistance) slows it down, the object will
spin indefinitely. This is the law of conservation of angular momentum.
Something that is spinning will keep spinning, unless something happens to
stop it.
I’m pretty sure that when I was a kid, dizziness was considered a sort of
internal toy. If you were bored, you could always spin around on the spot,
partly to see who could keep going the longest, and partly because it was
funny that everyone fell over as soon as they stopped. The spinning itself
didn’t seem to cause too many problems – the brief and entertaining
disorientation comes when you stop. It’s a shame that adults don’t play this
game very often; we might understand ourselves better if we did. The
disoriented feeling happens because of something going on inside your ears
that you can’t see, but your brain is certainly aware of.
Let’s go back to the spinning raw and boiled eggs that I was talking about
in the introduction. Each egg, still in its shell, is put down on its side and
spun around. After a few seconds of both eggs spinning, you quickly put
your finger on the shell of each to stop the rotation. Both eggs stop
spinning. You take your fingers away. Then one egg starts revolving again.
The egg that is solid has to stop spinning completely when you stop the
shell. Both egg and shell have to move together. But when you stop the raw
egg, you only stop the shell. The fluid inside is still spinning around; it’s not
connected to the shell, and so there’s no reason for it to stop spinning. So
the fluid pushes on the shell until the shell starts to rotate again.
When you spin yourself around, most of you (fortunately) is like the
hard-boiled egg. It all has to move together. So when you stop spinning,
your brain and nose and ears all stop, too. But not your inner ears. There are
small, semicircular canals in each ear that are filled with fluid precisely
because that makes them behave like the raw egg. The fluid doesn’t have to
match the movement of its container because it’s not attached to it. This is
one of the ways that your body senses where you are; tiny hairs detect how
the fluid is moving around, and your brain matches that up with what you
see. If you rotate your head, the fluid in the curved canal doesn’t rotate as
quickly, so it flows around the canals because it hasn’t caught up yet. But if
you spin for a while, this fluid starts to spin, too. It only takes a few seconds
to catch up, and then the fluid in your ears is spinning steadily with the
canals, matching the movement of its container. When you stop suddenly,
the fluid doesn’t stop. Just like the raw egg, the container has stopped, but
the fluid keeps going. So your inner ear is telling your brain that you’re
moving, but your eyes are telling your brain that you’re not. That is when
you feel dizzy, as your brain tries to work out what’s really going on.
Eventually, the fluid in your inner ear does stop spinning, because its
container has stopped. The dizziness fades away.
This is one of the reasons why pirouetting ballerinas keep facing in one
direction as they spin, and then very quickly move their head all the way
around to get back to the same direction as their body catches up. With this
very quick stop–start motion, the fluid inside doesn’t get up to a steady
spin, so the ballerina doesn’t feel disoriented when she stops.
There are two aspects to the conservation of angular momentum. The
first is that something that isn’t spinning needs a push to get it going. It
can’t just start spinning by itself. And the second is that something that is
already spinning will keep spinning unless something pushes on it to make
it stop. In our everyday lives, it’s often friction that provides the push to
slow things down. So the spinning top eventually comes to a halt and the
spinning coin slows down so much that it falls over. But in situations where
there is no friction, things really will keep spinning indefinitely. That’s why
the Earth has seasons.
In northern England, the seasons are the rhythm that gives all my
memories a cosy home. Long walks along the Bridgewater Canal on hot
summer days, hockey matches in the autumn drizzle, driving back from
Polish Christmas Eve dinner in a frosty chill, the excitement of spring days
getting longer – the variety was part of the joy of it all. One of the hardest
things about living in California was the absence of that rhythm; it felt as
though time wasn’t moving, and that was intensely disconcerting. I continue
to feel the seasons very strongly today. I like to be able to identify my place
in the yearly cycle by the cues that still mark it, even in a modern society:
animals, the air, the plants and the sky. And the foundation of all those
riches is the bit of physics that keeps things spinning unless something
stops them.
Spin has a direction, the axis around which everything is rotating. We
imagine the axis of the Earth as a line that goes from the South Pole to the
North Pole, sticking out slightly, pointing away into space. But because it’s
been clonked by solar system debris in the past (especially the huge
collision that made the Moon), the spinning top that is the Earth doesn’t
point straight up relative to the rest of the solar system. Imagine looking
down on the solar system, with the Sun in the middle and the planets
circling around on a flat plane. The axis of the Earth points slightly off to
the left. And now that it’s spinning about that tilted axis, it has to stay
spinning about that same axis. So when the Earth is on the left of the Sun as
we’re looking at it, the north end of the axis is pointing away from the Sun,
out into space. But six months later, when the Earth is on the right of the
sun, the north end of the axis is still pointing to the left – which is now
towards the Sun. The spin axis of the Earth does not change direction as it
goes around the Sun – there’s nothing pushing on it, so it must keep going
as it did before. But that means that the North Pole gets more or less
sunshine, depending on where the Earth is on its orbit. This is where our
fn10
seasonal cycle comes from. We have a day/night cycle because the Earth
fn11
keeps spinning, and a seasonal cycle because the axis of that spin is tilted.
Spinning is part of our lives in lots of ways. But there’s one device in
particular that relies on spin, and it’s one we might all be seeing more of in
the future: a flywheel. Anything that’s spinning has extra energy because of
its spin. So if a rotating object will keep spinning indefinitely, that also
means it can act as a store of energy. If you can get the energy back as you
slow the spin down, you’ve effectively got a mechanical battery. This is
what a flywheel is, and it isn’t new; they’ve been around for centuries. But
a new wave of flywheels is about to arrive in our society, a set of very
efficient modern devices that could help solve a really thorny problem.
One of the biggest challenges for any energy grid is matching up supply
and demand on very short timescales. If everyone cooks dinner at about the
same time, energy use across the country will rise for an hour or so and then
fall. Ideally, someone monitoring the system would allow energy into the
grid as it’s needed, to match that spike. But that’s a problem if the energy is
coming from a coal-fired power station that takes hours to start and stop.
And you may not even be in control of the rate or timing of energy
generation. One of the difficulties with many renewable energy sources is
that you can’t dictate when they’re generating energy – it’s easy to make
hay (or store energy) when the sun shines, but what if that’s not when you
need it?
Surely, you might say, all you need is a battery to store the extra energy
until you can use it? But electric batteries aren’t really up to the job.
They’re expensive to manufacture, they’re often based on relatively rare
metals, they have a limited number of charge/discharge cycles, and there are
limits to how quickly they can store and release energy. In response, over
the past few years, along have come some prototype flywheel projects. And
it looks as though this technology may offer a workable solution, at least
some of the time. A flywheel is a heavy spinning disc or cylinder, with
bearings that are as frictionless as possible. Once it’s spinning, it will keep
spinning. And since there is energy associated with rotation, that spin can
store energy. You use any excess energy in the grid to get your flywheel
spinning, and it will just keep spinning, holding on to the energy. Then,
when you want energy back, you slow the wheel down by converting the
energy into electricity. There’s no limit to the number of times you can
charge and discharge the flywheels, and they can release their energy very
quickly. You only lose about 10 per cent of the energy you had to start with,
and very little maintenance is needed. Even better, you can make them to
suit your needs: a small one to go with the solar panels on your roof, or a
vast bank of them to moderate spikes in the whole energy grid. Small
portable flywheels are even being tried out on hybrid buses, storing energy
as the bus brakes and supplying it back to the wheels when the bus needs to
speed up again. Flywheels are appealing because they’re based on a
beautifully simple idea – the conservation of angular momentum. Eggs and
spinning tops and stirred tea all follow the same principle. But it takes
efficient modern technology to turn it into a practical solution. It’s still early
days for the new incarnation of this technology, but you may well be seeing
spinning flywheels around much more often in the future.
8
When opposites attract
Electromagnetism
Static electricity is a start, but the real power comes when you start to move
electrons and electrical charges more systematically. Our electrical grid, the
network we use to move energy around, is an astonishing resource. By
pushing electrical charge down wires, and controlling it using tiny switches
and amplifiers, we can deposit energy wherever we need it. An electrical
circuit is just a way of redistributing electrical energy. The most important
thing about a circuit is that it is just that – a circuit. It has to be a loop, so
that electrons can keep shuffling around it without building up at the far
end. Every circuit must begin and end at a power supply, something that
will keep the electrons on the move, taking them in at one end, pushing
them along and putting them back into the circuit at the other end. The
power supply is a bit like a lift that carries people up to the top of a very
long slide. The people can go down the slide and up to the top again, round
and round all day, as long as there’s a lift to give them enough energy to get
back to the start. The rule of every circuit is that you have to lose all the
extra energy from the power supply before the electrons get back to where
they started from.
An electron shuffling along a wire is all well and good, but what’s
pushing it around the circuit? We’ve said that the first thing is to have an
electrical conductor, something that provides a path down which an electron
can move. But the other thing you need is a force to push it with.
A fridge magnet and a balloon charged with static electricity are both
weird for the same reason: they show that it’s possible to have an invisible
force field. That is, one stationary object is pushing or pulling on another
one nearby, but you can’t see what’s doing the pushing. This similarity isn’t
accidental, but the real link is only obvious once you start moving the
electrical or magnetic fields around. First, let’s go back to that principle of a
force field. It’s not just humans that can make use of them.
The stream bed is a murky brown maze of rocks, plants and tree roots.
It’s dusk and the muddy water is flowing lazily through and over the
obstacle course. A metre below the surface, two small antennae are poking
out from beneath a pebble, twitching as they taste the water. Something
moves nearby and the antennae vanish. This freshwater shrimp is a
scavenger, hungry but vulnerable. Upstream, a hunter slides into the dark
water. It paddles along the surface towards the centre of the stream with two
webbed front feet, then shuts its eyes, closes its nose, seals its ears and
dives. The platypus is ready for dinner.
If the shrimp stays perfectly still, it will be safe. The platypus swims
quickly, picking its way confidently through the maze even though it’s
currently blind, deaf and unable to smell anything. Its flat bill sweeps from
side to side, scanning the mud. Another foraging shrimp feels the water
move as the platypus approaches and snaps its tail, jerking backwards into
the gravel. The hunter swerves towards it. The signal forcing the tail muscle
of the shrimp to contract was an electric one. That electric pulse created a
temporary electric field centred on the shrimp. This electric disturbance
flashed through the surrounding water, exerting tiny pushes and pulls on
nearby electrons. It lasted for a fraction of a second, but it was enough. A
platypus has an array of forty thousand electrosensors on the upper and
lower surfaces of its bill. The simultaneous water movement and electric
pulse were all it needed to get a direction and range. The bill hammers into
the sand in exactly the right place, and the shrimp is no more.
The shrimp’s movement condemned it because the act of moving
changed its electric field. Every electric charge pulls or pushes on other
electric charges around it. An electric field is just a way of describing how
strong that push or pull is in different places, while talking about electric
signals means that an electric charge moved somewhere, and something
nearby noticed the change because the push on it increased or decreased.
Since all muscle movements involve moving electric charges around inside
the muscles, they all generate electric fields. So electrosensing is a reliable
hunting technique underwater if you’re close enough to your prey, because
no amount of colourful camouflage can disguise an electric signal. Any
animal has to move eventually, and the tiniest motion will generate an
electric signal that can give it away.
If that’s the case, why aren’t we more aware of the electric fields that we
generate ourselves? It’s partly because those fields aren’t very strong, but
mostly because electric fields decay quickly in air, which doesn’t conduct
electricity. Stream water (and especially salty ocean water) is a very good
conductor of electricity, so electric signals can be detected from much
further away. Almost all the species that use electrosensing are aquatic
(bees, echidna and cockroaches are the known exceptions).
In an electric circuit, the electrons move because there’s an electric field
inside the wire. That electric field is pushing on each electron, shoving it
along. But where does the electric field come from? A good place to start is
with a battery. Batteries come in all shapes and sizes, but there is one set in
particular that I will never forget. They were chunky sea batteries, and I
worried about them because they were floating free in a giant storm,
powering my one shot at an important experiment.
To study the physics of the ocean surface in storms, we need to go and
look at that surface. The ocean is such a complicated environment that
theorizing from a nice warm office is of limited use unless you’re sure that
what you’re working on is definitely based on reality. But even when you
get ‘there’, on a ship miles from shore in rough seas, it’s still difficult to
touch the region I’m interested in – the water just a few metres below the
sea surface. Knowing what happens there will improve our understanding
of how the oceans breathe, and will contribute to better weather forecasts
and climate models. But to see the details, you need to be in it; and it’s a
violent, messy, dangerous place to be. I can’t swim in that water, but my
experiments have to. The experiments need power, an electricity supply,
and they need it while they’re bobbing up and down in the waves, free of
the ship. You can’t plug them in, so you have to rely on batteries. And
fortunately for me, electrical circuits work just as well when they’re
bobbing up and down as they do when on dry land.
The bosun scowled at the horizon, stuffed his hands deep into the pockets of
his paint-splattered hoodie and swayed along the deck of the ship towards
me. It was November in the North Atlantic, and I hadn’t seen land in four
weeks. Everything was always either going up or going down as we clung
to a heaving grey sea that merged with grey sky in every direction. The roll
of electrical tape I’d just put down on the deck took advantage of my
temporary distraction and skidded across the deck until it met the bosun’s
boot. His thick, cheery Boston accent seemed comically out of place in this
forbidding environment. ‘How long you gonna be?’
For me, the worst bit about doing experiments at sea has always been
these final checks before letting the experiments float free. I was nervous,
and this bit was my responsibility alone. To measure the bubbles just
beneath the breaking waves, I was using a large yellow buoy with a variety
of measurement devices strapped to it. The bosun was in charge of
manoeuvring this beast off the ship and into the rolling sea, but I had to
make sure that it was ready. The storm that was coming would be a big one,
and I desperately wanted good data from it. ‘I’m just about to plug in the
batteries, and then I’m ready to go,’ I said. The monstrous yellow buoy, 11
metres long, that carried my experiments was strapped to the deck, shackled
securely until it was safe to release it. I started with the armoured camera
near the top and put my hand on the power connector, following the wire all
the way down to the bottom of the buoy where the chunky batteries were
and then plugging it in. Then back up to the acoustical resonators. Hand on
the power cable, follow it down to the batteries, plug it in. Check that the
connection is secure. Check again. Back again to the other camera. These
experiments could carry out incredibly delicate and sophisticated
manipulation of the physical world, but only when provided with electrical
energy. And the providers were four cumbersome lead–acid sea batteries
that weighed 40 kg each, and whose basic design hadn’t really changed
since they were invented in 1859. But they worked.
When it was time, we scientists huddled in our oilskins at the far end of
the deck and the crew and crane took over, manoeuvring the swaying
monster over the side and into the dark ocean. As the last rope slipped free,
there was a weird shift of perspective, and the huge yellow beast became a
vulnerable bobbing piece of flotsam, tiny compared to the vast ocean and
frequently hidden by the waves. A burst of chatter spread along the ship’s
rail about how the buoy was sitting in the water, and the speed at which it
was drifting away from the ship. But I wasn’t thinking about any of that. I
was thinking about electrons.
Below the waterline, the dance of the electrons had started. They were
shuffling out of the battery, around the circuits carried by the buoy and then
back into the other side of the battery. There were a fixed number of
electrons held in the circuit, all just going around the same loop. The
electrons don’t get used up – they just go round and round. The trick is that
it takes energy to push them around, and they give that energy away as they
travel. The source of that energy is the battery, and a battery is a very
ingenious device.
The clever thing about batteries is that they join up a chain of events.
Each link in the chain supplies the electrons that the next link needs; and so
once a battery is connected to a circuit, everything is in place for electrons
to flow around the loop. These sea batteries had two terminals sticking out
to connect them to the outside world. Inside, each terminal was connected
to one of two sheets of lead, but those two sheets weren’t touching. The
space in between the lead sheets was full of acid, which is why they’re
called lead–acid batteries. There are two ways in which the lead can react
with the acid. There’s one that needs some extra electrons from somewhere,
and there’s another one that gives away extra electrons. A lead–acid battery
is charged when those two reactions have been pushed as far as they can go.
When I plugged the equipment into each battery, I was providing a path
all the way from one lead sheet through my experiments to the other lead
sheet. And then there was the crucial last piece to the jigsaw: because of the
chemistry at the lead plates, there was an electric field down the wire.
Electrons were being pushed along the wire, away from one lead sheet
towards the other. They couldn’t get there across the acid, so the only
option was the outside circuit, the long way round. Once the electrons have
a path with an electric field pushing them down it, the reactions can undo
themselves because the chain is complete. One set of lead plates gives
electrons to the acid, and then the acid passes this charge on to the lead at
the other plate. The lead there takes electrons as it reacts, and the whole
thing keeps going because the electrons can then shuffle around the circuit
back to the first set of plates. The really important fact is that on that trip
through the camera around the back, the electrons have some extra energy
to get rid of. This is electricity. And if you arrange it so that on their way
they have to pass through a sophisticated electrical circuit, hey presto: you
can put that energy to work, and you’ve got a useful battery.
As I leaned over the rail of the ship watching that bobbing yellow buoy, I
was imagining this dance. The camera would switch on, creating a pathway
for electrons from the battery, and they would bounce their way up the stem
of the buoy, into the camera housing. You have control over where the
electrons go because you know that they’ll take the easiest path. So you
arrange a path through the maze that is made of conducting material. The
power cable is metal, easier for an electron to move through than the plastic
coating around the wire, so you know that electricity will flow down the
wire rather than escaping into the surrounding material. Beyond that, the
most basic element of control is a switch. A closed switch is just a place in
the circuit where two parts of electrical wire touch. They’re not glued
together, but when they’re touching, electrons can move between them. To
stop the flow, you just move one of the wire ends away from the other.
Electrical flow stops because there is no longer an easy route through.
Once inside the camera, the path of the electrons would split, some
shuffling into the computer and some into the camera itself. And the thing
about electrical circuits is that in the end all roads lead to Rome, or in this
case, back to the battery. The massive yellow buoy was just the skeleton for
this branching flow of electrons, and the electrons themselves were
generating electric and magnetic fields, pushing and pulling on camera
shutters, acting as timers, generating bursts of light and recording data in a
huge, intricate synchronized sequence, before shuffling back to the battery.
And all of that was happening while the buoy was being shoved around
by the huge waves (8–10 metres high in some cases) of an Atlantic storm.
We bobbed about on the research ship and waited, living a life where
gravity was an uncertain friend and a tenuous grasp on order was
maintained only by imprisoning things with Velcro or elastic bungees or
rope. After three or four days, the chemical reaction in the battery had
finished – it was back in its original uncharged state. There was no more
stored energy left, the electrons could not be pushed around the circuits, and
the dance was over. The buoy went back to being an inanimate shell of
metal and plastic and semiconductors. But the data had been stored in solid
state computer memory, and it was safe.
A few days later, when the storm was over, we tracked the buoy down
and hauled it back on board. I’m always extremely impressed by the skill of
research ship crews at fishing things out of the water. Ships don’t move
sideways, and they’re slow to turn around or change direction. To stand a
chance of getting the buoy back, the captain had to bring his 75-metre
vessel right alongside it, managing both to avoid running it over and to get
close enough for the bosun to reach over and catch it with a long boat-hook.
And they usually succeeded on the first try.
Then it was our turn again. The batteries were plugged into the ship’s
power supply, providing the energy to push the chemical reactions back the
other way ready for the next deployment. The experiments were detached
and brought inside, with the exception of the camera. This got left outside
in the freezing cold, because the dance of the electrons has a downside, and
my poor PhD student was about to pay the price for it.
Possibly the most fundamental physical law we know of, one that has
been shown to be accurate time and time again and has never been
disproved, is that of conservation of energy. It states that energy can never
be created or destroyed, but only shifted around from one form to another.
The battery had chemical energy, and the chemical reactions converted that
to electrical energy, and then somewhere between one terminal of the
battery and the other one, that energy moved on. But where did it go?
Things happened – the camera took pictures, the computer programs ran
and data were recorded. But none of that stored the electrical energy in a
new place. The energy just leached away, unnoticed. There is a price to be
paid for moving electrons around, and it’s the generation of heat. Any
electrical resistance inflicts an energy tax on the electrical energy moving
through it. Even though the electrons will pick the path of least resistance,
fn5
some tax must still be paid.
The camera was housed in thick plastic, a material that transmits heat
very badly. When the camera was running, all the energy of the electrons
was eventually converted to heat as they flowed around the system. That
didn’t matter in the water, because the ocean where we were was about 8°C
and stole the heat away, cooling the housing efficiently. But air isn’t up to
that task. In the lab, when the computer was running to download the data,
the camera kept overheating. We did our best, but the only solution we
found was to leave it outside in a bucket of iced water (helpfully, the ship
had an ice machine), and so my PhD student had to spend nine or ten hours
starting and stopping the downloads to keep the data flowing while
preventing the camera cooking itself. Such is the glamour of field science.
This is why laptops, vacuum cleaners and hairdryers heat up as you use
them. The electrical energy must go somewhere, and if it’s not converted
into other kinds of energy, heat is the inevitable end. Hairdryers use this to
heat air; their circuits are arranged to dump energy as heat in a very
concentrated way. But laptop manufacturers hate heat, because hotter
circuits work less efficiently. There is no way of using electrical energy
fn6
without paying a heat tax.
So the electrons flow because an electric field is pushing on them. A
battery doesn’t really provide electrons – there are plenty of those in the
world. What it does is provide the electric field to move electrons. And if
the circuit is complete, this electric field will push electrons around the
loop. So far, so simple. But what are all those numbers on plugs and in tiny
font on the safety warnings? Perhaps it’s best to take the typical British
approach to all problems: find the biscuit tin and put the kettle on.
The most important thing about a teabreak is that it involves both tea and
a break. Some of my American co-workers never really understood this,
and used to bring along work to continue discussing it over tea. But for the
British, the act of ‘putting the kettle on’ signifies a change of pace. I’m
going to do it now, and in this case my kettle is an electric one that I simply
fill with water and plug into the mains. My mind is allowed to stop working
for a bit, while the kettle gets on with its job.
Pushing down on the switch does one very simple thing. It shifts a bit of
metal and thereby slots the last segment of a circuit into place. Now there’s
a route through the maze of the kettle, a path made entirely of electrical
conductors that electrons can easily travel along. This path is now
uninterrupted and it runs from one pin of the plug, through the kettle, and
back to the second pin of the plug. In this case, the electric field comes not
from a battery, but from a plug socket.
A standard three-pin plug has one long pin at the top. That’s called the
ground pin. It’s completely separate from the rest of the circuit. Effectively,
it’s doing the job that my car did on those cold snowy mornings – it’s there
to provide an escape route if any electrons start to build up in the wrong
place (say, on the outside of the kettle). So that’s not part of the path that’s
going to power the kettle.
The other two pins, the smaller ones, are going to do the electron-
pushing. One of them behaves like a fixed positive charge, and one like a
fixed negative charge. As I press down on the switch, I connect up a path
that now has an electric field running along it. Electrons along that path are
feeling a push away from the negative side and a pull towards the positive
side. So as I find the teapot and dig out teabags, the electrons start to
shuffle. They’re jiggling around quite a bit anyway, but now they have a
slight tendency to drift down the wire. And what that means is that overall,
there’s a movement of electric charge from one pin of the plug, through the
kettle, and out through the other pin of the plug.
On the bottom of my kettle, a label tells me that it’s designed to work at
230 volts (230V). The voltage is related to the strength of the electric field
that’s pushing electrons along the circuit. The stronger the electric field, the
more energy each electron has to get rid of along the way. That’s what a
high voltage is telling you – it’s saying that this is the amount of energy
available for use along the path of the circuit. In terms of the slide analogy
from earlier on, the voltage is the height of the slide that the electrons have
to shoot down before getting back to the other pin of the plug. The higher
the voltage, the more energy each electron has to dump on the way.
I’ve swilled out the teapot and put the teabags in it, the milk and a mug
are out and ready. Now I’m just waiting for the water to heat up. It only
takes a couple of minutes, but when I’m thirsty, I’m very impatient. Hurry
up! I know what the voltage of the electrical supply is, but that’s only part
of the story. The higher the voltage, the more energy each electron can give
up. But that doesn’t say anything about how many electrons are passing
through. The fastest way to dump lots of energy in the water is to make sure
that lots of electrons are flowing around the circuit. That’s what an
electrical current is, and we measure it in amps. The higher the current, the
more electrons are moving past one point in the wire in any one second.
When you multiply the voltage of the supply by the current (in amps)
flowing through the circuit, you get the total amount of energy deposited
per second. My kettle runs from a 230V supply, and can draw a current of
13 amps, and 230 × 13 = 3,000 (approximately). The base of the kettle
agrees – it says that the kettle power is 3,000 watts (3,000W), which
equates to 3,000 joules of energy released per second. That’s enough to heat
my water to boiling in just less than two minutes, but it will lose a bit of
heat to the surroundings, so in practice it takes closer to two and a half
minutes.
I’ve no intention of testing this out while I’m waiting for my tea, but they
say ‘volts jolt, current kills’. The voltage difference between me and my car
on that snowy day in Rhode Island was probably 20,000 volts. But only a
tiny amount of electrical charge went anywhere, so it didn’t do me too
much harm. The current was tiny and very little energy was transferred. If I
connected up the path between the two plug terminals with my fingers, so
that my body took the place of the kettle, it would be a different story. A
high current means that there are lots of electrons, each carrying the same
amount of energy. The total amount of energy is huge, because so many
electrons are rushing through. It would be far more dangerous than the
shock from the car, even though the voltage difference across the pins of the
kettle is only about a hundredth of the voltage difference between me and
my car. It’s the current that matters most when it comes to potential harm to
you.
As the electrons shuffle through the metal of the heating element, they’re
being pushed by the electric field. That makes them speed up slightly, but
the conductor is made up of lots of atoms, and so these sped-up electrons
inevitably bump into things. When they bump, they lose energy, heating up
whatever they bump into. And so forcing lots of charge to move means that
there’s lots of bumping and lots of heating. That’s all the kettle is doing –
speeding up electrons so that they bump into things and pass on their energy
as heat. The electrons themselves don’t travel very far at all – they might
drift at about 1 mm per second. But it’s enough.
Boiling water has loads of extra energy, and it’s amazing that it gets there
just from minuscule electrons shuffling about and bumping into things.
Amazing, yet undeniable; my tea is ready, heated by electric fields pushing
on electrons in a conductor. This is the simplest possible use for electrical
energy: converting it directly into heat. But once people had worked out
how to build circuits and power supplies and batteries, things got much
more sophisticated very quickly.
There is a fundamental difference between the electron dance generated
by batteries (any batteries) and what happens when you plug a device into
the mains. In any device powered by a battery, the electrons are always
flowing in one direction only. This is called direct current, or DC. A
standard AA battery will supply about 1.5 volts DC. But the mains current
is different – it’s alternating current, or AC. That means it switches
fn7
direction about a hundred times a second. It turns out to be more efficient
if you run your electricity supply like this.
You can switch between DC and AC, but it’s a bit of a nuisance. Anyone
who carries around a laptop power cable will be familiar with this kind of
nuisance – it’s the small heavy block that sits in the middle of the cable. It’s
called an AC/DC adaptor, and its job is to convert the AC current from the
mains into the kind of DC current that your laptop wants (which is what the
laptop battery provides directly). To do that, it needs coils of wire and a bit
fn8
of circuitry, and it’s still tricky to make all the necessary bits any smaller.
So for the time being, we’re stuck with carrying around the adaptors.
We take electronics for granted today. But in its early days, it was a
capricious and uncertain beast. My own grandfather was getting involved in
that world just as all its new sophistication was making its way into our
homes.
My grandfather, Jack, was one of the first television engineers. Back in
his day, electronics could be loud and hot and was certainly capable of
generating quite a pong – as my grandmother readily recalls. Her
description of the sort of problem that he used to have to fix reminded me
of that physicality about early electronics that’s easy to forget in these days
of smartphones and wifi on tap – and also surprised me with her familiarity
with all the components and processes. I’d never really heard her talk about
anything technical in my life, and yet when it came to these old TVs, she
was comfortable with specialized electrical terms I’d never come across.
‘Well,’ she told me one day, ‘one important component was a line-output
transformer, and when that went on the TV, it sometimes went with a bit of
a bang but it also produced a burning sensation and a smell.’ Her northern
accent reminds me that this is almost certainly flat-capped understatement.
Electrons have always been invisible, but from the 1940s to the 1970s you
could definitely tell they were up to something. There was always the risk
of a bang or a pop or a hiss, the sudden appearance of a sooty burnt patch or
a flash of light that told you lots of energy had just been shunted somewhere
it shouldn’t have gone. Jack found himself in at the beginning of the new
world of television, part of the only generation that had a real feel for the
electrical world. By the end of his career, transistors and computer chips
had hidden it all away. The tiny exterior of these components conceals a
vast and sophisticated interior, incomprehensible from the outside. But
before they came along, there were a few decades when you could almost
see the magic at work.
In 1935, at the age of sixteen, Jack had started a trade apprenticeship with
Metropolitan Vickers, locally known as MetroVick. This giant of the heavy
electronics engineering world was based in Trafford Park near Manchester,
turning out world-class generators, steam turbines and other large-scale
electronics. When he finished his apprenticeship in electrical engineering at
the age of twenty-one, he was classed as having a reserved occupation,
deemed too useful to go to war, so he spent five years testing aeroplane gun
electronics at MetroVick. The first test of these systems was called
‘flashing’. You put 2,000 volts across it and if it didn’t go bang, it passed.
This was the raw end of the taming of the electron, the early stages of
wrangling it into submission.
After the war, EMI was looking for people with electronics experience,
because early televisions were skittish, complex beasts, needing an expert
to set them up and frequent adjustment throughout their lifetimes. So EMI
sent Jack to London to train as a television engineer. The tools of this trade
were valves and resistors and wires and magnets, the components that could
coax electrons to do your bidding. This visually beautiful potpourri of glass
and ceramic and metal could be made to do something that sounds very
simple, something that was at the heart of every television set until the
1990s. It could make a beam of electrons and then bend it; and if you do
that right, you can make moving pictures.
Jack learned about ‘CRT’ televisions, and I love that name because it
connects us to the world that existed before electrons had even been
discovered. CRT stands for cathode ray tube, and cathode rays were
distinctly odd when they were first discovered. Imagine the early German
physicist Johan Hittorf, in 1867, looking at his latest creation. In the gloomy
lab, there’s a glass tube with two bits of metal sticking into the space inside
at either end, and all the air inside the tube has been removed. This sounds
fairly mundane. But imagine how odd it must have been to discover that if
you connect a large battery to the two bits of metal, mysterious invisible
stuff flowed from one end of the tube to the other. He could tell it was there,
because it made the far end of the tube glow, and he could make shadows
by putting things in the way. Even though no one knew what was flowing, it
needed a name, so it became known as cathode rays. The cathode is the
terminal attached to the negative end of a battery, and that’s where the
strange stuff was coming from.
It was another thirty years before J. J. Thomson discovered that what was
flowing wasn’t really rays at all, but a stream of individual negatively
charged particles – the particles we now call electrons. By then it was too
late to change the name of the apparatus, though, and it was still always
called a cathode ray tube. We know today that applying a voltage across it
generates an electric field stretching from one end to the other, and so
electrons will hop off the negative end and rush towards the positive end.
Any particle with an electric charge will be accelerated by the electric field,
which means it will be constantly pushed along. So the electrons don’t just
move towards the positive end because they’re attracted by it, they’re
speeding up as they go. The higher the voltage difference between the two
ends, the faster they’re going when they get to the other side. In a CRT TV,
they can be going at several kilometres per second by the time they hit the
screen. That’s a significant fraction of the speed of light, the fastest that
anything in the universe can travel.
So the same basic process that led to the discovery of the electron in the
first place was in use inside every TV in the world until a couple of decades
ago. Each CRT TV has a device at the back that produces electrons. The
middle of the TV is a completely empty chamber – a vacuum with no air in
it – so there are no obstacles at all, and so the electrons ‘fired’ from the
‘electron gun’ stream across that empty space until they hit the screen. It’s
the purest form of electrical current – charged particles moving in a straight
line.
My aunt opens a box full of bits and pieces she saved from Jack’s workshop
when he died. There are glass tubes that look like cylindrical lightbulbs,
with a weird metallic insect-like structure inside each one. These are valves,
used to control the flow of electrons in circuits. Early on, most of Jack’s job
seemed to involve working out which one of these had malfunctioned, and
replacing it. My mother, my aunt and my grandmother clearly have a lot of
affection for these, because there were so many of them around back then,
and so many different types. And then in the corner of the box there’s a
large circular magnet, now broken in half.
This is the great connection, and it represents the moment the penny
dropped for the physicists of the late 1800s. If you want to control
electricity, you need magnets. If you want to control magnets, you need
electricity. Electricity and magnetism are part of the same phenomenon.
Both an electric field and a magnetic field can push on a moving electron.
But the result of the push is different. An electric field will push an electron
in the direction of the field. A magnetic field will push a moving electron
sideways.
Creating a beam of electrons is all well and good, but the real cleverness
of these old television sets was to control what that beam was pointed at.
And the deep connection between electricity and magnetism is at the heart
of it. As an electron zooms through a magnetic field, it gets pushed to one
side. The stronger the magnetic field, the more it gets pushed. So by
changing the magnetic fields inside an old TV, the electron beam could be
pushed and pulled to a point wherever it was needed. The large, permanent
magnet that my aunt showed me was used very close to the electron gun, to
do the basic focusing. But the steering electromagnets positioned a bit
closer to the screen were being controlled directly by the signal from the
aerial. They pushed the electron beam so that it scanned horizontally across
the screen, one line at a time. The beam itself was being switched on and
off during each line, and where it hit the screen, a bright spot was created.
The ‘line-output transformer’ that Nana mentioned was the bit of kit that
controlled the scanning. To make a smooth picture, 405 lines were scanned,
50 times each second, with the electron beam flicking on and off at exactly
the right time for each pixel.
This is an incredibly intricate electronic dance. To see any picture at all
as a result of it requires a lot of fiddly components all doing the right thing
at the right time. So early televisions had lots of knobs and dials to make
adjustments, and it sounds as though the temptation to mess with them was
too much for many TV owners. Jack had the knack of knowing how to
readjust them. It must have seemed like magic at the time. For centuries,
craftsmen had been respected for what they could produce, and everyone
could appreciate what they’d done even if they couldn’t do it for
themselves. Now the world had changed. Electronics engineers could make
a device function, but it was impossible to see what exactly they’d done or
why it had worked.
It’s odd that silent, invisible electrons locked away in a vacuum were the
key to the huge richness of visual broadcasting with all its sound and
spectacle. And for fifty years televisions were based on the same simple
principle. Put an electron in an electric field, and you’ll speed it up or slow
it down. Put that moving electron in a magnetic field, and its path will
curve. Leave it there long enough, and it will go round in circles.
The massive physics experiment at CERN in Geneva, famous for the
fn9
discovery of the Higgs boson in 2012, works on exactly the same
principles as a cathode ray tube, although the particles it can shift around
aren’t just electrons. Any charged particle can be accelerated by an electric
field and have its path curved by a magnetic field. The Large Hadron
Collider, the experiment that finally confirmed the existence of the Higgs
boson, had protons zooming around in its guts. In this case, the speeds
reached were incredibly close to the speed of light, so fast that even with
extremely powerful magnets to steer the zooming particles, the circle had to
be 27 km in circumference.
So the basic setup used both to discover the electron itself and to run the
Large Hadron Collider at CERN, a controlled stream of charged particles in
a vacuum, also sat in a corner of many homes until very recently. These
days, the bulky CRT televisions have been replaced almost entirely by flat
screens. In 2008, sales of flat panel displays overtook CRT screens
worldwide, and the world has never looked back. The switch made laptops
and smartphones possible because it made them portable. The new displays
are also controlled by electrons, but in a much more sophisticated way. The
screen is split into many tiny boxes called pixels, and electronic control of
each pixel determines whether or not it gives out light. If you have a screen
resolution that is 1280 × 800 pixels, that means that you’re looking at a grid
made up of just over a million individual dots of colour, each separately
switched on and off with tiny voltages, and each updated at least sixty times
every second. It’s an astonishing feat of coordination, but it’s still trivial
compared to what your laptop gets up to.
Let’s return to the magnets. A magnetic field can push electrons around,
and so it can control electric currents. But that’s not the limit of the
interlinking of electricity and magnetism. Electric currents also make their
own magnetic fields.
It’s easy to take a compass for granted, especially if you do a lot of walking,
when it’s very handy to be accompanied by a needle that always points
north. But imagine getting ten compasses, or twenty, or two hundred. You
spread them out across the floor, they all point north, and suddenly you see
that this isn’t just something that happens when you get a compass out. It’s
there all the time, and it’s consistent. You can take your compass collection
anywhere on the globe, unpack it, set it out, and all the compasses will
swing round and agree on where north is. The Earth’s magnetic field is
always there, flowing through cities, deserts, forests and mountain ranges.
We live inside it, and although we never feel it, a compass will always
remind us that it’s there.
A compass is a brilliantly simple measuring device. The needle is a
magnet, and so one end of it behaves very differently from the other end.
Unhelpfully, these two ends of the magnet are called the north and south
poles, but it’s just a way of saying that one behaves like the magnetic north
pole of the Earth, and the other behaves like the magnetic south pole. If you
take two magnets and move them about near to each other, you’ll see very
quickly that it’s very hard to push the two north poles together, but that a
north pole and a south pole will attract each other very strongly. This is why
it’s easy to detect the direction of a magnetic field; if you put a small mobile
magnet inside a magnetic field, it will spin around until its north and south
ends are aligned with the field. And that’s all a compass is: a mobile magnet
that gives away the direction of any magnetic field you bring it into. We
can’t see the vast magnetic field of the Earth, but we can see the compass
needle respond to it. It’s not just the Earth’s field that compasses sense,
either. Take a compass around your home, and you’ll detect for yourself the
magnetic fields that surround plug sockets, steel pans, electronics, fridge
magnets and even any iron that’s been close to a magnet recently.
Compasses, obviously, are mostly used for navigation. Finding your way
about on the surface of a sphere is always going to be tricky, but the Earth’s
magnetic field has provided a fabulously reliable tool for explorers for
centuries. The Earth has a magnetic north pole and a magnetic south pole,
and anyone with a compass can orient themselves towards one or the other.
As a navigational tool magnetism is straightforward, it’s cheap, and it never
runs out. However, it does have a few caveats attached to it. Caveat number
one sounds unexpectedly serious: the magnetic poles aren’t fixed in one
place. They wander, and they can travel a very long way.
On the day I’m typing this, the magnetic north pole is in the far north of
Canada, about 270 miles from ‘true north’, which is the real North Pole
defined by the Earth’s spin axis. Since this time last year, the magnetic
north pole has moved 26 miles – it’s on its way across the Arctic Ocean
towards Russia. This sounds spectacularly unhelpful for navigators,
although since the world is a big place it’s not as bad as it seems. But the
magnetic field moves because of where it comes from, and it’s a reminder
that the innards of our planet are more than just a static ball of rock.
Deep down beneath our feet, the iron-rich outer core of the Earth is
churning slowly. It’s shifting heat from the centre out towards the surface,
and the rotation of the planet forces the molten rock to rotate, too. Because
of the iron, the sluggish outer core is an electrical conductor, and that means
it can behave like the electromagnet in the toaster. It’s thought that the
currents running through the Earth’s outer core as it turns are responsible
for generating our planet’s magnetic field. The process is based on the slow
shifting of molten rock; and because the details of the rock movements
change with time, the magnetic poles meander. They stay approximately
aligned with the spin axis of the Earth because the rotation of the iron-rich
rock is caused by the rotation of the whole planet, but the alignment is only
approximate.
So if you really care about accurate navigation, you need to correct for
the current position of the magnetic pole, because it’s not the same as the
true North Pole. Today’s maps show the direction of both poles. I just had a
look at an Ordnance Survey map of part of the south coast of the UK, and
both magnetic and spin north are marked at the top. I can see that if you
followed a compass directly north for 40 miles, you’d end up about 1 mile
west of the line towards true north. A map seems like such a permanent
record, and yet the magnetic field that you may use to help you navigate
with it is fickle. Modern technology means that you and I won’t often get
lost because of this. But the aviation industry, with one of the most
sophisticated modern navigation systems humans have developed, certainly
pays attention. For a start, it has to keep relabelling its runways.
Next time you’re at or near an airport, take a look at the large signs at the
start of each runway. Every runway around the world is labelled by a
number, which is its direction in degrees from north, divided by ten. So the
runway at Glasgow Prestwick airport was given the number 12, because a
plane landing on it will fly in on what’s called a ‘heading’ of 120 degrees.
Each runway will have a specific designation that’s a number between 01
fn10
and 36. But this heading is relative to magnetic north, because that’s what
a compass is telling you. So in 2013, runway 12 at Glasgow became runway
13, to keep up with the movement of the magnetic pole. The runway hadn’t
moved, but the Earth’s magnetic field had. Aviation authorities keep an eye
on it all, and correct the runway designations when it becomes necessary.
Since the poles move relatively slowly, the changes are manageable.
But the meandering of the poles is only the start. The Earth’s fickle
magnetic field has far more to offer than navigational assistance. And the
clues it leaves behind have provided the final confirmation of one of the
most controversial, simple and profound ideas that geologists have ever
had. The continents, the immense rocky masses that dominate Earth’s
surface, are moving.
Human
I’m breathing, and so are you. Our bodies need to take oxygen molecules
from the air and send carbon dioxide back out. Each one of us walks around
in our own personal life-support system, a body with an inside and an
outside. Our insides can do all sorts of things, but only when supplied with
stuff from the outside: energy, water and the right molecular building
blocks. Breathing is just one of the supply routes. It’s ingenious: expand
your ribcage, increase the volume of your lungs, and the mob of tiny
jostling air molecules close to your mouth is pushed down your windpipe
by the air further away. Take a deeper breath and your chest expands even
more, making space for more of the atmosphere to rush in and touch the
smallest structures in your lungs. Then, as you relax the muscles around
your ribcage, your ribs are pulled downwards by the Earth, pushing the air
molecules in your lungs closer together until they jostle each other back to
rejoin the outside. Oxygen isn’t the only molecule pulled into your lungs
that your body can use. As the air passes by the sensors inside the top of
your nose, some of the billions of molecules bumping into the walls will
happen to collide with a larger molecule attached to the wall, one they
temporarily fit into like a key into a lock. The underlying cell senses the
molecular click as they fit together. That’s the start of our sense of smell – a
few floating molecules of the right type bumping into the right place. Inside
now has some information about what’s outside.
A human body is a vast coordinated collection of cells, about 37 trillion
of them last time anyone tried to count, each one a tiny factory. Every
single cell needs supplies, but it also needs a safe environment, with the
right temperature, pH and moisture level. As you walk through the world,
your body is constantly adjusting to adapt to the conditions around it. If you
spend too long in a warm room, the molecules close to your skin surface
vibrate faster because they’ve got more energy. If those vibrations were
passed deeper into your body, they could start to disrupt the workings of
your cells. So as you’re sitting in the warm room, you need to give away
energy. That sounds easy; water molecules evaporate easily in the warm,
taking energy with them. You’ve got lots of water in you that could
evaporate. But the water is stuck inside because you’re waterproof. You
need to sweat.
Your skin has a very thin layer of fat molecules just underneath the
outermost skin cells, a barrier preventing any liquid moving between inside
and outside. But while you’re sitting in the warm room, your skin will open
up tunnels through the barrier: your pores. Sweat seeps through the pores,
penetrating the waterproof layer, and reaches the outside. Individual water
molecules bump into each other and the warm skin surface until the most
energetic are travelling fast enough to escape. One by one, they float away,
leaving your skin cooler. When you are cool enough, the pores close, and
you are waterproof again. Your skin isn’t waterproof just to keep water out.
It’s also waterproof to keep water in, because the internal supply of water is
limited. Water is transported round your body in your blood, the internal
supply system that lets your body share out its resources. This supply
system must run continuously to keep your cells alive. And we can check
that it’s working: we all have a pulse.
Our pulse is a three-dimensional disturbance, a travelling pressure wave
that provides clues about blood flow. Our hearts are constantly squeezing
the blood in their chambers, raising the fluid pressure and so forcing the
blood out into our arteries. It’s a powerful push, and as it comes to an end,
the fluid pressure in the heart chambers drops. The forces on the blood have
now been reversed, and the recently expelled blood would rush back in if it
weren’t for the one-way valves guarding the exit. The sudden rush of fluid
backwards closes the valves, and the liquid thumps against the valve tissues
as it’s stopped. This thump is so strong that it pushes outwards on the
tissues around it, and they push on the tissues beyond, and a pressure wave
travels through the body, slightly compressing the muscle and bone in its
path as it travels. It takes this pressure wave about six milliseconds to reach
the outside of the body, and if you put a stethoscope or your ear up against
someone’s body, you can hear it. This is your heartbeat. If waves didn’t
travel through our tissues, we wouldn’t ever be able to hear our hearts. In
fact it’s a double beat, with two pulses: ‘lub-dub’, because the heart has four
valves and they close in pairs, one pair just after the other. This accidental
combination of physics and physiology broadcasts the most significant sign
of life throughout our bodies.
After sweating, your blood carries fewer water molecules than before.
Now, your body needs to replenish itself from outside. In order for you to
take a simple drink of water, your cells need to coordinate their activities.
Decisions, and the actions required to coordinate the necessary body parts
to carry them out, are made first subconsciously and then consciously in the
brain.
One brain cell is no use by itself. It only works because it’s connected to
others, and the network of connections seems to be as important as the brain
cells themselves. As a decision about finding a drink emerges from the
connections, the brain cells need to connect with other cells further afield.
The vehicle for this internal communication is a nerve fibre, a thin strand of
cell that is the body’s equivalent of an electrical wire. By shunting
electrically charged particles across a membrane at one end of the nerve
fibre, the brain cell starts an electrical signal which ripples along the nerve
fibre by playing electrical dominoes. At the end of the first nerve fibre,
there’s another one. The dance of the electrically charged particles sends the
message across the gap, and then more electrical dominoes carry the
message onwards. The message is relayed from cell to cell for the fraction
of a second that it takes to reach one of the muscles in your leg. At around
the same time, the messages from other nerve fibres carrying coordinated
signals to other leg muscles also arrive, and your leg muscles contract to lift
you off the sofa. The feel of the floor beneath your feet and the temperature
change on your skin as the movement generates a slight breeze are
conveyed back to your brain via yet more electrical signals.
There is a phenomenal amount of information being shunted about inside
us, carried either by electrical nerve signals or by chemical messengers such
as hormones. All the disparate organs and structures of a human constitute a
single organism because we are connected not just by resources, but by
information, vast, coordinated, overlapping streams of it. Long before the
‘information age’, we ourselves were information machines.
That information falls into two categories. The first is the travelling
information: nerve signals and chemical signals that are moving right now,
seeping and flashing and flowing through us. But we also carry vast
quantities of stored information, the molecular library that is filed away in
our DNA. In the world around us, millions of similar atoms clump together
to form large agglomerations of glass or sugar or water. But in the giant
molecule that is a DNA strand, each minute atom sits in its prescribed
place, and the precise placing of individual atoms of different types gives
our bodies an alphabet. A piece of the cell’s molecular machinery can walk
along the strand, reading the genetic alphabet of A, T, C and G, and use that
information to build proteins or regulate the activity of the cell. We have to
be gigantic compared with atoms because each factory-like cell has to
contain so much.
Our bodies are immense machines; even a single cell might contain a
billion molecules, and there are around 10 million million (1013) cells in our
body. We need impressive signalling and transport systems to coordinate all
these constituent parts, and that coordination takes time. No human has
‘lightning reactions’, because the cost of our wonderful complexity is the
huge amount of time it takes us to get anything done. The shortest time that
we can appreciate is approximately the blink of an eye (about a third of a
second), but in that time millions of proteins have been built inside us and
billions of ions have diffused across our nerve synapses, while the simpler
world outside our bodies has just been getting on with things.
Our internal information engine carries on churning as we walk from one
room to another. But this gigantic system needs information about what’s
around it. Just at this moment, we need to find water. We have sensors built
in, body parts that change in response to the environment and share that
information with our brain. The sense that we’re most aware of is probably
our sight.
We live submerged in light, but our body keeps most of it out. That sea of
light carries information about the world, because light’s nature offers clues
to its origins, but most of this information goes straight past us. A tiny
fraction of this luminous cornucopia falls on to the pupils of our eyes, two
circles at most a few millimetres in diameter. A small subset of what arrives
at the pupils, the visible light rays, is allowed in. From this tiny sample
comes all of the visual richness that we take for granted. As they cross the
boundary, these light waves must be marshalled so that the information can
be harvested. Our windows on the world are guarded by soft, transparent
lenses which slow the light to 60 per cent of its speed in air. As these light
rays are slowed down, they swerve, and the lens shape is tweaked by tiny
muscles to make sure that all the rays from a single object out there, outside
the body, will meet again at the back of the eye. This selection process is
astonishing. We assume we see all there is, but really we sample only the
tiniest fraction of what is out there to build our picture.
The light rays hitting our retina may have travelled from the Moon or
from our fingers, but they have the same effect. A single photon is absorbed
by a single opsin molecule, twisting the molecule around to start a chain of
dominoes that sends an electronic signal into our control systems. As our
thirsty body walks into the kitchen, photons that have bounced off a sink, a
tap and a kettle stream into our eyes, and our brain processes that
information in the blink of an eye to tell us what to pick up first. If it’s
slightly dark in the kitchen, we turn on a lightbulb, releasing a fountain of
light waves. They radiate outwards, and as soon as their journey starts
they’re being modified by the world, reflected, refracted and absorbed until
perhaps our eyes pick up what’s left. And it isn’t just light that’s flowing
around us.
Humans are sociable animals. We hold our social networks together via
communication, sending and receiving signals from others. Our voice is one
of our most special features, a flexible musical instrument capable of
producing and shaping sound waves which are then broadcast through our
surroundings. No Brit could imagine making a hot drink without asking
others in the room whether they wanted to share the experience, and we ask
with sound. The others in the room pick up the signal with their ears, and
hearing the question will trigger a new flow of information inside their
bodies, splitting and rejoining and gathering meaning until their nerve fibres
instruct their vocal muscles to provide the appropriate response. Once the
message has come back to us, we alter the world appropriately, rearranging
the ceramics and metals in front of us.
Many different atoms make up our bodies, and wonderful though the
variety is, there are limits to what we can do directly because of the way
those atoms are arranged. But humans are experts at manipulating the world
to produce tools that can do what we can’t. We can’t hold water in our
hands as it boils, but a steel kettle can. We can’t turn a part of ourselves into
an air-proof container for dried leaves, but a glass jar will do that job. We
don’t have claws or a shell or tusks, but we can make knives and clothes
and can-openers. A ceramic container can hold a hot drink for us without
transmitting the heat energy to our vulnerable and sensitive fingers. Metals,
plastics, glass and ceramics are all our proxies in the world, helped by the
materials with biological origins: wood, paper and leather.
The kettle has held the water molecules while giving them energy in the
form of vibrations at the tiniest level. Now they’re jostling far more quickly
than before, and we transfer them to their new ceramic home. Frustratingly,
we can only see a hint of the splash of milk that bounces back up just after
we add it to the drink. It’s right there, right in front of you in plain sight, but
it’s virtually invisible because we just can’t process the signals fast enough.
You can no longer see the bottom of the cup: the liquid that had been partly
transparent is now opaque because light is bouncing off millions of tiny fat
droplets.
As we’re manipulating the world around us, we take for granted that we
are stuck to the floor by a force that’s manageable because our bodies have
evolved to cope with it. If the Earth’s gravity were stronger, we’d probably
need sturdier legs and we might find bipedal life difficult. If gravity were
weaker, we might have evolved to be taller, but life would be slower
because everything would take longer to fall. As we lift one leg to step,
we’re relying on the pull of gravity to make us fall forwards. We pivot
around our stationary foot and by the time we stop the fall with the stepping
foot, our whole body has moved forwards. Walking doesn’t work without
gravity, and our bodies have evolved to suit the Earth’s gravity. We’re just
the right size and shape for bipedal walking to work for us. As we pick up
the drink and move to the door, we’re using our own bodies as an upside-
down pendulum, swinging each leg forwards as we pivot about our other
foot and hip. The rhythm in our walking caused by that regular swinging
affects the liquid in the cup, forcing it to slosh with the same rhythm.
As we walk, we use fluid tucked inside our skulls to help with our
balance. The fluid sloshing around deep inside the tiny cavity of our inner
ear keeps going when we stop and lags behind when we start. Sensors on
the walls of that cavity feed that information into the giant connected
network of our brain, helping with the decisions about which muscle to
move next.
On this occasion, we reach a door, push it open with our free hand, and
step outside.
Earth
Outside, we can look sideways through the invisible atmosphere, at the rest
of the world. Our planet is a system made of five interacting components:
rocks, atmosphere, oceans, ice and life. Each one has its own rhythm and
dynamics, but the sumptuous variety that we see on Earth is the result of the
timeless dance that connects them. The same forces drive them all, and
there are similarities in surprising places. As we look out through the
invisible molecules that fill the sky, packets of air are shifting according to
their buoyancy. Air that has been warmed by the building we’ve just
stepped out of is rising because it’s less dense than the air around it.
Columns of rising air from warm ground could be a few kilometres tall,
taking perhaps five minutes or so to rise each kilometre. Cooler, denser air
is flowing in beneath to take its place, hauled downwards by Earth’s
gravity. These patterns of convection stretch out across the landscape we’re
looking at. Air is never completely still.
If we were looking sideways across the surface of the deep ocean, our
view might encompass similar buoyancy flows, also invisible. Cold, salty
water in the North Atlantic sinks downwards towards the centre of the
Earth, just like the cooler, denser air. Once it reaches the ocean floor, it
flows sideways across the seabed until it warms or mixes with less salty
water and floats back to the surface. In the sky, one cycle of floating up and
sinking down might take a few hours. In the ocean, one cycle might take
four thousand years, and the water will be carried around half the globe as it
happens.
And then, down below our feet right now, the rocks themselves are also
moving. The Earth’s mantle makes up most of the planet, a thick layer
between the outer core and the thin crust floating on top. It’s liquid but
viscous, sluggish and slow. This melt is being heated, both by the Earth’s
hot core and by the slow decay of radioactive elements buried deep inside
it. This shunting of energy around the deep rocks is happening now, beneath
all of us. As the hot mantle rock becomes buoyant, it floats upwards and
cooler rock sinks down to take its place. But molten rock at these
temperatures and pressures takes time to move. Deep below us, a mantle
plume might take a year to float upwards by 2 cm. A full cycle from the
bottom to the top and back again might take 50 million years. But the centre
of the Earth bows to the same physics as the atmosphere and the ocean,
continuously shifting heat from within to without.
A vast amount of heat energy is continuously moving outwards from the
centre of the Earth, but it’s utterly insignificant compared to the amount of
light energy from the Sun striking our planet. And in almost every
environment on Earth, tucked away in corners or dominating the landscape,
there is green. It might be a furtive veneer of moss on a brick wall or the
luxuriant biological architecture of a rainforest, but plants are everywhere.
Each leaf is the support structure for layers of chlorophyll-stuffed cells,
each one a tiny molecular factory turning sunlight and carbon dioxide into
sugar and oxygen. A fraction of the energy in the flood of light washing
over each leaf is captured and stashed away as sugar: fuel for the future.
Even on the calmest of sunny days, in a field where everything looks still
and unchanging, the plants are busy. One molecule at a time, they are
producing the oxygen that we breathe, enough to keep all the other living
things on Earth alive, enough to maintain an atmosphere that is 21 per cent
oxygen. These tiny molecular machines are continuously remaking a fifth
of our entire planet’s atmosphere. As we look sideways through the air,
we’re looking through the jostling molecular output of millions of ferns,
trees, algae, grasses and more, produced over thousands of years: the
bounty of a green army.
From our position on the ground, just outside our home, we can see only
a small fraction of our planet. Suppose we can levitate, and we’ll see far
more. As we rise through the atmosphere, the air molecules spread out.
Gravity is pulling them downwards, and it can only hold a very thin layer to
the Earth’s surface. As we rise past the top of the largest thunderstorm,
approximately 20 km up, 90 per cent of the molecules in the atmosphere are
beneath us. The deepest point in the ocean is 11 km below sea level, and
below that, there is dense rock for about 6,360 km before you get to the
centre. Without a rocket, we humans are confined to a vertical range of a
measly 30 km, playing on the edge of the giant planet we call home. A layer
of paint coating a ping-pong ball has the same thickness relative to the
sphere it covers.
At a height of 100 km, we are officially at the boundary between Earth
and outer space, and we can see the globe rolling past beneath us – green,
brown, white and blue, spinning in the blackness of space. From up here,
the scale of the ocean is shocking: a planet-sized shell made of one simple
molecule repeated over and over again. Water is the canvas for life, but only
fn1
in the Goldilocks zone, the energy range within which the molecules move
about as a liquid. Give those molecules extra energy, and their vibrations
will shake apart any complex molecules that they house. More energy still,
and they will float away as a gas, useless for protecting fragile life. At the
lower end of the Goldilocks range, as you reduce the energy, the vibrations
slow until the molecules must slot themselves into an ice lattice. Immobility
like that is the enemy of life. Even the process of building these inflexible
ice crystals can burst any living cell that contains them. Our planet is
special not just because it has water, but because that water is mostly liquid.
From our vantage point here on the edge of space, Earth’s most precious
asset dominates the view.
Perhaps, down there, as the Pacific Ocean glides past, a blue whale is
making sound waves, calling into the gloom. If we could watch that sound
travelling beneath the ocean surface, we would see it travelling outwards
like ripples on a pond, taking an hour to reach California from Hawaii. But
the sound is hidden in the water, and no evidence of it is visible from up
here. The oceans are filled with sound, overlapping pressure oscillations
pulsating outwards from breaking waves, ships and dolphins. The deep
rumblings of Antarctic ice can travel underwater for thousands of miles.
From our viewpoint on the edge of space, you would never know that any
of it was there.
Everything on the planet is spinning, travelling once around the Earth’s
axis every day. As winds travel across the spinning surface, they tend to
keep going in a straight line, although friction with the ground and
confinement by the air around them constrains their path. From up here, we
can see that the winds in the northern hemisphere tend to turn to the right,
relative to the ground, as they carry on in spite of the spin of the Earth. So
weather, especially the weather further away from the equator, spins.
Hurricanes rotate, and so do the smaller storms that we can see rolling
across the oceans. The eye of the storm is the hub of each wheel, and each
wheel must spin because the Earth spins.
Over Antarctica, thick snow clouds are gathering. Inside each one,
billions of individual water molecules exist as a gas, jiggling around with
the oxygen and nitrogen. But as the cloud cools, they are giving up their
energy and slowing down. When the most sluggish molecules bump into a
nascent ice crystal, they lock on, each in a fixed place in the ice lattice. As
the snowflake is buffeted up and down inside the cloud, the molecules on
all six sides of the original crystal find themselves in the same conditions,
and stick in the same way. Molecule by molecule, a symmetrical snow
crystal is built. After hours of slow growth, the crystal is large enough for
gravity to win the battle and it tumbles from the bottom of the cloud. Below
is the Antarctic ice sheet, the largest agglomeration of ice on Earth,
stretching sideways for thousands of kilometres and down for thicknesses
of up to 4.8 km. The accumulation of ice is so heavy that the continent itself
has been pressed downwards under the additional weight. But every
molecule of that white expanse fell in a snowflake, and the pile of
snowflakes has been growing for a long time. Some of the water here has
been frozen for a million years. In that time, the molecules have vibrated
about their fixed crystal lattice location continuously, but never fast enough
to become a liquid again. In contrast, the molecules being pushed out of
Hawaii’s volcanoes as lava are only just dropping below 600°C for the first
time since the Earth was formed, 4.5 billion years ago.
At the heart of Earth’s outer engine is the energy supply from the Sun. As
it heats up the rocks, ocean or atmosphere, or as it fuels sugar production in
plants, it’s pushing the engine away from equilibrium. As long as there’s an
imbalance in the distribution of energy, there is always the potential for
things to change. The movement energy of falling rain can erode mountains
as it splashes down on bare rock. The vast excess of heat energy at the
equator drives tropical storms, battering palm trees, redistributing water
from sea level to high mountains, and sending waves crashing on to
beaches. The energy stored in a plant will be used to build branches, leaves,
fruit and seeds, eventually running out of usefulness as low-level heat. Only
the seed will be left, a package of genetic information destined to restart the
cycle with new energy from the Sun’s fountain of light. Our planet lives
because of the constant injection of energy from above, feeding the engine
and preventing Earth winding down into stable, unchanging equilibrium.
From up here on the edge of space, we can’t see the tiny details, but we can
see the big picture: energy flows on to Earth from the Sun, trickles down
through the ocean, the atmosphere and life, and eventually carries on out
into space as the planet radiates heat away. The same amount of energy
goes in and comes out. But the Earth is a gigantic dam in the energy flow,
storing and using this precious resource in myriad ways before it’s released
to the universe.
As we drift back down to ground level, a beach now looks like a process
instead of a place, a patchwork of timescales and size scales. The ocean
waves are carrying energy from storms far out at sea. As they break on the
beach, they rattle the sand and rocks, grinding them together. One speck at a
time, the stone is chipped away, each pebble sculpted by millions of random
collisions. It takes a millisecond to remove one minuscule chip, but years of
slow attrition to make the pebble smooth. In geological time, a beach is
temporary. It will last only if the supply of new pebbles and sand is greater
than the loss as they wash out to sea. Over months and years, sand will shift
into the sea and back out again in response to the ocean. We love our tidal
beaches precisely because we can see the ebb and flow reshaping the sand
twice a day; it’s as though the slate is wiped clean, and we find the
simplicity of the newly smoothed sand satisfying. But this daily
remodelling hides the decadal shifts as our coastline grows and shrinks in
front of us. The life in the rock pools thrives on change, adapted to periods
of being high and dry alternating with spells of complete submersion.
Though a casual glance at a rock pool can give the impression of a museum
exhibit behind glass, in every pool there is a fierce battle for resources
going on. The resources on offer are all ultimately very simple: access to
the drips of energy oozing through the Earth’s system or the chance to
gather the molecular building blocks needed to construct a life. More than
anywhere else, a beach exemplifies the transience of life. When the energy
and nutrients are available to support life, rock pools flourish. During the
barren periods, life will be found elsewhere. Species evolve by altering their
use of the physical toolbox available to them one genetic mutation at a time.
Whether they’re harvesting energy, moving around, communicating or
reproducing, they are all just using the same principles in different ways.
Energy passes through, but the Earth itself is constantly recycled. Almost
all the aluminium, carbon and gold that make up our planet has been here
for billions of years, shifting from one form to another. It might seem that
after this long, these different substances should all be jumbled up, mixed
together in a giant planetary soup. But the physical and chemical processes
around us are continually sorting the pile, so that pockets of similar atoms
group together. Gravity allows liquids to drain through porous solids, so
that water soaks into the soil and joins huge underground aquifers while the
soil stays where it is. When vast blooms of tiny calcium-based marine
creatures live and then die at the ocean surface, it’s gravity that coaxes them
downwards so that they drift towards the ocean floor. The vast marine
cemeteries that sometimes form as a result in shallow seas compress, shift,
and become distinctive white limestone. Salt deposits are formed because
water molecules will evaporate easily to become a gas when they are given
more energy, but salts won’t. The lava produced at volcanic mid-ocean
ridges is far more dense than water, so it stays on the ocean floor, building
new crust. And life itself is constantly plucking materials from the world
around it, reshaping and reorganizing them, and then leaving the detritus to
be reused when it dies.
On a dark night, looking up at the sky, we see waves that have travelled
across our solar system or our galaxy or our universe to reach our eyes. For
millennia, light waves were our only connection to the rest of the universe,
the only reason we knew that there was anything else out there. A couple of
decades ago, we started to observe the thin streams of matter that reach us:
neutrinos and cosmic rays. And then gravitational waves came along, only
the third way that we have of touching the rest of the universe. In February
2016, it was finally confirmed that catastrophic astronomical events like the
merging of black holes also send out waves, ripples in space itself.
Gravitational waves have been passing through all of us our entire lives,
and we’re finally about to discover what we’ve been missing out on. The
light and gravitational waves whooshing past our planet weave a rich
tapestry that lets us map out our universe and then add an arrow labelled:
‘We are here.’
But on an average day on Earth, there are more immediate
considerations. Standing outside our home and watching the world go by is
a reminder of the gigantic system that we’re part of. We are a small sliver of
the life that keeps the system running in its current configuration. When
Homo sapiens first emerged, each human only had two life-support
systems: a body and a planet. But now there is a third.
This planet has been altered by many species, but only in the past few
thousand years has a single species knowingly rebuilt its environment to
suit itself. It is almost a single organism now, a sprawling planet-sized web
of interconnections between individual consciousnesses. Each individual is
almost entirely dependent on others in the system for survival, but still has
its own contribution to make. An understanding of the laws of physics is
one of the pillars holding up our society, and we could not manage our
transport, resource management, communication or decision-making
without it. Science and technology make possible the greatest ever
collective human achievement: our civilization.
Civilization
A. A. Michelson and E. W. Morley, ‘On the relative motion of the Earth and
of the luminiferous ether’, Sidereal Messenger, 6, 1887, pp. 306–10,
http://adsabs.harvard.edu/full/1887SidM....6..306M
Sindya N. Bhanoo, ‘Silvery fish elude predators with light-bending’, New
York Times, 22 Oct. 2012,
http://www.nytimes.com/2012/10/23/science/silvery-fish-elude-
predators-with-sleight-of-reflection.html?_r=0
Alexis C. Madrigal, ‘You’re eye-to-eye with a whale in the ocean: what
does it see?’, The Atlantic, 28 March 2013,
http://www.theatlantic.com/technology/archive/2013/03/youre-eye-to-
eye-with-a-whale-in-the-ocean-what-does-it-see/274448/
Leo Peichl, Günther Behrmann and Ronald H. H. Kröger, ‘For whales and
seals the ocean is not blue: a visual pigment loss in marine mammals’,
European Journal of Neuroscience, 13 (8), 2001, pp. 1520–8
Jeffry I. Fasick et al., ‘Estimated absorbance spectra of the visual pigments
of the North Atlantic right whale (Eubalaena glacialis)’, Marine Mammal
Science, 27 (4), 2011, pp. E321–E331
University of Oxford, press pack for Marconi exhibition:
https://www.mhs.ox.ac.uk/marconi/presspack/
Bill Kovarik, ‘Radio and the Titanic’, Revolutions in Communication,
http://www.environmentalhistory.org/revcomm/features/radio-and-the-
titanic/
RMS Titanic radio page, http://hf.ro/
Yannick Gueguen et al., ‘Yes, it turns: experimental evidence of pearl
rotation during its formation’, Royal Society Open Science, 2 (7), 2015,
150144
Chapter 6: Why don’t ducks get cold feet?
The page references in this index correspond to the printed edition from
which this ebook was created. To find a specific word or phrase from the
index, please use the search feature of your ebook reader.
AC (alternating current), 236–7
AC/DC adaptor, 237
Adams, Douglas, 204
adiabatic heating, 23
air: atoms, 13–14; baking, 17–20; breathing, 256–7; bubbles, 76–9; buoyancy, 57–9, 62; buoyancy
flows, 264; convection currents, 61, 264; dogs panting, 114; floating particles, 75–6; hot, 60–1;
katabatic winds, 21, 22–3; molecules, see air molecules; pockets, 40–1, 85–6; pressure, 12, 23–6,
28, 40; resistance, 43–4, 209, 222; sucking, 26–8; swirls of warm and cold, 1–2; thunderstorm,
131–3; viscosity, 73, 74; water molecules, 161–2, 221; weather, 36–7; whales, 14–15
air conditioning, 274
air molecules: air pressure, 24–6; Antarctica, 21; atmosphere, 265; bees flying, 222; breathing, 256–
7; eggs, 58; elephants drinking, 27–8; katabatic winds, 23; speed, 23; thunderstorms, 36–7;
vacuum pump, 24–5
alcohol molecules, 164, 175
Amundsen, Roald, 20, 168
Antarctica: Andes, 45; ice sheet, 267; katabatic winds, 21, 22; mid-Atlantic ridge, 250; polar
explorers, 21, 22, 168; sounds of ice, 266–7; thermohaline circulation, 63–4; winter temperature,
176–7
anthocyanins, 8–9
Archimedes, 56
Arctic Ocean, 165–8, 247
Aristotle, 24
Asimov, Isaac, 157
astronauts, 192–3, 205–6
Atlantic Ocean: buoyancy flows, 264; experiments at sea, 228, 231; mid-Atlantic ridge, 250–1;
thermohaline circulation, 63–4; Titanic radio messages, 139–40, 150; weather, 1; width, 252
atoms, 13–14; Brownian motion, 158–9; conductivity, 181–2; crystals, 156–7, 169–70; DNA, 260;
existence, 14, 69, 157–8, 182; glass, 171–5; ice crystals, 176–7; ice cubes, 169, 184; ions, 155n;
magnetic fields, 217; nucleus, 13, 219; size, 69
Earth: atmosphere, 36, 146, 159, 192, 265; axis, 203, 205, 211–12, 267; centre, 45–8, 56, 189, 197,
264–5; circumference, 47; coldest recorded temperature, 176; continental drift, 249–52; energy
supply from Sun, 265, 268, 272, 273, 275; formation, 268; global warming, 176, 275–6; gravity,
41, 44, 45–6, 48–50, 56, 204–5, 263–4; ‘greenhouse effect’, 146–7; life-support system, 263–71;
magnetic field, 246–8, 250–2; mantle, 264; mass, 48, 49, 50, 216; oceans, 47, 62, 64, 129, 176,
266, 272–3; plate tectonics, 251–2; poles, 164; radio waves, 141; recycled, 269–70; seasons, 211–
12; shape, 196–7; spin, 186–7, 196–7, 203, 205, 211–12, 267; start of life on, 119; tallest
mountain, 197; temperature, 146–7, 176; weather, 1, 131, 159, 267
earthquakes, 98–9, 115–16
eggs: freshness, 58; hard-boiled, 32; spinning raw and boiled, 5, 209–10
Einstein, Albert, 46n, 49, 127, 158–9
electric(al): batteries, 109, 213, 231; charge, 131, 219, 222–4, 226, 234, 235, 240; circuits, 224–5,
227, 230–1; conductors, 221, 224, 225, 226–7, 234, 247; current, 100, 131, 235–7, 240, 243–4,
253–4; energy, 224, 228, 232–3, 236, 252, 254; fields, 125, 225–7, 230, 231, 233–4, 236, 239–42;
insulators, 220, 224; kettle, 234–6; plug (three-pin), 234; power cable, 230; resistance, 232; shock,
218–19, 221, 222, 252; signals, 100, 226–7, 252, 259; voltage, 234–6, 237n, 239–40, 243
electricity: bees, 221–2, 224; flywheels, 213–14; grid, 187, 213–14, 217, 224, 253; magnetism, 241,
243–5, 252–5; mains, 233, 236–7; power stations, 254; static, 222–4, 225
electromagnetic spectrum, 142
electromagnetism, 142, 217–18, 252–5
electromagnets, 241, 244–5, 247
electronics, 237–8, 246
electrons: atomic structure, 13, 155n, 181; cathode ray tube (CRT), 239–40; direct current (DC), 236;
electric charge, 219–20, 222–4; electric circuit, 224–5, 227, 229–31, 233–6; electric shock, 221;
electromagnetism, 252–4; electronics, 238–42; flat screens, 243; heat conduction in metals, 181–2,
236; heat generation, 232, 236; power network, 272–3; toasters, 244–5
elephants: infrasound, 136; trunks, 26–8
equilibrium: batteries, 109–10; controlling, 105–7, 118–19, 273; flow towards, 117–20, 179; fossil
fuels, 109, 276; Hoover Dam, 107–9; pigeons, 102n; sloshing of liquid in mugs, 110–12; solar
energy, 268, 276; swinging, 112; whales, 15
evaporation: solar energy, 273; sweat, 81, 113–14; TB bacteria, 74; towels, 86; transpiration, 90;
water molecules, 11–12, 66, 257–8, 270; wet clothes, 161–3
expansion, thermal, 175–6
explosions, 11–13
gas molecules: distance between, 160–1; Earth’s atmosphere, 36; focaccia, 19–20; global warming,
275–6; popcorn, 12; push of, 28, 31–2, 34; viscosity, 73
gases: behaviour, 13–14, 20; candles, 59–61; fizzy lemonade, 39–40; flow, 59; frying cheese, 163–4;
greenhouse, 146–7, 176; ideal gas law, 20–1, 22, 37; laws, 13, 36, 38, 183n; pressure, 19–20, 25;
rockets, 32–6; snow clouds, 267; steam engines, 28–32; transition from gas to liquid, 164–5;
viscosity, 73; water evaporation, 113–14, 161–2, 270
glass: blowing, 170–2, 173–4; old windows, 172–3
goggles, 80–3
Goldilocks zone, 266
gravitational waves, 270
gravity: air molecules, 264, 265; astronauts, 192–3, 205–6; bacteria, 74; bubbles, 58, 59; canal locks,
106; candles, 59–61; convection patterns, 264; cycling on an indoor track, 189–90, 197; falling,
42–4, 73n; force, 41–2, 48–9, 70; glassblowing, 170, 171; ketchup, 93–4; kitchen scales, 49–50;
milk, 71–3, 191; ocean currents, 62–4; pizza, 196, 197; raisins in fizzy lemonade, 40–1; redwood
trees, 89–90; seedlings, 46–7; shipboard life, 44–6, 231; sloshing liquid in mug, 110–11;
snowflakes, 267; Sputnik, 203–5; toast falling, 207–8; Tower Bridge, 51–3, 59; walking, 263;
water soaking through soil, 269–70; water supply, 273
‘greenhouse effect’, 146–8
greenhouse gases, 147, 176
gyroscopes, 5–6, 186
ice: Antarctica, 21, 176–7, 266–7; crystals, 165, 176, 266, 267; cubes, 169, 184; ducks on, 177–80;
sea, 63, 165–8; spinning skaters, 54
inertia, 54, 70, 80, 90
information: aural, 262; decision making, 263; flow of, 152–3, 259, 274, 276–7; network, 274, 276;
portable, 271; selecting wavelength, 144; sharing, 271; smells, 257; stored, 252, 260; travelling,
260; visual, 102, 260–1; wave-based, 125, 133, 136–42, 274
infrared wavelengths, 134–5, 136, 144, 146–7, 243
infrasound, 136
International Space Station, 35, 205–6, 277
ions, 155–6, 260
iron, 182, 215–17, 244, 246–7, 250
Jeanette, USS, 166–7
Magdeburg sphere, 25
magic, 217, 241–2
magnetic field: changes in, 247–51; Earth’s, 245–51; electrical grid, 253; electromagnets, 244–5;
electrons, 231, 241–3, 244; light waves, 125; north and south poles, 216–17, 246, 250–1;
permanent, 245; temporary, 245
magnets: coin sorting, 215–17, 219; electromagnets, 244–5, 247; electronics, 239, 240–2; fridge, 10,
216, 217, 225, 245, 246; north and south poles, 216–17, 223, 246; permanent, 245; toasters, 243–5,
247; wind turbine, 253–4
Matthews, Drummond Hoyle, 251
methane, 58, 120, 146, 147, 164
Mexico City earthquake, 115–16
Michelson–Morley experiment, 125n
microfluidics, 91, 92n
microscopes, 67–9, 157–8
microwave wavelengths, 136, 144, 145n
milk: adding to drink, 1, 100, 262; cream, 70–2, 74, 75n, 76, 191; homogenized, 72–3, 74;
refrigeration, 184; single drop of, 100; wiping up spilt, 84–6
mobile phones, 148, 150–4
molecules: air, see air molecules; alcohol, 164, 175; atoms, 13–14; behaviour, 17; Brownian motion,
158–9; complex, 119, 266; conductivity, 181; DNA, 260; evaporation, 161–3; fat, 257; formation,
119, 220n; gas, see gas molecules; greenhouse effect, 147; heat transfer, 178–9; human body, 260,
274; lightning, 131; long, 94, 98; manipulation, 69; melting, 164; opsin, 261; oxygen, 256–7; pH
value, 8–9; protein, 71; refrigeration, 182–3; sugar, 71, 156; temperature, 267–8; water, see water
molecules; wax, 60
momentum: angular, 5, 54n, 208n, 209, 211, 214; of fluid, 83n, 111
monsoon, 159–62
Morley, Lawrence, 251n
mug of tea: sloshing, 110–13; stirring, 2, 185–6
ocean: colour, 129–30; depth, 14, 62, 265; global warming, 176; katabatic winds, 21–2; movement of
water, 62–4, 264; saltiness, 62–3; sound waves, 138–9, 266–7; surface, 44, 122, 129, 138, 162,
227; thermohaline circulation, 64; underwater mountains, 250–1, 270; waves, 122–5, 127, 268–9
Ocean Conveyor Belt, 64
oxygen: breathing, 256–7; candle burning, 60; dogs panting, 114; Earth’s atmosphere, 265; fish
breathing, 58–9; melting, 164; snow clouds, 267; speed of atoms, 13; water molecules, 78, 103;
whale’s lungs, 15–16
oysters, 148–50
particles: charged, 239–40, 242, 253, 259; floating, 75–6; jiggling, 158–9
pearls, 148–50
pelicans, 121–2
pendulums, 112, 115, 116–17, 263, 277
Penzias, Arno, 145n
Perrin, Jean, 159
pH value, 8–9, 257
Phillips, Jack, 140, 141
photosynthesis, 88–90, 118
pigeons: droppings, 145n; head-bobbing, 101–3
pivot, 51–3, 199–200, 202, 207–9, 263
pizza, 193–6, 197
platypus, 225–6
Pockels, Agnes, 77–9
popcorn, 11–13, 20, 38
pressure cookers, 12
protons, 219, 222, 242
pulse, 258
quicksand, 100
water: boiling, 180–1, 235–6, 262; bubbles, see bubbles; buoyancy, 56–8, 264; canal locks, 105–6;
capillary action, 85, 88, 90–2; colour, 129–30; conducting electricity, 226–7; crystals, 165–6;
cycle, 272–3; depth, 15–16; evaporation, 66, 86, 90, 113–14, 161–4, 257, 273; freezing, 165, 167,
169–70; Hoover Dam, 107–9; human body, 257–8; liquefaction, 99; molecules, see water
molecules; oceans, 62–4; popcorn, 11–13; rain, 37, 103–4, 107, 131, 160, 268; refraction, 128;
salty, 62–4, 165, 227, 264; sound waves, 137–9, 148, 266–7; steam engines, 30–1; surface tension,
77–83; temperature, 174–6; tonic, 3; transition from liquid to solid and back again, 164–5, 169–70;
viscosity, 71, 73; waves, 121–5, 127, 133, 148
water molecules: capillary action, 84–6, 89; coffee puddle, 65–6; evaporation, 11–12, 66, 257–8, 270;
frying cheese, 163; Goldilocks zone, 266; humid air, 221; ice, 165, 167, 169, 267; raindrops, 103–
4; snow, 267; speed, 13; steam, 12, 30; surface tension, 78–9, 81, 83, 89; thermal energy, 181, 262;
wet clothes, 161–2, 163
wavelengths: all waves, 125; light, 128n, 130, 134–6, 139, 143–6; microwave, 144, 145n; radio, 139,
144–5, 151–3, 274; sound, 138–9; water, 124–5
waves, see gravitational waves; light waves; radio waves; sound waves; water
Wegener, Alfred, 249–50
weightlessness, 192, 205–6
wellie throwing, 198–202
whales: colour-blind, 139; communication, 137, 138–9, 148, 266; diving, 14–16, 36, 56; feeding, 14–
16, 126, 165; sonar, 14–15
Whitesides, George, 92
wildfires, 22–3
Wilson, Robert, 145n
wind turbines, 253–4
windows, glass, 173
winds: buildings vibrating, 115; katabatic, 21–3; monsoon, 159–60; ocean currents, 63–4; spinning,
267; storms, 37, 123–4
‘wise women’, 8, 9
Helen regularly presents BBC programmes on physics, the ocean and the
atmosphere – recent series include Colour: the spectrum of science, Orbit,
Operation Iceberg, Supersenses and Dara O’Briain’s Science Club – as
well as programmes on bubbles, the Sun and our weather. She is also a
columnist for Focus magazine, was shortlisted for PPA columnist of the
year in 2014, and has written numerous articles for the Guardian.
TRANSWORLD PUBLISHERS
61–63 Uxbridge Road, London W5 5SA
www.penguin.co.uk
Helen Czerski has asserted her right under the Copyright, Designs and
Patents Act 1988 to be identified as the author of this work.
Every effort has been made to obtain the necessary permissions with
reference to copyright material, both illustrative and quoted. We apologize
for any omissions in this respect and will be pleased to make the
appropriate acknowledgements in any future edition.
A CIP catalogue record for this book is available from the British Library.
1 3 5 7 9 10 8 6 4 2
1: Popcorn and rockets
fn1
We’ll get to the meaning of absolute temperature in chapter 6.
fn2
This substitution is not the recommended way of doing science today.
fn3
We don’t know how much of the air Otto’s vacuum pump removed. It won’t have been all the air,
but it must have been a substantial proportion of it.
fn4
And also when we breathe. Every breath you take gets into your lungs because the atmosphere
pushes it there.
fn5
If you’ve ever wondered what Thomas the Tank Engine’s tank is all about, it’s all about the water.
The water can be stored in a separate carriage with the coal (a tender) or it can be stored in a tank that
sits around the engine. Thomas stores his water around the engine – that’s why he was rectangular –
and so he’s a tank engine.
fn6
The Indian Airmail Society also experimented with rocket mail around the same time. They
managed 270 flights, sending parcels as well as letters, but never established it as a long-term
success. In the end, rocket mail was never going to be able to compete with the regular ground-based
mail delivery systems on reliability and cost.
3: Small is beautiful
fn1
Sorry. Really. If it helps, what I’m about to say is just as true for instant coffee, so you don’t ever
need to waste shots of fancy coffee on science.
fn2
We can go a long way into the world of the small without having to deal with the strangeness of
quantum mechanics. That really kicks in when you explore what’s happening to individual atoms and
molecules, and there’s an awful lot that’s bigger than that and still smaller than what we can see. That
middle bit is interesting because we can understand it intuitively (something which is impossible by
definition when it comes to the rules of the quantum world), even though we can’t see it clearly.
fn3
As someone who loves the variety and spice of life, I’m always a bit sad when I see this word.
Making everything the same definitely has its uses, but sometimes it does just sound as though it’s
taking the fun out of life. Especially if you’re a blue tit.
fn4
Their rise is slowed even more by the extra protein coat that surrounds each of the new smaller
globules; this weighs them down a bit, so they’re even less buoyant than they were before. This has
been measured in quite a lot of detail. You’d be surprised at how much science has gone into a pint of
milk.
fn5
If you’re interested in reading more on this, the biologist J. B. S. Haldane wrote a very famous
short essay in the 1920s called ‘On being the right size’. It’s here: http://irl.cs.ucla.edu/papers/right-
size.html. The most memorable quote from it is very painfully true: ‘To the mouse and any smaller
animal it [gravity] presents practically no dangers. You can drop a mouse down a thousand-yard mine
shaft; and, on arriving at the bottom, it gets a slight shock and walks away, provided that the ground
is fairly soft. A rat is killed, a man is broken, a horse splashes.’ As far as I’m aware, no one has
actually done this specific experiment. Please don’t be the first. And certainly don’t blame me if you
do.
fn6
If you keep stirring your milk, the cream won’t rise to the top because it keeps getting mixed back
in. The same principle applies here – particles don’t sink down very far because they keep getting
mixed back in by air currents that move faster than they’re falling.
fn7
Especially mine. I am a bubble physicist, after all.
fn8
You can see this effect for yourself if you put a droplet of water on something that is fairly
hydrophobic – a tomato does the job. The droplet will sit up, mostly off the surface. Then just touch
it with a cocktail stick with a tiny bit of detergent on the end, and the drop will immediately spread
out sideways. I recommend washing the detergent off the tomato before eating it.
fn9
This balance – how much water is attracted to a solid surface compared with how much it’s
attracted to itself – helps with all sorts of problems. The most important one for any Brit is the
question of why some teapots dribble from the spout as you finish pouring, sending tea down the side
of the pot and on to the table rather than into the cup. The answer is that the teapot is just too
attractive to water. As the flow slows, the forces sticking the water to the spout dominate over the
momentum carrying the water forwards. You can solve this by having a hydrophobic teapot, one that
doesn’t attract the tea at all. Sadly, at the time of writing, no one seems to be selling them.
fn10
It was actually an apology. On a trip to Krakow, I had promised a fabulous dinner in the Jewish
quarter of the city, but this was before the days of smartphones, so I got lost. I led twelve hungry
people on a merry dance around many dark and empty streets, failing to find any restaurants at all,
never mind the excellent ones I was aiming for. We ended up eating in a McDonalds instead. I felt
that apple pie was the least I could do to make up for this.
fn11
Of course, the fat and protein and sugar in the milk don’t evaporate, so they’re left behind and
the towel still needs a wash.
fn12
The clock tower of Westminster, the one that houses Big Ben, is 96 metres tall. These trees really
are giants.
fn13
One nanometre really is tiny – there are a million of them in one millimetre.
fn14
But there’s a limit. To increase the tension in the water to pull it up further, the stomata must get
smaller. And smaller stomata let in less carbon dioxide, so there is less raw material for
photosynthesis. Theory suggests that the tallest a tree could possibly grow is 122–130 metres,
because beyond that it wouldn’t be able to take in enough carbon dioxide to do any actual growing.
fn15
There’s also some evidence to suggest that the fog might actually go the other way too – entering
the stomata to keep them full of water, not just preventing evaporation.
fn16
These devices go by the catchy name of ‘microfluidic paper-based electrochemical devices’ or
uPADs for short. A non-profit organization called Diagnostics for All has been set up to move this
idea into the real world.
4: A moment in time
fn1
This behaviour is called ‘shear-thinning’ and it’s also handy for snails, as we’ll see shortly.
fn2
Of course, there is a third option: that the snail had been an egg or a juvenile hiding inside the
compost. But it was pretty large, and I couldn’t imagine it growing so big in such a short time.
fn3
There’s a genuinely funny bit in Frost’s paper when he describes what happened when they
accidentally set the treadmill to a very low speed. It’s not often that I’d quote a scientific paper for
comic effect, but in this case it’s absolutely justified: ‘After completing the filming of a particular
bird, the treadmill was inadvertently turned to a very slow setting instead of completely off as
intended. After a short time we noticed that the bird’s head was slowly and progressively pushed
forward until it eventually toppled over. Further observations indicated that toppling, or extreme
changes in posture, could also be produced by very slow forward (opposite direction to that eliciting
normal walking) treadmill movements. It appeared that the extremely slow (imperceptible to us)
speed of the treadmill was not sufficient to induce walking in the bird, but was sufficient to stabilize
its head even though this sometimes resulted in loss of equilibrium.’
fn4
When I first moved to the American south-west, I couldn’t shake off a nagging curiosity about
exactly where all the water came from in this dry environment. The book that answered many of my
questions (and tells the fascinating story of the battles over water supply in that area) is Cadillac
Desert by Marc Reisner, and I highly recommend it. California is suffering from a severe drought as I
write this, and the tough decisions about how to deal with it cannot be delayed any longer.
fn5
There is actually another solution: start drinking cappuccino. Having a foam layer on top dampens
down the oscillations a lot, so foam-covered drinks don’t slosh as easily. This is also useful in the
pub. Beer snobs may not like too much of a head, but at least it stops them spilling their drink.
fn6
There are also two smaller pendulums that help with this, just below the main one.
5: Making waves
fn1
One of the unintentional discoveries from my time at sea is that the best way to wind up a bird
enthusiast is to casually ask about seagulls. There are gulls (lots of different types), and some of them
live in or on or near the sea. But there is no such thing as a seagull. Bird enthusiasts will either spend
hours explaining this to you or leave in a huff.
fn2
If you get the chance to see them from the side, you’ll see that they’re actually going round in
small circles. The point is that they’re not travelling with the wave.
fn3
Other Pacific islanders, most notably the Tahitians, also had surfboards. However, it seems that
they only lay or sat on them. The Hawaiians pioneered the idea of standing up on the board, and so
‘surfing’ as we understand it today.
fn4
Experiments showing that light behaves like a wave were relatively straightforward. It took an
extremely clever experiment the size of the Earth’s orbit around the Sun to reveal the most counter-
intuitive thing about light: there isn’t any ‘stuff’ that’s doing the waving. Instead, the waves travel as
disturbances in electric and magnetic fields. The test became known as the Michelson–Morley
experiment, and it’s one of my favourites of all time, because it’s simple to understand, it’s extremely
elegant, and it used our whole planet as a vehicle to test their hypothesis.
fn5
Like many materials, diamond slows down different colours of light – different wavelengths – by
different amounts. So part of the sparkle comes from the diamond separating out the colours, as well
as bouncing them back at you.
fn6
It would be interesting to see the colour chosen by children to draw water in a culture that doesn’t
have this habit. I think that we identify water as blue because we know about the oceans, and we
have aerial photography and very clean swimming pools. But few cultures had those things until
recently. Are there enough hints that they would unconsciously colour it blue? Or is that entirely a
learned habit?
fn7
Although astronomers haven’t always believed that it’s the majesty of the cosmos they’re looking
at. In 1964, Robert Wilson and Arno Penzias detected waves from the sky at microwave wavelengths
that shouldn’t have been there. They spent a long time trying to work out which bit of the sky or their
telescope was messing up the measurement, sure that something was generating extra microwave
light. They also cleared out some nesting pigeons from the telescope, along with their droppings
(euphemistically described as ‘white dielectric material’ in the paper they wrote). The unwanted
background light persisted. It eventually turned out to be the signature of the Big Bang, some of the
most ancient light in the universe. There is something special about an experiment that has to be very
careful to distinguish between the after-effects of pigeon poo and the after-effects of the formation of
the universe.
fn8
Which has very little to do with how a real greenhouse does anything.
fn9
That’s why mobile phones are called cellphones in US English – the network is cellular.
9: A sense of perspective
fn1
Not too hot, not too cold, but just right.