0% found this document useful (0 votes)
704 views221 pages

147214094X Elemental

Uploaded by

Nguyễn Hiệp
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
704 views221 pages

147214094X Elemental

Uploaded by

Nguyễn Hiệp
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 221

Copyright © 2019 by Tim James

Published in 2019 by Abrams Press, an imprint of ABRAMS. All Rights Reserved.


No part of this publication may be reproduced, stored in a retrieval system, or
transmitted in any form or by any means, mechanical, electronic, photocopy,
recording, or otherwise without permission in writing from the publisher.
Cataloging-in-Publication Data is available from the Library of Congress
ISBN: 978-1-4683-1702-2
Abrams books are available at special discounts when purchased in quantity for
premiums and promotions as well as fundraising or educational use. Special
editions can also be created to specification. For details, contact
specialsales@abramsbooks.com or the address below.
Abrams Press® is a registered trademark of Harry N. Abrams, Inc.

ABRAMS The Art of Books


195 Broadway, New York, NY 10007
abramsbooks.com
Dedicated to
the students of Northgate High School
Contents

INTRODUCTION A Recipe for Reality

CHAPTER ONE Flame Chasers

CHAPTER TWO Uncuttable

CHAPTER THREE
The Machine Gun and the
Pudding

CHAPTER FOUR
Where Do Atoms Come
From?

CHAPTER FIVE Block by Block

CHAPTER SIX
Quantum Mechanics Saves
the Day

CHAPTER SEVEN Things that Go Boom

CHAPTER EIGHT The Alchemist’s Dream

CHAPTER NINE Leftists

CHAPTER TEN Acids, Crystals, and Light

CHAPTER ELEVEN It’s Alive, It’s Alive!

Nine Elements that Changed


CHAPTER TWELVE the World (and One that
Didn’t)
APPENDIX I Sulfur with an “f”

APPENDIX II Half a Proton?

APPENDIX III Schrödinger’s Equation

APPENDIX IV Neutrons into Protons

APPENDIX V The pH and pKa Scales

APPENDIX VI Groups of the Periodic Table

Acknowledgments

Notes

Index
INTRODUCTION

A Recipe for Reality


Fourteen billion years ago, our Universe decided to begin. We
don’t know what came before (if there was a before), we just
know it started stretching in every direction and has been
doing so ever since.
In the first few nanoseconds after the big bang, all of reality
was a glowing soup of particles, frothing at temperatures
millions of times hotter than the Sun. As everything spread
out, however, things cooled, particles stabilized, and the
elements were born.
Elements are the building blocks nature uses for cosmic
cooking; the purest substances making up everything from
beetroot to bicycles. Studying the elements and their uses is
what we call chemistry, although sadly that word has come to
mean something sinister for many people.
A writer on a popular health website was recently moaning
about “chemicals in our food” and what we can do to keep
food “chemical free.” These scaremongers seem to think that
chemicals are toxins created by lunatics in lab coats, but this
view is far too narrow. Chemicals aren’t just the bubbling
liquids you see in test tubes: they are the test tubes themselves.
The clothes you’re wearing, the air you’re breathing, and
the page you’re currently reading are all chemicals. If you
don’t want chemicals in your food then I’m afraid it’s too late.
Food is chemicals.
Suppose you mix two parts of the element hydrogen with
one part oxygen. In scientific notation, you’d write that as
H2O, water, the most famous chemical in the world. Chuck in
a bit of the element carbon and you get C2H4O2—household
vinegar. Multiply each of those ingredients by three and you’ll
get C6H12O6, more commonly known as sugar.
The only difference between cooking and chemistry is that
while a recipe might specify a vegetable, chemistry wants to
go deeper and find out what the vegetable itself is made of.
There’s practically no limit to what you can describe once you
know the elements involved. Consider this beast for example:1
It looks like something you might find in a barrel of toxic
waste but it’s the chemical formula for a human being. You
have to multiply each number by seven hundred trillion, but
those are the correct chemical ratios for one human body. So,
if you hear someone say they distrust chemicals, feel free to
reassure them. They are a chemical.
Chemistry is not an abstract subject happening in dingy
laboratories: it’s happening everywhere around us and
everywhere within us.
In order to understand chemistry, therefore, we have to
understand the periodic table, that hideous thing you probably
remember hanging on the wall of your chemistry classroom.
Glaring down at you with all its boxes, letters, and numbers,
the periodic table can be intimidating. But it’s nothing more
than an ingredients list, and once you’ve learned to decode it,
the periodic table becomes one of your greatest allies in
explaining the Universe.
So, yes, the periodic table is seriously weird and seriously
complicated, but so is the rest of nature. That’s what makes it
worth studying. That’s what makes it beautiful.
CHAPTER ONE

Flame Chasers
THE MOST FLAMMABLE SUBSTANCE EVER MADE

Chemistry really began when we mastered our first reaction:


setting fire to stuff. The ability to create and control fire helped
us to hunt, cook, ward off predators, stay warm in winter, and
manufacture primitive tools. Originally, we burned things like
wood and fat, but it turns out that most substances are
combustible.
Things catch alight because they come into contact with
oxygen, one of the most reactive elements out there. The only
reason things aren’t bursting into flame all the time is that
while oxygen is reactive it needs energy to get going. That’s
why starting a fire also requires something like warmth or
friction. Oxygen has to be heated in order to combust.
The most flammable chemical ever made, though, far
worse than oxygen, was created in 1930 by two scientists
named Otto Ruff and Herbert Krug.1 Meet chlorine trifluoride.
Made from the elements chlorine and fluorine in a one-to-
three ratio, chlorine trifluoride is unique in being able to ignite
literally anything it touches, including flame retardants.
A green liquid at room temperature and a colorless gas
when warmed, ClF3 will set fire to glass and sand. It will set
fire to asbestos and Kevlar (the material from which
firefighters’ suits are made). It will even set fire to water itself,
spitting out fumes of hydrofluoric acid in the process.2
There are very few instances of ClF3 being used, though,
because it has the inconvenient property of setting fire to
almost anything with which it comes into contact. It takes a
special kind of maniac to think, “Hmm, I’ll give that a go.”
The most spectacular ClF3 incident happened on an
undisclosed date at a chemical plant in Shreveport, Louisiana.
A ton of it was being moved across the factory floor in a
sealed cylinder, refrigerated to prevent it reacting with the
metal. Unfortunately, the cold temperature made the cylinder
brittle and it cracked, spilling the contents everywhere. The
ClF3 instantly set fire to the concrete floor and burned its way
through over a meter in depth before extinguishing. The man
moving the cylinder was reportedly found blasted through the
air 150 meters away, dead from a heart attack. That was
refrigerated chlorine trifluoride.3
During the 1940s, a few cautious attempts were made to
use it as a rocket fuel, but inevitably it kept setting fire to the
rockets themselves so the projects were abandoned.
The only people who made a serious attempt to harness its
power were the Nazi weapons researchers of Falkenhagen
Bunker.4 The idea was to use it as a flame-thrower fuel, but it
set fire to the flame-thrower and anyone carrying it so, again,
it was deemed unusable.
Just think about that. Not only will it set fire to water,
chlorine trifluoride is so evil even the Nazis didn’t mess with
it. What makes it so potent?
The answer is that fluorine behaves in a very similar way to
oxygen but needs less energy to get started. It’s the most
reactive element on the periodic table and effectively out-
oxygens oxygen at breaking other chemicals down. So, when
you combine it with chlorine, the second most reactive
element, you get an unholy alliance that starts fires without
encouragement.

FIRE FROM WATER

The Greek philosopher Heraclitus was so enamored with fire


he declared it to be the purest substance—the basic matter
from which reality was made. According to him, everything
was somehow made from fire in one form or another. Fire was,
in other words, elemental.
It’s an understandable assumption to make since fire does
appear to possess magical properties. Then again, Heraclitus
lived on a diet of nothing but grass and tried to cure himself of
dropsy by lying in a cow shed for three days covered in
manure … after which he was eaten by dogs.5 So perhaps we
don’t need to take Heraclitus’s views too seriously.
The reason it was so difficult to identify elements in the
ancient world was because, unknown to the early
philosophers, very few elements occur in their pure state. Most
of them are unstable and combine to form element fusions
called compounds.
It works a bit like a singles’ bar. Each person is unhappy on
their own so they link up with others to form stable pairings.
At the end of the evening, most individuals have formed
compounds leading to greater stability all around. Only a
handful of elements like gold, which doesn’t mind being
single, remain in their native state.
Almost everything we come across in nature is a
compound, so while something like table salt may look pure,
the game is being rigged. Table salt is actually a compound of
sodium and chlorine—the true elements.
You’ll never find a lump of sodium in the ground or a cloud
of chlorine drifting on the breeze because both are violently
reactive. This makes them virtually undetectable, especially if
you’re working with the crude lab equipment of the first
millennium.
There’s also the fact that many elements are shockingly
rare. Take the element protactinium used in nuclear physics
research; the entire global supply comes from a single flake,
weighing 125 g owned by the UK Atomic Energy Authority.6
With the odds stacked against them, Greek philosophers had
no chance of getting things right.
It wasn’t until the late seventeenth century that a German
experimenter named Hennig Brandt proved everyday
substances had elements locked inside them and most of the
stuff we thought to be pure, wasn’t at all.
On an unknown night in 1669, Brandt was boiling vast
quantities of urine in his lab (you’ve got to have a hobby),
probably because urine is gold-colored and he was hoping to
make a fortune by solidifying it into the precious metal.
After many hours of what must have been unpleasant work,
Brandt was finally left with a thick red syrup and a black
residue similar to the gunk you get after burning toast. He
mixed these two things together and heated the mixture once
more. What happened next made no sense.
His mixture of urine syrup and cooking schmutz suddenly
formed a waxy solid, which smelled powerfully of garlic and
glowed blue-green. Not only that, it was extremely flammable
and gave off blinding white light as it burned. He had
somehow extracted fire from water.
Brandt named his chemical phosphorus from the Greek for
light-bringer, and spent the next six years experimenting with
it in secret. And it wasn’t a fun six years, either. Each 60-g
batch of phosphorus required five and a half tons of urine to be
boiled.
Eventually, running out of his wife’s money, Brandt went
public with the discovery and began selling phosphorus to
Daniel Kraft, one of the first science popularizers, who took it
around Europe giving demonstrations to amazed royals and
scientific institutions.7
Brandt, however, kept the method of extraction a closely
guarded secret. Although how nobody figured it out has
always been a puzzle. He must have had one hell of a cover
story to explain why he wanted all that urine.
Nowadays we understand exactly what was going on in
Brandt’s methods. The human body’s recommended intake of
phosphorus is between 0.5 and 0.8 g a day, but since
everything we eat contains it, we tend to consume over twice
that amount. All this excess is passed into the urine and Brandt
was just boiling everything else away.
His discovery marked a crucial moment for chemistry
because the extracted phosphorus was so markedly different
from its source. Urine doesn’t glow in the dark (sadly) but it
obviously contains a chemical that does. It was proof there
were chemicals hiding in plain sight. The elements weren’t out
of reach.

THE MEN WHO PLAYED WITH FIRE

At the beginning of the eighteenth century, the German


chemist Georg Stahl, armed with this new knowledge that
everyday substances could be made from hidden elements,
decided to put forward an explanation for fire.
When metals burn they form colored powders, which were
called calxes at the time. Calxes were notoriously difficult to
set alight, so Stahl concluded that they were elements, difficult
to ignite because their fire had been removed.
According to this hypothesis, anything flammable
contained a substance that escaped into air when heated,
leaving behind the charred remains. This substance was named
phlogiston from the Greek phlogizein (to set alight) and Stahl
argued that a fire was phlogiston being separated from a calx.8
Stahl’s fire hypothesis was important because, unlike
previous ideas in chemistry, it was testable. If correct, it should
be possible to trap phlogiston and combine it with a calx to
regenerate the original metal. By putting forward an idea that
could be proven wrong, Stahl gave us a genuine scientific
hypothesis and, like most scientific hypotheses, it was quickly
destroyed.
The first chink in the armor came from the French-British
scientist Henry Cavendish. He was a notoriously shy man with
a penchant for collecting furniture, beloved by physicists
because he helped provide evidence for the force of
gravitation. His greatest contribution to chemistry though was
a series of experiments involving acid and iron.
The reaction between these two always released an
invisible gas, which Cavendish collected. His first thought was
that he had successfully got hold of phlogiston until he
discovered something odd. The gas was explosive.9 If fire was
the result of phlogiston escaping, how could phlogiston itself
be burned? How could phlogiston escape from itself?
Stranger still, when Cavendish’s gas (which he called
flammable air) exploded, it generated pure water. If you could
make water from other things, maybe water wasn’t elemental
either.
The next mystery came in 1774 from the heretical English
clergyman Joseph Priestley. Priestley was experimenting on
calx of mercury (the red powdery substance you get when
mercury is burned) and directing beams of sunlight at it with a
magnifying glass.10
He collected the gas given off and found that other things
burned very well inside it, better than they did in normal air.
Whatever it was, it was clearly good at removing phlogiston.
Logically this gas had to be dephlogisticated because it was
able to absorb phlogiston, so he called it “dephlogisticated
air.”
About two hundred years previously, the Polish magician
Michał Sędziwój had discovered air to be a mixture of two
gases, one of which was “the food of life” and one of which
was useless.11 Could this be related?
Priestley decided to seal some mice in a box with his
dephlogisticated gas and they survived without harm. He also
discovered, after testing it on himself, that it was actually
preferable to regular air and made him feel euphoric to breathe
it. Sędziwój’s food-of-life gas was apparently the same as his
dephlogisticated gas.
Priestley also discovered that plants seemed to breathe the
gas out, replenishing a room after a fire had burned. The whole
thing was very confusing. Fires generating water, metals
generating fire, plants generating air … What was going on?

BRINGING ORDER

The answer to all the riddles came in 1775 when Priestley


shared his phlogiston results with the French chemist Antoine
Lavoisier.
Lavoisier worked for the French government collecting tax
contributions but his real passion was science. He had already
been experimenting on calxes by the time Priestley’s
experiments came to his attention12 and decided it was time to
put the phlogiston hypothesis through its paces. If fire was the
result of phlogiston leaving a substance, the leftover calx
should weigh less.
Priestley had tried taking measurements with his
magnifying glass and mercury calx, but precision equipment
didn’t exist in the eighteenth century. Imagine trying to
distinguish a powder weighing 1 g from a powder weighing
1.1 g. Quite the challenge.
Lavoisier decided to scale up Priestley’s experiment in
order to get a clear result. The difference between 1000 kg and
1100 kg is a difference of 100 kg, which you could see with
the naked eye. So, Lavoisier ordered the construction of a
nine-foot magnifying glass and blasted a plateful of mercury
calx with sunlight.13
The results were unmistakable—calxes weighed more than
the original metal. Everyone had it backward. Fire wasn’t the
removal of phlogiston: it was something being added from the
air itself. Substances like metals and phosphorus were the
elements and fire was what happened when they combined
with Priestley’s gas.
As brilliant as this insight was, Lavoisier wasn’t perfect
and mistakenly thought Priestley’s gas was also responsible
for the sour taste of acids. He called it oxygène from the Greek
oxys-genes (sour-maker), which translates into English as
oxygen.
The exploding gas Henry Cavendish had isolated was a
different element (contained within the acid, not the metal)
and, when heated with oxygen, combined to form water.
Lavoisier named this gas hydrogène from the Greek hydros-
genes (water-maker), which translates into hydrogen.14
This new way of looking at things also explained why you
couldn’t breathe in a room after a fire had been burning. It
wasn’t because the fire was giving out a toxic substance: it
was because air was partly made from oxygen and fires
absorbed it, leaving the other gas behind.
This useless gas was eventually shown to react under
extreme conditions and could make niter, one of the key
ingredients in gunpowder, so the French statesman Jean-
Antoine Chaptal named it nitregène—nitrogen.
Science always progresses when a hypothesis is proven
wrong and Lavoisier’s experiments signed the death warrant
on phlogiston. Air was an unreacted mixture of nitrogen and
oxygen, water was a fused compound of hydrogen with
oxygen, and fire was a reaction between oxygen and any
available chemical. None of them was an element.
For his efforts, Lavoisier was taken to the guillotine in May
1794. Possibly because he worked as a taxman in pre-
revolutionary France (never a good idea), but more likely
because he criticized the inferior science of Jean-Paul Marat,
who became a leading figure of the revolution. An unlucky
end for a great mind, although that’s nothing compared to the
bad luck of a chemist named Carl Scheele.

THE UNLUCKIEST MAN IN THE HISTORY OF CHEMISTRY

Cavendish, Lavoisier, and Priestley were geniuses of a new


science and other people quickly joined the hunt. Everyone
wanted the glory of discovering a new element, although
agreeing on who makes a discovery isn’t always obvious.
Some elements have been around since antiquity so it’s
impossible to know who originally discovered them. The Old
Testament contains passages dating back three thousand years
that refer to gold, silver, iron, copper, lead, tin, sulfur
(correctly spelled with an f—see Appendix I), and possibly
antimony.15
Then there are instances of someone predicting an element
without actually obtaining a sample. Johan Arfwedson
deduced there was an element hidden within petalite rock and
named it lithium from the Greek lithos (rock), but it wasn’t
until 1821 that William Brande extracted it.16
In order to avoid confusion and settle debates we tend to
talk about the first person to isolate an element rather than
discover it. Credit goes to the first person who manages to
hold a pure sample of an element and recognize it as such.
Which brings us to the Swedish chemist Carl Scheele.
In 1772, Scheele successfully made a brown powder, which
he named baryte from the Greek barys, meaning heavy. He
knew there was an element hidden inside (barium) but it was
Humphry Davy who isolated it and got the glory.
In 1774, Scheele discovered the gas chlorine (from the
Greek chloros, meaning green) but didn’t realize it was an
element. It was again Humphry Davy who made this link in
1808, thus getting the credit.
That same year, Scheele discovered calx of pyrolusite but
failed to isolate the elemental manganese inside, achieved a
few months later by Johan Gahn.
Then it happened again in 1778 when Scheele identified
molybdenum, before it was isolated by Peter Hjelm. And then
again in 1781 when he deduced the existence of tungsten but
failed to isolate it before Fausto Elhuyar, who got the credit.17
Scheele even discovered oxygen in 1771—three years
before Priestley—but his manuscript was delayed at the
printers and, by the time it was published, Priestley had got his
results out.18
To commemorate his many contributions to chemistry, the
mineral Scheelite was named after him … until it was
officially renamed calcium tungstate and Scheele was once
again nudged out of the history books. If there is a god of
chemistry, he apparently hates Carl Scheele.
CHAPTER TWO

Uncuttable
DIAMONDS, PEANUTS, AND CORPSES

In 1812 the German chemist Friedrich Mohs invented a 1 to


10 scale to classify the hardness of minerals. Tooth enamel has
a score of 5, for example, while iron ranks as a 4. This means
your teeth will technically dent a lump of iron but not the other
way around. Although I don’t recommend you try it because if
you accidentally bite steel (iron with carbon impurity), which
has a hardness of around 7.5, you’ll regret it.
Diamonds were given a value of 10 because they were the
hardest things known at the time. Their claim to the crown was
only overthrown in 2003 when a group of researchers from
Japan managed to make something even harder—a
hyperdiamond.
The most common explanation given for how diamonds
form is that coal (fossilized plant) gets compressed
underground until it turns hard and transparent. It’s what
everyone gets told in primary school but it’s a complete myth.
Diamonds are made in a much more extreme environment.
The same year hyperdiamonds were manufactured,
Hollywood birthed its own unbelievable creation: The Core, a
sci-fi film, which has to be seen to be believed. A few
highlights from the movie involve a man hacking the entire
global internet from a laptop, sunlight melting the Golden
Gate Bridge, and Hilary Swank landing a space shuttle in the
San Fernando Valley.
One scene in particular stands out for me. A team of
scientists is launched into the Earth’s mantle in order to nuke
the Earth’s core and find themselves dodging diamonds the
size of buildings.1
What’s interesting about this scene is that, while giant
diamonds are unlikely, it’s otherwise fairly accurate.
Diamonds really are made in the Earth’s mantle, not in the
crust.
A diamond is made solely from carbon and it takes billions
of years to grow one. Plants do contain carbon but haven’t
been around long enough to create the gems we extract from
mines today. To fuse carbon into a crystal also takes a
staggering amount of pressure and temperature—far more than
you could achieve in a planetary crust.
Diamonds are really made a few hundred kilometers into
the upper mantle, where pressures are hundreds of thousands
times greater than atmospheric pressure and temperatures are
comparable to the surface of the Sun. Once they’ve been
made, the crystals are vomited to the surface in volcanic
eruptions, which solidify, and we eventually dig them up.
The compressed-plant myth probably arises because we
also mine coal and that is made from heat-compressed plant,
but it forms at wussy temperatures and pressures, inadequate
for diamonds.
It is also true that one naturally turns into the other, but it’s
the opposite of what the myth claims. Diamonds are slightly
unstable and will decay into coal over thousands of years. So,
the obvious question is: could we reverse the process?
In 2003, Tetsuo Irifune from the Tokyo Institute of
Technology decided to try compressing coal into a diamond
for real. By using the engineer’s equivalent of an extreme
pressure cooker, Irifune took a lump of coal-like carbon and
subjected it to pressures far in excess of what you’d get in the
mantle. The result was a hyperdiamond, a chemical never seen
before in nature.2
Hyperdiamonds will have a Mohs value greater than 10 but
the precise number hasn’t been calculated because the original
piece of carbon is compressed so much the resulting
hyperdiamond is tiny. We’re talking a few millionths of a
gram.
But we don’t have to use coal as our starting material. Dan
Frost from the Bavarian Geological Institute in Germany
managed to make a diamond by compressing peanut butter,3
and the Illinois-based company LifeGem can make artificial
diamonds by compressing your deceased loved one’s ashes.
Provided you’ve got the carbon, it can be crystalized.
The fact that coal, diamond, and hyperdiamond are all
made from the same element yet have different properties (we
refer to them as “allotropes of carbon”) suggests that elements
can somehow arrange themselves in different ways.
In order to explain this phenomenon, we’re going to have
to look closely at the notion of something being diamond-like
or “uncuttable.” And in ancient Greek the word for uncuttable
is one you probably know already: atom.

THE MAN WHO PROVED GOD

Imagine holding a grain of sand between your fingertips. It’s


hard to make out details with the naked eye but logically the
grain would have two halves; a left hemisphere and a right
one. You could imagine a knife small enough to chop the grain
right down the middle, splitting it in two. Then, once you had
these half grains, you could repeat the process, slicing to
quarter grains and so on.
Theoretically, you could do this forever. No matter how
small the grain fragment, you’d always be able to zoom in and
divide in half again.
The alternative would make no sense. Imagine chopping a
grain up so small that it no longer had a left or right half. A
piece so small it didn’t have any size and just was. For an
object like this, the very concept of dividing by two would be
meaningless. It would be like trying to divide by two on a
calculator and the calculator replying with “Sorry, you have
reached the smallest thing, you can’t divide anymore.” You’d
have to be crazy to suggest the existence of a smallest object.
Cue Democritus.
Democritus was a philosopher/stand-up comedian living in
the fifth century BCE and he took the idea of elemental
substances very seriously. He believed everything was made of
microscopic uncuttable pieces (atoms) that combined to make
the world around us.
Say you’ve got a packet of M&M’s. Rather than eating
them in mixed handfuls, every sane human being divides them
into piles organized by color and eats them one pile at a time.
Don’t trust anybody who does otherwise.
This sifting of a mixture into purity is what we’re really
doing when we break a substance down into its elements;
we’re grouping the atoms according to type. This would also
explain where allotropes come from. Diamond, coal, and
hyperdiamond could all be made from carbon atoms stacked
and arranged differently, leading to a variety of properties.
And, as if the atom hypothesis wasn’t strange enough,
Aristotle later used Democritus’s idea to prove the existence of
God. Because atoms were constantly in motion, bouncing off
each other and flying through the emptiness between, every
atom’s movement could be back-tracked to a collision with an
earlier atom, whose movement could be explained as a
collision with an earlier one still. Cause led to effect and every
effect had a preceding cause.
If you went back far enough there must have been a first
movement that caused everything but had no cause itself. Such
a thing (an uncaused cause) would be outside the normal laws
of nature while still being able to influence them. God, in other
words.4 Make of that what you will.

LORD OF THE SWAMP

Sadly, along with many other great ideas, Democritus’s atomic


hypothesis was shelved as the Holy Roman Empire took hold
of intellectual Europe. It wasn’t until the late 1700s that atoms
were given serious attention thanks to the work of an English
scientist named John Dalton.
At the age of twelve, most people in England are getting
acquainted with being a student in high school. John Dalton
was teaching at one. The son of a weaver, Dalton had already
taught himself science, mathematics, English, Latin, Greek,
and French, and achieved the rank of headmaster by his late
teens.5
Don’t be fooled though. While a fierce academic, Dalton
still knew how to have a good time and, like any youngster,
spent his free moments collecting samples of swamp gas from
local bogs. Surprisingly, he never married.
It was while burning these samples of gas that Dalton
learned gases don’t react all willy-nilly but combine in specific
ratios. Hydrogen and oxygen, for instance, always combine in
a two-to-one mix and nothing else. If you have three times as
much hydrogen as oxygen, you end up with a third of your
hydrogen left at the end. It’s as if there’s only a limited amount
of oxygen “bits” to go around.
Dalton decided the best way of explaining these findings
was to assume there were tiny particles making up each
elemental gas. Thanks to his proficiency in Greek he was
familiar with the work of Democritus and began referring to
these particles as atoms.
The idea, however, was not widely accepted. Dalton had a
habit of overcomplicating things and the book he published in
1808 to outline his atomic hypothesis was a notoriously
difficult read.6 His ideas were rigorous but his explanations
were boring and his chemistry was cumbersome.
Nevertheless, Dalton was greatly respected and was
eventually given the privilege of being presented to King
William IV. This also led to him committing the biggest faux
pas of his career because Dalton was a Quaker and forbidden
to wear scarlet clothing, which happened to be the color of
robes required for meeting the king. Dalton was color-blind
(incidentally, he was the first person to document its
existence), and the event’s organizers “forgot” to tell him he
was wearing robes that would offend his fellow Quakers.7
So Dalton went parading around in front of other Quakers
in the most outrageous clothing imaginable. The unluckiness
of being simultaneously color-blind, a Quaker, and publicly
dressed in scarlet is remarkably unfortunate. Somewhere, in a
dark corner of purgatory, Carl Scheele is probably cackling to
himself.

UNDER PRESSURE

The real watershed for the atomic hypothesis came in 1899


when the French physicist Émile Amagat began experimenting
with pressure chambers. Amagat had spent his youth lowering
samples of gas into mineshafts to measure how much they got
compressed, and by adulthood had designed sophisticated
machinery capable of compressing gases to three thousand
times atmospheric conditions.
Through these experiments he discovered there was a limit
to how far a gas could be squeezed. Once you got to a certain
point, the gas fought back and refused to get smaller.8
This couldn’t be explained with the infinitely-smaller-
particles hypothesis. If matter was made from infinitely small
chunks, then any gas would contain an infinite number of gaps
in between them as well. No matter how small you
compressed a gas there would always be enough space for the
matter to fall into.
The physicist Robert Boyle, son of the Earl of Cork, had
conducted experiments on gas pressure and argued that it was
possible to compress a gas forever because of this very reason.
Amagat’s research showed otherwise. A gas had a fixed
amount of matter, which meant it probably wasn’t made from
an infinity of smaller bits.
Combined with Dalton’s swamp-gas discoveries, Amagat
made the idea of atoms look less like a hypothesis and more
like a theory—something that has evidence in its favor.
However, there was one big problem or, rather, a very small
one. In order to make sense of Amagat’s readings you had to
accept that atoms were tiny. Unthinkably so.
Imagine looking at planet Earth from space and trying to
pick out a single grape on its surface. That is the equivalent of
looking at a grape and trying to pick out a single atom on its
skin.
If atoms were real they would need to be so small that even
waves of visible light would be too big to bounce off them. It
wouldn’t matter how powerful your microscope was, atoms
would be impossible to discern by their very nature.
Scientists are in the business of testing theories once
they’ve been established, but how could you test this one?
How could you see the unseeable?
EINSTEIN WAS HERE

Albert Einstein was a legend in his own lifetime. What’s more


impressive is that he deserved the reputation. Publishing over
three hundred scientific papers and essentially inventing the
landscape of modern physics, Einstein was the epitome of
genius.
It would be foolish to summarize his many achievements in
a few paragraphs, so we’ll focus on the one most relevant to
chemistry: a paper he published on July 18, 1905, in which he
made the atomic hypothesis testable rather than speculative.
While working at the Swiss patent office, Einstein
stumbled across some research from 1827 by the Scottish
botanist Robert Brown. Brown had noticed that grains of
pollen floating on water appeared to jiggle in random patterns.
Originally, he had assumed the grains were alive but found the
same thing happened with sand or dust. The phenomenon was
known as Brownian motion and, although unexplained, it was
nothing more than a curiosity.
Einstein decided to model the pollen’s trajectory through
the water and found it could only be explained as the result of
bombardment from water particles. To accurately describe
how the pollen moved, you had to factor in the friction of
pollen against water, which meant you had to accept the
existence of “water atoms.”
Despite the persistent rumors that he failed math in school,
Albert Einstein was a mathematician par excellence and drew
up an equation that related water temperature to the pollen
grain’s likely movement. By introducing an equation with a
measurable outcome, Einstein changed the game completely.
An idea can be debated but a number cannot, so if you can
predict a specific value from your hypothesis you have
something to search for directly.
He finished his paper with the phrase, “It is to be hoped
that some enquirer may succeed shortly in solving the problem
suggested here.”9 As was usually the case with Einstein, his
equation was soon tested and confirmed. The zigzagging
wasn’t random at all, but the result of minor fluctuations in
water movement on either side of the grain. Pollen looked like
it was undergoing constant collisions because it genuinely
was.
In finding this, Einstein did for the atomic hypothesis what
Lavoisier did for the elemental one: he provided indisputable,
quantitative evidence. You couldn’t sensibly discuss elements
without atoms anymore, or vice versa. There was no argument
to be had. Atoms were real.
CHAPTER THREE

The Machine Gun and the Pudding


THE SMALLEST MOVIE IN HISTORY

In 1989 researchers at IBM pushed the boundaries of


marketing by creating a sculpture of their company logo using
only thirty-five atoms. Then in 2013 they went even further
and created a sixty-second film, A Boy and His Atom, by
drawing images with atoms and animating them through stop-
motion, earning a Guinness World Record for the world’s
smallest stop-motion film.
By taking hundreds of photographs in different positions
and playing them at high speed, the IBM researchers were able
to tell the story of an atomic stick-figure who plays with a pet
atom. This wasn’t an easy thing to do, because as we saw in
the previous chapter atoms are too small to be seen.
The trick to their film is that each photograph of the atomic
models is not really a photograph. They are images obtained
from a scanning tunneling microscope (STM), a device that
allows us to peer at distances smaller than visible light can
access.
Imagine standing beside a dark hole and dropping a rock
over the edge. By timing how long it took to reach the bottom,
you could calculate how deep the hole was without being able
to see it. STMs work on a similar principle.
The business end of an STM is not a lens but a thin nozzle
with tiny particles clinging to the tip. These particles are
bound loosely so when you apply an electric current they fall
off and land on the surface beneath. As they fall they lose a
certain amount of energy, which the STM can measure,
calculating how far away the surface is.
By scanning the tip back and forth across an object, any
bumps and blips will correspond to a different amount of
energy being lost, and the STM can indirectly create a map of
what the object must look like.
The filming of A Boy and His Atom was carried out by
creating a flat sheet of copper and bonding carbon monoxide
particles to it in specific positions. As the microscope scanned
across the copper, it picked up these carbon monoxides like
dots on a Braille picture and created the corresponding image
in the computer.1
It’s a novel idea but how can it be possible? In order to
detect the outline of an atom, our STM would need to be
dropping particles even smaller than atoms. Where can we
find particles that small?

CALL ME “J. J.”

At the turn of the twentieth century, the main pursuit of any


serious physicist was trying to understand electricity. There
were two leading nineteenth-century hypotheses under
consideration, each supported by some of the biggest names in
science. In one corner was the legendary Hermann von
Helmholtz, a staunch believer in particles. He argued that
since arcs of electricity cast shadows, for example, it had to be
made from matter—electrical atoms.
Leading the opposition was his student, Heinrich Hertz,
who preferred to explain things with invisible force fields.
Having recently shown that magnetic fields could be used to
bend the path of electric current, Hertz argued that electricity
also had to be a disturbance in some sort of electric field.2
Their disagreement was passionate, although Helmholtz
and Hertz remained good friends until the end. Sadly they both
died in 1894, shortly before the matter was finally settled by a
brilliant British physicist named Joseph John “J. J.” Thomson.
It was Helmholtz who had been right.
J. J. Thomson was, by anyone’s account, a wunderkind of
science. He was admitted to the University of Manchester at
the age of fourteen and was later appointed to the most
prestigious physics post in Britain, taking over from Lord
Raleigh as the Cavendish Professor of Experimental Physics at
Cambridge University.
The precise details of Thomson’s electricity experiments
are very mathematical but the premise is straightforward. Fill a
small chamber with gas and connect two ends of a circuit to
the front and back. At a high voltage it is possible to generate
streams of electricity through the gas and, if you place
magnets at certain points, you can manipulate their behavior.
By carrying out a variety of studies on this theme,
Thomson made several crucial observations. Most important
was the fact that electricity moved slowly. Hertz’s field
hypothesis predicted electricity should move at the speed of
light but Thomson’s measurements clocked it as practically
sluggish by comparison. This meant electricity had mass and
was therefore made from particles.
The Irish scientist George Stoney called these particles
electrons, from the Greek electron, meaning amber (which
could be rubbed to create shocks of static electricity), and it
caught on. Except electrons were notably different from other
particles.
For one thing, the atoms discovered by Dalton and Einstein
were two thousand times bigger. In fact, it was possible to
shoot a stream of electrons through a plate of solid iron
because they could apparently fit through the gaps.
Normal atoms are also happy to approach each other
whereas electrons actively repel. This repulsive property was
named charge and, in all honesty, it’s still a mystery. We can
measure its influence and describe the mechanism that causes
one electron to repel another, but why electrons have charge is
not yet understood.
More pressing for Thomson was the issue of where
electrons were coming from. Batteries are composed of regular
atoms (bigger and chunkier) so electrons had to be somehow
hidden within them. Apparently, atoms weren’t the smallest
things after all—they contained electrons.
So how come atoms didn’t have this property of charge? If
the electrons within them were repelling, how were two atoms
able to approach each other and even bond?
Thomson concluded that atoms had to contain some
additional substance with an anti-charge, canceling the
electron charge and giving atoms the appearance of being
neutral overall.
Thomson proposed that electrons were nestled within a
kind of atomic sponge. Slice away a segment of an atom and
you’d see the electrons arranged like plums in a traditional
British Christmas pudding. A bit like this:

The electrons and dough had opposite and attractive


charges, which was why it took so much effort to pull
electricity from an atom—you had to rip electrons away from
their complementary dough.
Thomson’s model of the atom was given the rather catchy
name of “The Plum-Pudding Hypothesis.”

THE IMPORTANCE OF BEING ERNEST


The name atom had stuck by the time Thomson published his
work, which is a shame as it’s notoriously misleading. What
we call atoms are not uncuttable at all, and neither are they the
smallest things. They’re just stable structures that prefer not to
be pulled apart.
Electrons are the truly uncuttable particles and, as far as
Thomson could figure, they were suspended in an oppositely
charged dough. But science makes progress by disproving a
hypothesis, not by proving it, and the plum-pudding idea was
eventually torn to pieces by Thomson’s student Ernest
Rutherford.
Raised on a New Zealand sheep farm, Rutherford was
known for rejecting expensive equipment and carrying out
ludicrous experiments because nobody else was doing them.
His unorthodox approach earned him the 1908 Nobel Prize in
Chemistry, though, so people tended to let him get on with it.
He won the prize for discovering that larger atoms could
spit out tiny pieces, which he called alpha particles, that are
much heavier than electrons and carry the opposite charge.
Rutherford assumed this happened because atomic dough
repelled itself and when the atoms were large there was more
chance of self-repulsive instability, leading ultimately to an
explosion. The alpha particles he discovered were believed to
be bits of atomic dough spat out by the micro-blasts.3
Most people would have accepted the Nobel Prize and
moved on, but Rutherford was a scientist to the core. He
wanted to put his own hypothesis on the chopping block and
see if he could disprove it. So he hired the world’s best
experimentalist, Hans Geiger, and together they developed a
method for probing the interior of an atom.
They discovered that alpha particles would produce tiny
flashes of light when they hit a piece of zinc sulfide (ZnS), so
they spent countless hours sitting in dark rooms firing alpha
particles at ZnS, looking for flashes through a lens.
The boredom was unbearable, so Geiger invented an
electronic counter that would detect the impacts automatically.
His invention was the crackling Geiger counter used in a
thousand spy movies ever since.
One morning in 1909, Geiger went to see Rutherford to
talk about one of their promising undergraduates, Ernest
Marsden. Marsden was only twenty but was gaining a
reputation for exceptional lab prowess.
Geiger wanted to give him a new project so Rutherford, in
his typical oddball style, came up with something peculiar:
“Why not let him see if any alpha particles can be scattered
through a large angle in the gold foil experiments?”4
The gold-foil experiments had been designed a few years
earlier. By taking a piece of radium (a highly alpha-spitting
metal) and pointing it at a thin foil you could shoot alpha
particles right through the foil. Placing a detector on the other
side would let you measure how much the particles were
affected by the foil and gave clues about the density of the
atomic dough. The best metal was gold because it could be
stretched into a leaf only a few atoms thick. The setup looked
like this:

For some reason, Rutherford wanted Geiger and Marsden


to put the detector at huge angles to the foil rather than directly
on the other side. Geiger must have been puzzled since surely
the detector would read nothing but, given Rutherford’s
reputation (the Nobel Prize medal on his desk probably helped
too), he just shrugged and set Marsden to work. The very next
day, Rutherford’s eccentricity paid off.
The detector began picking up scattered alpha particles
even when the detector was moved to the same side as the
alpha source. This couldn’t be explained with the plum-
pudding hypothesis because how could an alpha particle get
bounced back by dough? It would be like setting up a machine
gun to fire at an actual plum pudding and have the bullets
bounce back and shoot you in the face. You would expect
them to cut right through the pudding and hit the opposing
wall, so why are you now in hospital? And what explanation
are you going to give the admissions nurse?

Rutherford described it in similar terms: “It was quite the


most incredible event that has ever happened to me in my life.
It was almost as incredible as if you fired a fifteen-inch shell at
a piece of tissue paper and it came back and hit you.”5
The results were published in February 1910 and, by the
following year, Rutherford had run the math. There was only
one possible explanation for the result. The atom wasn’t a soft
sponge all the way through but had hard lumps for the bullets
to bounce off. The plum pudding apparently contained nuts.
These nuts were most likely small and clustered in one
place within the atom, since only a few bounces were detected
for every thousand bullets. It would also make sense for them
to have the same charge as alpha particles in order to scatter
them when impacted, not to mention holding the electrons in
place.
Rutherford called this clump of particles the nucleus, from
the Latin for nut, and proposed that electrons orbited it like
planets around the Sun. Thomson’s plum-pudding idea had to
be abandoned. It was ingenious, but it had no evidence and in
science no evidence means no theory.

YOU WANNA GET NUTS? LET’S GET NUTS!

Did Rutherford have a hunch about the nucleus or was he just


fooling around when he suggested moving the detector? Was
he trying to think of some task for Marsden and that was the
only thing he could come up with at short notice?
Personally, I like to imagine Marsden putting the detector
on the wrong side as a sulky middle finger to Rutherford. Here
was this great man giving him a stupid task to perform. Oh,
you want wide angles? How does the wrong side of the foil
sound to you? That wide enough for you, Rutherford?
We’ll probably never know but whatever happened in that
lab, and whatever went through the minds of the three men,
the results have become a part of science lore.
There was still a niggling question that needed answering,
though. Rutherford’s idea was that the nucleus contained
particles with a charge opposite to an electron, but if this was
so, why wasn’t it ripping itself apart? Particles with the same
charge repel each other so the nucleus shouldn’t exist at all.
The answer was discovered by another of Rutherford’s
students, James Chadwick, in 1932.
Using a piece of polonium, known to eject alpha particles,
Chadwick bombarded a lump of beryllium metal and set up a
piece of wax on the other side to cushion any impacts.
Every time there was an emission from the polonium,
something inside the beryllium came flying out the other side
as if a pool-ball collision was taking place within the nuclei.
These ejected particles were obviously heavy but they didn’t
repel each other, meaning they had to be neutrally charged.
They also had to have some glue-like property that held
charged particles together more powerfully than they could
repel themselves.
The nut of the atom was apparently made of two types of
particle. Neutrons (the neutral ones), which had the glue
property, and charged protons (from the Greek word for first),
which held the electrons in place. Further research from Niels
Bohr, Werner Heisenberg, and Oskar Klein elaborated on
Rutherford’s findings, and the popular view of the atom was
eventually established.
Atoms were like solar systems. Protons and neutrons
formed the central nucleus with oppositely charged electrons
whizzing around the edge with apparently nothing in between.
If you imagine expanding an atom to the size of a football
stadium, an electron would become the size of a dust mote
while protons and neutrons would be huddled together in a
nucleus, roughly the size of a golf-ball hovering in the center.
The strangest conclusion from this is that most of an atom
is empty space. Even something like osmium, the densest
element, is apparently 99 percent nothing. As are you.
HIDDEN ELEMENTS

In the Superman movie Man of Steel, the spaceship that brings


Kal-El to Earth is analyzed by chemists and found to be made
of elements that don’t fit on the periodic table.6 The periodic
table is a list of all known elements so the Kryptonians
obviously have different elements on their planet to ours.
The idea of hidden elements is a tantalizing one and it’s
been thrown around in fiction for decades. In the H. P.
Lovecraft story “The Dreams in the Witch House,” the
protagonist discovers a small statue made of an element that
cannot be identified by any scientist.7 Lovecraft was inspired
by a physics lecture he attended in the same year neutrons
were discovered, but could such things really exist? Could
there be exotic elements tucked away in unknown corners of
the Universe?
Not that I want to destroy these fictional stories, but the
answer is no. You can’t have hidden elements for a
straightforward reason: atoms aren’t atomic. Yeah, I know,
right: what the hell?
The word atom was obviously supposed to mean uncuttable
but it’s really the electrons, protons, and neutrons that fit this
description. The word atom had stuck, however, so we still
call them that even though “electron-proton-neutron-
superstructures” would be a more accurate term.
The smallest possible atom would logically contain one
proton (and one electron since the charges always cancel).
This would be element number 1, which turned out to be
Cavendish’s exploding gas—hydrogen.
The next element would have two protons (plus some
neutrons to glue them together). That turned out to be helium.
It wouldn’t be possible to have element 1.5 in between
because there is no such thing as half a proton (see Appendix
II).
Once you’ve got a list of all the elements, you can be sure
you haven’t missed anything because nature is only able to
make atoms in whole numbers. The elements you find on
Earth are the same elements you find everywhere in the
Universe. Which is where we’re going in the next chapter.
So I’m sorry, Superman, your spaceship isn’t possible.
Interestingly, though, kryptonite is real. The chemical formula
for kryptonite is LiNaSiB3O7(OH)F2, a mineral which was
discovered at a Serbian mine in 2007.8
CHAPTER FOUR

Where Do Atoms Come From?


THE COLDEST PLACE IN THE UNIVERSE

The temperature scale we use for our everyday lives was


invented in 1742 by Anders Celsius. He took the freezing and
boiling temperatures of fresh water, divided the scale into a
hundred chunks, and called them “centigrades” from the Latin
for hundred steps.
Celsius’s original thermometer defined 100°C as freezing
and 0°C as boiling, but this was reversed after his death and
the scale was renamed the Celsius scale in his honor. The
Fahrenheit scale, used more widely in the United States, was
invented by Daniel Fahrenheit, who used salted ice and
created a scale going up to human body temperature.
Whichever scale you’re using, the behavior of particles is
the same: as you heat up something, the average speed of its
particles increases. Because higher temperatures cause
particles to fly around more, this also tells us that when gases
get hotter they take up more space. Conversely, if particles
become colder they occupy a smaller volume because they
move around less. Hotter gas = bigger. Colder gas = smaller.
This simple relationship between temperature and volume
is called Charles’s law after Jacques Charles, the physicist who
discovered it. But obviously this relationship can’t go on
forever. If you keep cooling things down further and further
then the volume shrinks with it, so eventually you should
reach a temperature where the volume drops to zero.
Charles’s law implies that there’s a temperature so cold the
particles would take up no space, effectively winking
themselves out of existence. This hypothetical temperature is
clearly impossible so we call it “absolute zero,” calculated to
be –273.15°C. It’s a temperature so cold you would have to
break the laws of physics to reach it.
The coldest place on Earth is usually reported as a point in
Antarctica near Dome Argus, which drops to –93.2°C during
the winter season.1 The emptiness of deep space has an
average temperature of –270°C while the Boomerang Nebula
stoops to –272°C, one degree above what’s physically
possible.2
But the all-time record for coldest place in the Universe is
right here on Earth, at the lab of Martin Zwierlein in
Massachusetts, where his team have been able to synthesize a
chemical called sodium-potassium, the coldest chemical ever
created.
Usually when two atoms bond (see Chapter 8), we attach
the suffix ide to whichever element isn’t a metal, e.g. iron
oxide. A bond between two metal atoms is so rare, however,
that we haven’t invented a naming system for it, hence the
rather unusual-sounding sodium-potassium.
Zwierlein’s experiment works by filling a chamber with
sodium and potassium atoms in the gas state and heating them
to around 7,300°C. By applying a magnetic field across the
chamber, the atoms lose their ability to move in as many
directions and they begin pairing up (a phenomenon known as
Feshbach resonance).
The next step is to zap the gas with two laser beams, one at
high energy and one at low energy. When blasted with the
high-energy laser, the atoms become stimulated and begin
glowing the same color as the beam. Giving out their own
light causes them to lose energy, of course, so this is where the
second laser comes into play.
Because it is emitting at lower frequency, it serves as a sort
of landing platform for the atoms to drop toward. The atoms
keep losing energy until they match the frequency of the lower
laser, leading to a colossal drop in temperature.
Zwierlein managed to strip the molecules of their heat and
plummeted the temperature to five-hundred billionths of a
degree above absolute zero, the current world record.3 But
since studying materials at their coldest temperature tells us a
lot about how particles behave, we want to go further.
The problem with Zwierlein’s experiment is that he
performed it on Earth and the gravitational field of our planet
pulls slightly on the atoms, causing them to wiggle, thus
raising the temperature. The obvious solution is therefore to
remove the effects of gravity.
That’s the aim of the Cold Atom Laboratory, a version of
Zwierlein’s experiment due to be performed aboard the
International Space Station (ISS). Because the ISS is orbiting
the Earth and changing direction constantly, the effects of
gravity average to zero. It might be possible to drop atoms not
only to billionths of a degree, but to trillionths.
The rules of astrochemistry are clearly very different to
those of Earth chemistry and space is where we need to look
next in order to understand where elements come from in the
first place.

FOR WHAT DO WE KNOW OF THE STARS?

Many centuries ago in the province of Miletus, the great


philosopher Thales was ambling through a dark field staring
up at the speckled lights that swam across the sky. There were
no street lamps in the sixth century BCE so Thales had a perfect
view of the Universe with stars uncountable stretching from
horizon to horizon.
It was in this moment, as he began to wonder what the stars
themselves were made of, that he took a step forward, found
nothing but air and toppled into a pit. As he crumpled to the
bottom, a Thracian servant girl came dashing up to the edge
and giggled hysterically after him, “Maybe you should look to
the ground, old man, and not only to the stars!”4
I know exactly how he must have felt. I once put my pants
on backward while trying to solve an equation in my head. I
even did the zip without noticing and only discovered the
mistake hours later when I tried to put my hand in my pocket.
Centuries after Thales, the philosopher Aristotle decided
that stars were made of an unreachable substance called ether
—the holy element of the gods.5 A nice hypothesis but a
completely untestable one since, by definition, gods are
beyond the human realm.
If you followed Aristotle’s logic, there were unobtainable
materials in the Universe and therefore no point in trying to
understand what everything was made of.
Unfortunately, his idea caught on and people stopped
searching for answers through experimentation and relied on
guesswork alone. This trend of trusting opinion over data is
why scientific progress died for a millennium and we got stuck
in the dark ages. So, nice going, Aristotle.

TWINKLE, TWINKLE

The stranglehold of Aristotle finally began to loosen in 1814


when the German physicist Joseph von Fraunhofer made an
important discovery. When you look at a beam of light from a
flame, you can split it with a prism and reveal a multitude of
colors. It’s the same effect that causes rainbows.
What Fraunhofer found was that not every beam of light
looks the same when you split it. Different types of fire
produce different types of rainbow.
Forty-five years later, Robert Bunsen (of the burner)
realized the implications of this discovery. Each element gives
off a particular spectrum when burned, like a unique rainbow
fingerprint. By studying the light from a fire using
Fraunhofer’s equipment, you could calculate exactly what
atoms are present in the reaction.
This technique, called spectroscopy, allows us to monitor a
reaction from a great distance, so if we turn our spectrometers
to the stars we should be able to deduce their composition.
The most interesting spectroscopic finding came in 1868
when the French astronomer Pierre Janssen and the British
astronomer Norman Lockyer simultaneously observed a
completely new elemental signature in the light of our own
Sun.6 It didn’t match any of the known elements on Earth so
Lockyer named it helium from the Greek helios, meaning Sun.
Twenty-seven years later, William Ramsay extracted it from
terrestrial rocks, making it the only element to be discovered
in space before it was isolated on Earth.7
The next breakthrough happened in 1925 when the
American astronomer Cecilia Payne-Gaposchkin successfully
calculated how much of each element was present in a typical
star.
Payne-Gaposchkin studied astrophysics at Harvard under
Harlow Shapley, one of the only astronomers in the world to
let women take the subject, and wrote her PhD thesis on the
star classes identified by another astronomer, Annie Jump
Cannon (possibly the greatest name in science).
Cannon was completing her nine-volume catalog of every
known star when Payne-Gaposchkin began perusing the data.
Being well versed in the new science of quantum mechanics
(which most astronomers weren’t), Payne-Gaposchkin showed
that the amounts of each element in stars were vastly different
to the amounts found on Earth. Stars weren’t just hot planets,
as was suggested by the world’s leading astronomer Henry
Norris Russell, they were something else entirely.8
On Earth, the most abundant elements are oxygen, silicon,
aluminum, and iron but stars are made almost entirely from
hydrogen and helium. The astronomers Otto Struve and Velta
Zebergs described Payne-Gaposchkin’s research as
“undoubtedly the most brilliant PhD thesis ever written in
astronomy”9 but her work was largely dismissed (three
guesses why).
Henry Norris Russell even advised her not to publish the
results because it would make her a laughing stock but, to his
credit, changed his mind when he repeated her methods and
found she had been right all along.
The Universe, it turns out, is made almost entirely from
hydrogen and helium. The other elements from which we
derive all the planets are merely trace impurities. This
humbling realization prompted the astronomer Lewis Fry
Richardson (or possibly George Gamow, the origin is unclear)
to write the following poem in tribute to the discovery:
Twinkle, twinkle little star,
I don’t wonder what you are,
For by spectroscopic ken,
I know that you’re hydrogen.
STARS UNCOUNTABLE

If you look up on a clear night sky, out in the countryside


where there’s no light pollution, you can see a pale ribbon of
light stretching from horizon to horizon. The ancient Greeks
thought it was breast milk from the goddess Hera and called it
galaxias kyklos, the milky circle.
Today, we know this glowing stream is made from suns. So
many suns it becomes impossible to count them as individual
points of light so it blurs into a beautiful haze.
The night is filled with what we usually think of as starlight
but in a very real sense what we’re talking about is sunlight.
Our Sun, the source of all our energy, is only one among
billions of others orbiting the super-massive black hole
Sagittarius A*.
If you were to somehow see our galaxy from the outside,
you wouldn’t even know our sun was there amid the glow. It
would be like looking at a cloud and trying to pick out a single
drop of water.
There are somewhere between one and four hundred billion
suns in the Milky Way but it’s hard to know for sure because
we’ve never been outside it to take a picture. And our galaxy
isn’t special either. In 964, the Persian astronomer Abd al-
Rahman al-Sufi saw what looked like a cloud sitting inside the
Andromeda constellation. Little did he realize he had just
discovered our nearest galactic neighbor, confirmed in 1923
by the astronomer Edwin Hubble. It sits about twenty
quintillion kilometers away from us and contains around a
trillion stars.
The telescope that bears Hubble’s name, quietly orbiting
our planet 547 km above ground, has probed even further than
Andromeda and revealed over 170 billion other galaxies in our
local region of space.
If someone were to ask how many stars there are in the
Universe the answer would sound comical. Even the lowest
estimate puts the number of stars at around ten quadrillion in
our local region of space alone.
The only kinds of people who talk in numbers that big are
preschool children who have no idea how ridiculous the
numbers sound and scientists who know exactly how
ridiculous the numbers sound.

HOW DO STARS FORM?

The usual answer to this question never does justice to the


truth. People are usually told that suns are either fires or balls
of burning gas. Both views are tragically inadequate. The
closest we’ve ever come to manufacturing the fabric of a star
here on Earth was on October 30, 1961. That was when the
human race stood in awe and terror as it detonated the Tsar
Bomba on the Russian island of Severny.
It is the most powerful nuclear explosion ever achieved to
date, with a blast radius of around 35 km. To put things in
perspective, our Sun is equivalent to roughly two billion Tsar
Bombas detonating in unison, every second. In one instant, the
Sun casually generates over a million times the amount of
energy our entire species consumes in a year.
Its light provides the energy needed for our crops to grow,
its warmth is what makes water evaporate, giving us rain, and
its gravitational pull is what stops us drifting into the cold
emptiness of space. It’s no exaggeration to say that the Sun
maintains the entire human species and permits it to live. And
it goes far deeper than that.
To understand what’s really going on we’ll need to
consider the effects of the all-pervading force of gravity, which
is usually ignored in chemistry.
All matter in the Universe has a gravity field to it, which
means everything is pulling on everything else. We don’t feel
it but our bodies are loosely gravitating to the objects in the
room around us and they are being drawn back toward us in
return.
The reason you don’t notice this effect is because gravity is
a very weak force (you need an entire planet’s worth to hold
things in place), but while gravity might be weak it is infinite
and has been around since the beginning.
Within the first half-second after the big-bang expansion
started, the earliest particles called photons and neutrinos
(Appendix II again) began colliding, forming the protons,
neutrons, and electrons we already know about. A few
hundred seconds after that, the protons and neutrons joined up,
creating hydrogen and helium nuclei with a tiny bit of lithium
and beryllium thrown in (elements 3 and 4). Then, for the next
380,000 years, nothing happened.10
During this time the Universe was a buffet of free-floating
nuclei and electrons. You wouldn’t have been able to see
anything in front of your face because there was light in all
directions and all of reality would have looked like a milky
fog.
Then, after about 1.6 million years, the temperature
dropped to a breezy thousand degrees and electrons got
snagged by nuclei, forming clouds of hydrogen and helium
atoms. The Universe finally became see-through and gravity
began exerting its influence.
As the hydrogen/helium clouds started collapsing under
their own weight, their gravity fields became more
concentrated, pulling more and more atoms into the mix. Over
millions of years, these clouds condensed into swirling knots
across the Universe, getting hotter and hotter until they
whipped themselves into such a frenzy that the nuclei of the
atoms began to fuse.
Gravity pulled things inward while the heat from fusion at
the core pushed outward. When a truce was finally reached
between these forces, the result was a stable sphere of nuclear
explosion. The very first sun.
The core of a sun like ours reaches a temperature of about
16 million°C, hot enough to vibrate hydrogen and helium
atoms into each other and mash them into heavier elements
like oxygen and carbon. Bigger and fiercer stars can go even
further, burning carbon atoms into magnesium and then fusing
all the way up to iron (element 26). This is how the light
elements are made.

TIME TO DIE

In about four billion years, the hydrogen in our Sun will be


depleted and things will start to cool. The thermal pressure
from within will no longer be hot enough to support its shape
and gravity will dominate, causing everything to contract.
This will temporarily increase the core pressure, giving it a
momentary second wind of heat and inflating the envelope of
gas around the outside of the Sun, making it significantly
bigger. At this point the Sun’s radius will stretch out to
encompass the Earth, burning our beautiful planet to cinders.
As we’ve already said, though, our Sun is peanuts
compared to what else is out there. When big suns reach the
end of their lives, something very different happens. A super-
giant star will keep burning until its entire core has been
converted to iron and, once again, the heat can’t support the
outer layers and gravitational collapse occurs. But this time
we’ve got a bigger star and more gravity so the contraction
happens within seconds. The iron core is too dense to be
compressed so when the outer layer shrinks, it bounces off the
core and the shockwave causes a catastrophic explosion,
which rips the whole thing to pieces.
We call it a supernova and it’s during these violent star-
plosions that iron atoms get fused together, generating
elements all the way up to ninety-two. The star’s body has
been shredded from the inside out and the newly formed heavy
elements are scattered into the dust of space.
And then the whole process repeats. Clouds form, gravity
makes them clump, and suns are born, except now we have
new atoms in the mixture. The clouds are no longer just
hydrogen and helium but colorful mixtures of heavier
elements too.
As this second generation of stars is weaved from the
corpses of supernovae, the heavier elements get sucked into
the star’s rotating gravity field. Some of this material gets
pulled into the furnace but a lot of it forms a ring, encircling
the sun like a moat around a castle.
Clusters of metal and rock gather in the eddies of this
current and eventually congeal into planets. Each planet in a
solar system is made from atoms that began life inside an
ancient star, blown to pieces by the colossal horror of a
supernova.
This is not idle speculation either. Thanks to spectroscopy,
we have witnessed all of these events happening. The
Universe truly is in a cycle of stellar reincarnation with planets
and their inhabitants being generated as by-products.

CHILDREN OF STARDUST

There are stories from many cultures about how we are drawn
from the dust of the Earth and that we are at one with nature.
What science gives us is something far grander: the
reassurance that these are not fairy tales.
The first nine months of your life involved your mother
building you out of the food she ate, but the atoms in that food
came from the Earth and the Earth is made from the remnants
of long-dead suns. With the exception of hydrogen, all the
atoms in your body started their lives in the heart of a sun,
which means you are, as Carl Sagan once observed, made
from star stuff.
The stars you see at night are not transcendent objects
made from ether as Aristotle believed: they are made from the
same material as you. They are your distant relatives and when
you die you will return to them. As our planet reaches its fiery
demise, your atoms will get spread across the Universe and
you will become part of another planet, perhaps even another
living being. Maybe the ancient humans who worshipped the
stars chose their gods wisely.
CHAPTER FIVE

Block by Block
RECORD-BREAKING FLAVOR

Classifying chemicals according to their properties has been a


goal for thousands of years. Today, we use sophisticated
equipment but an astonishing amount can be gleaned using our
senses.
The human tongue is coated with receptors coming in at
least five varieties: sour, bitter, salty, sweet, and umami
(sometimes called savory). If the right-shaped chemical docks
with a sweet-receptor for instance, a signal is sent to the brain
and the food is perceived as sweet. Smell receptors work in the
same way, except there are thousands of potential shapes,
allowing us to distinguish thousands of fragrances.
The food in our mouth is sensed by the tongue and nose
simultaneously. This combination of smell and taste is what
gives each food a “flavor.” That is, with the exception of spicy
foods. They work by accident.
As well as taste, your mouth also needs to monitor the
temperature so you don’t consume things that are too hot. The
heat sensors in your body have names like “TRPV1 receptors”
and there are plenty on the tongue and in the gut. Certain
chemicals are coincidentally shaped in such a way that they
trigger the heat sensors and tell your brain the area is hot, even
though the rest of your mouth is cold. The resulting confusion
is what we perceive as “spiciness.”
In 1912, the American scientist Wilbur Scoville devised a
test to measure the spiciness of food mathematically and we
still use it today. The spicy chemical is dissolved in water
repeatedly until it can no longer be tasted by a panel of
volunteers. The number of dilutions required to make the taste
imperceptible is then expressed as a Scoville Heat Unit or
SHU.
Since the tongue is good at tasting even trace amounts of a
chemical, SHU values are typically enormous. The oil from a
jalapeño pepper is undetectable after about 8,000 dilutions so
jalapeños are given an SHU of 8,000, while something like
Tabasco sauce scores closer to 50,000.1
The world’s spiciest pepper at the time of writing is the
Dragon’s Breath chili, bred into existence by Welsh spice-
master Mike Smith and possessing an SHU of over 2.4
million.2 That’s basically the same as pepper spray. This chili
is so hot it would trigger anaphylactic shock if you ate it, but
that’s nothing compared to the world’s spiciest chemical:
resiniferatoxin.
Produced in the latex of Euphorbia resinifera plants (also
called resin spurges), nobody has ever carried out taste trials
with resiniferatoxin because it is acutely toxic and causes
severe burns to the skin, meaning we have to calculate its SHU
indirectly.
A study carried out by Arpad Szallasi in 1989 (on rats)
found that resiniferatoxin was one thousand to ten thousand
times better at binding to TRPV1 receptors than the chemical
in chili peppers.3 Since we know chili peppers have an SHU of
around 16 million, resiniferatoxin is going to score somewhere
in the region of 16–160 billion SHU. That’s spicy enough to
kill you.
There are many other chemicals that have record-breaking
effects on our senses too. The sweetest chemical, so sickly it
induces vomiting, is called lugduname, 230,000 times sweeter
than table sugar.4
The darkest chemical, so black you can’t even see a torch
shining on it, is called vantablack.5
And the worst-smelling chemical is a tie between
propanthione and methanethiol, substances that have caused
mass unconsciousness, spontaneous vomiting, and even death
from smelling them at a distance.6

THE ELEMENT WAR

The first attempt at properly identifying elements was done by


none other than Pythagoras himself, although it was a little bit
weird. Most people know Pythagoras from the square-on-the-
hypotenuse law they learned in school. What’s not usually
mentioned is that Pythagoras was also probably the world’s
first cult leader.
Not a great deal is known about the Pythagorean order
because revealing their secrets would get you exiled, but we
know they were forbidden from touching white chickens or
eating beans.7 Pythagoras was murdered because an angry
mob chased him to the edge of a bean field and, rather than
entering it, he turned toward the crowd and chose a fatal
beating.8
The only other thing we know about the Pythagoreans is
that they considered numbers to be elements. Pythagoras and
his cult worshipped numerical order, believing math to be the
true face of reality. Their table of elements was simply a list of
numbers going from one upward. Okay then.
Other people chose more tangible substances as their
elemental matter. We met Heraclitus, who proposed fire as a
candidate, in Chapter 1. Thales, who we met in Chapter 4,
favored water because it took many forms, while the
philosopher Anaximenes declared air to be the purest material,
and so on.
It was a man named Empedocles who brought order to all
the squabbling in the fifth century BCE. Rather than backing
any of the other thinkers, he took the diplomatic approach and
suggested that maybe everyone was right. Perhaps there
wasn’t just one element but several.9 Empedocles’s periodic
table would have looked like this:
This surprisingly simple solution ended the arguments and
everyone was happy. Thales could keep his water, Anaximenes
his air, Heraclitus his fire, and Pythagoras was dead in a bean
field so nobody cared what he thought.
Nowadays, some people still think of these substances as
being elemental but there is really no justification for this.
They were chosen for peace-keeping politics rather than
accurate knowledge, although sadly a lie can remain popular if
people like it and it is easy.

THE TABLE MAKES ITS DEBUT

Once Antoine Lavoisier discovered that air was a


nitrogen/oxygen mixture and water was a hydrogen-oxygen
compound, scientists abandoned Empedocles’s four-element
idea and began burning or dissolving everything they could lay
their hands on to obtain the true elements.
By 1789, a lot of new ones had been discovered so
Lavoisier gathered all the information and published a
complete list, totaling thirty-three elements in all.10
He put them in four categories: gases, which were invisible
but occupied space; metals, which were shiny and burned in
oxygen; non-metals, which could be used to make acids; and
earths, which didn’t fit the category of metallic or acid-
making.
Lavoisier’s table was the first not to be based on guesswork
or gut feeling and it looked like this:

The substances indicated with an asterisk were later


discovered not to be elements but for a first attempt his table
was pretty good.
Other chemists had their own methods of grouping things,
of course. The German chemist Johann Döbereiner grouped
elements into families of three based on how similarly they
behaved. The metals lithium, sodium, and potassium behave
identically, for instance. They react violently with water,
tarnish in air, and can be sliced with a knife (if you’ve never
had the joy of cutting a piece of lithium metal, it feels like ice
cream straight from the freezer).
A similar observation worked for sulfur, selenium, and
tellurium. All three were powdery solids that reacted with
oxygen to produce strong-smelling compounds. Döbereiner
called these groups triads, but there was no apparent reason for
the patterns.11 The finished table of elements would have to
somehow explain these mysteries.

A MUSICAL INTERLUDE

The most famous stab at a periodic table, before the one which
actually worked, was a doomed attempt by the Englishman
John Newlands in 1863.12 Methods had already been devised
to measure the weights of atoms pioneered by Swedish
chemist Jöns Berzelius (who also introduced the element
symbols we use today)13 so Newlands obtained the data and
wrote a list of the elements in order of ascending mass. As he
did so, he discovered that the elements almost followed a
cyclic pattern the way musical notes do.
In Western music theory, there are only seven principal
notes. If you start at any particular tone and play up the scale
you’ll discover that the eighth note is identical to the first, just
a higher version. Note nine is a higher version of note two and
so on. One complete set of notes is called an octave and the
notes spiral up and up until the human ear can no longer catch
them.
John Newlands applied the same logic to his table of
elements, claiming there were seven categories that repeated
over and over as we got to higher masses. The first seven
elements made the first row, while the eighth element would
be the first entry on row two, having similar properties to
element 1 directly above it.
He called the seven columns of his table “families” and the
eight rows “periods,” meaning something that repeats
regularly. Thus, John Newlands introduced the idea of
elements being “periodic.”14
The idea of periods turned out to have some truth to it, but
his table had one minor flaw, which can sometimes prove
inconvenient for a hypothesis: it was wrong.
At the time Newlands composed his table (pun very much
intended), there were sixty-three elements known, which
didn’t fit into an eight by seven grid. So rather than adding an
extra column or abandoning the octaves idea, Newlands
shoved a bunch of elements into the same grid squares.
The metallic element cobalt, for instance, having the
audacity to exist, nudged later elements out of their correct
families, which didn’t match the hypothesis. Newlands
decided that cobalt and nickel were therefore the same
element.
They aren’t. (Although, fun fact, both get their names from
German sprites, Kobold and Nickel.)
Newlands knew these elements weren’t the same as each
other but this kept his table neat, so best not to worry about it.
He then had to do the same thing with awkward vanadium and
again with lanthanum. In doing so, Newlands fudged the data
to fit his idea. We have a word for that in science: cheating.
It would be like claiming there were three types of animal:
cows, goldfish, and pigeons—then when someone shows you
a tiger you decide it’s a cow really and put it in the same
column.
Newlands also cherry-picked the elemental features. Cobalt
is a lustrous metal with magnetic properties but his table
aligned it with fluorine, chlorine, and hydrogen, all reactive
gases. Newlands was happy to point out that chlorine,
hydrogen, and fluorine belonged together but ignored the fact
that cobalt didn’t.
As a scientist, your job is to recognize when your
hypothesis has failed. If nature says your idea is wrong then
you get a new idea, you don’t tell nature what to do.
As a result, Newlands’s table was rejected by the scientific
community of the day, although the story does have a happy
ending. Every scientist has published a dodgy idea at some
point, so scientists are a forgiving bunch who try not to hold
grudges. If one idea turns out to be wrong, your others are still
given a fair hearing. It’s useful to have that approach because,
although Newlands’s octave hypothesis was wrong, his idea of
periodic repetition turned out to be on the money. Elements do
obey a cyclic pattern but a much more complicated one than
he had assumed. He was, for this realization, awarded the
Davy Medal for Chemistry by the Royal Society in 1887.

THE DREAMER

Dmitri Mendeleev was born in Siberia in 1834, the youngest


of probably thirteen children (historians can’t agree on the
number, but I’m sure his parents knew).
When his father went blind, Dmitri supported the family
financially by tutoring science and, according to those who
saw him in action, he was a fantastic communicator, full of
passion and enthusiasm for both the subject and the art of
explanation.
At the age of fifteen, his mother decided he needed a higher
education and took him across Russia on foot, applying to as
many universities as they could along the way. The expedition
took close to a year and sadly her health worsened as the
months drew on. She died when they reached St. Petersburg,
but lived long enough to see her son get admitted to study
joint-honors in chemistry and teaching at St. Petersburg State
University.
She would have been proud of his accomplishments, as he
soon became one of the most outstanding chemists in Russia,
with a reputation for writing huge textbooks from memory in a
matter of months, and helping establish the country’s first oil
refinery in Tutayevsky.
He was also an imposing character, who shaved his beard
once a year and had fiery clashes with other students and
professors. His greatest contribution to science though was
creating the very first periodic table that actually worked.15
A few days prior to his breakthrough, Mendeleev made a
deck of playing cards with elements instead of suits on their
faces. He invented a version of solitaire based on chemical
properties and hoped it would help him discover a deep pattern
about their organization.
According to his friend Alexander Inostrancev, Mendeleev
had been awake for three days and nights playing the game
when he finally collapsed from exhaustion on the afternoon of
February 17, 1869.
Mendeleev fell asleep surrounded by his playing cards and
had the most vivid dream of his life. In the dream, he saw the
playing cards dancing before his eyes and dropping into place
perfectly, revealing the pattern for which he had been
searching.16 The elements did follow a cycle, but nobody had
figured it out because there were still elements missing!
Up until then, people had been discovering elements at
random and grouping them based on color, reactivity,
conductivity, thermal properties, and anything else you could
name.
Mendeleev realized that the elements were arranged in a
sequence of increasing mass but that some were still hidden
inside rocks. The elements that seemed to be in the wrong
place weren’t: they were just next to elements that were
unknown to chemists of the day.
Element 32 hadn’t been isolated yet, nor had 61 or 72. If
we assumed Mendeleev’s law of increasing integers worked,
we should find elements that matched those values and, sure
enough, germanium, promethium, and hafnium were
eventually identified and slotted into their respective gaps.

THIS WAY MADNESS LIES

By 1932, we knew that elements were made of atoms,


themselves made from protons, neutrons, and electrons. But if
every atom was made of the same three particles, why were
they so different from each other?
Take element 35, bromine. It’s a thick, mauve liquid that
sets fire to metal and corrodes human skin. The next element
is number 36, krypton. That’s a harmless, invisible gas with no
odor or reactivity. The only difference is that krypton has one
extra proton/electron than bromine, so why don’t they behave
similarly?
And what can we make of Döbereiner’s triads? Elements
29, 47, and 79 are copper, silver, and gold—all malleable
metals with a lustrous finish. Why do those three numbers in
particular end up with the same properties?
Why is element 4 a shiny solid while element 5 is a brown
powder? Why is element 9 one of the most reactive known to
man but element 10 one of the least? Why do elements 11 to
14 conduct electricity while elements 15 to 18 do not?
Any attempt to find order resulted in failure and a
hypothesis has to account for all evidence, not just a
convenient portion of it. If we couldn’t use the
proton/neutron/electron model to account for the differences in
behavior then we would have to abandon it.
The only conceivable explanation was that although each
atom was made of the same three particles, they were
somehow arranged differently in space. Democritus had
already suggested that atoms came in different shapes (fire
atoms were spherical, which allowed them to move easily,
while “bitterness” atoms were sharp and jagged). Could he
have been on the right track?
The answer finally came when physicists discovered one of
the most important theories in modern science, the one that
gave the periodic table its final form. Quantum mechanics.
CHAPTER SIX

Quantum Mechanics Saves the Day


QUANTUM CRASH COURSE

Quantum mechanics is infamous. Everyone has heard about it


and its reputation for being weird (a reputation that is well
deserved, by the way). However, in recent years, some of the
vocabulary has been hijacked by spiritualists to mean all sorts
of unrelated things, which sadly confuses the issue. Don’t
misunderstand me; there’s nothing wrong with talking about
spirituality but repurposing words from quantum mechanics to
mean something else is unhelpful. So we’ll tread carefully.
The first thing to say is that quantum mechanics is not one
idea but a sophisticated collection of theories that explain the
world at its smallest level. The behavior of electrons, the
nucleus, light, and their interactions are all explained by
quantum mechanics so it is of great importance to chemistry.
Covering it in detail would take a separate book entirely so
we’ll limit the discussion to the part, developed by Austrian
physicist Erwin Schrödinger, that helped build the periodic
table.
Schrödinger caused a lot of discomfort during his life and
was politely asked to leave a number of universities and
institutions. This wasn’t because of his academic
achievements, which were outstanding. It was because he
lived in a three-way relationship with his wife Annemarie and
their girlfriend Hilde. He also wore a lot of bow ties.
Scandalous.
Schrödinger’s most important contribution to science is
called the Schrödinger wave equation. It’s the equation that
tamed the periodic table and explains why elements behave the
way they do. It looks like this:

I know equations can sometimes put people off but this one
is vital to the story, so we can’t just brush it under a rug. I’ve
included a short explanation of what it means in Appendix III
if you’re feeling adventurous but don’t worry, we can still
understand what the equation does without having to go into
any mathematical detail.
Nobody is sure how Schrödinger came up with his equation
because there are no clear records of him deriving it. Some
claim he simply woke up one morning, went downstairs, and
wrote it based on gut feeling. It was only later that it was
tested and proven correct.
What the equation does is tell us where electrons are likely
to be as they zip about the nucleus. You start by taking the
electron’s properties (things like its mass, velocity, etc.) and
then figure out how much attraction there is from the protons
of whichever atom you want to describe.
By solving the equation for a given atom we can map out a
three-dimensional region of where electrons are going to be
and what patterns they will trace out in space.
When we do this, we find that electrons don’t move in
circular orbits at all. They surround the nucleus in regions that
come in a variety of shapes, the same way animals inhabit
different-shaped enclosures at a zoo. We call these regions
“orbitals” or sometimes when we’re being lazy, “electron
clouds.”
Some electrons hang out in spherical orbitals while others
occupy a dumbbell-shaped region protruding from the top and
bottom of the atom. Each orbital can hold up to two electrons,
so the more electrons you have on your atom the more orbitals
end up being used and the more extravagant your atomic shape
becomes.
The reason certain orbital shapes arise is because electron
movements are sort of wavy. They don’t move in simple lines
like marbles but seem to ripple as they travel from one point to
another. Since ripples can only come in certain shapes (you
can’t have half a wave, for example) so do the electron
orbitals.
A boron atom, which has five electrons, will distribute
them into orbitals shaped like the diagram on the left on page
65. Carbon, however, has six electrons so a new orbital shape
gets introduced and the atom looks like the diagram on the
right.
The fact that different atoms come in different shapes
explains why they have difficult chemical behaviors. They
stack together differently, bond at different angles, fit into
different spaces, and so on.
Solving the Schrödinger equation for a particular element
explains why it can be different to the element next door. Just
because they have a similar number of electrons doesn’t mean
they will have the same shape. It also answers the question
every student asks when they see the periodic table for the first
time.

WHY IS IT THAT SHAPE?

A table is supposed to be a neat rectangle with columns and


rows. You know, like the one Lavoisier created. The periodic
table we use today looks like a chimp accidentally vacuumed
up a computer keyboard and tried to glue it back together with
silly putty. It doesn’t look table-ish at all. So, who came up
with the design and why did everyone else say “Yup, looks
good to me”?
The man to thank is one Alfred Werner, the Swiss Nobel
Prize–winning chemist who published a short article in 1905
with the catchy title “Contribution to the development of a
periodic system.”1 It was here that the periodic table first took
shape.
Let’s consider the first ten elements. Actually, let’s not.
Let’s ignore elements 1 and 2 and begin with element 3. (I’ll
explain in a moment.)
We could line the elements up in a nice long row and be
done with it:

But now we can do better thanks to the Schrödinger equation.


The first two elements of this row put their electrons into
spherical orbitals while the next six go into dumbbell-shaped
orbitals. This means we can split the line like so:

The next eight elements have the same orbital shapes. The
atoms will be bigger but they will otherwise have very similar
chemistry. To represent this, we use Newlands’s periodic idea
and add a second row to our table, still dividing into two
blocks:
Each column of elements represents a particular orbital
shape. The only difference is that, as we go down, the orbitals
get larger.
When we get to element 21 a new shape gets introduced
(quantum mechanics is like that). The outer electrons of the
atom of this element, scandium, up to element 30, zinc, are
shaped like bundles of balloons rather than dumbbells so we
need to introduce a new block to the table. Element 31 goes
back to the dumbbell shape and so our table now looks like
this:

It’s a bit irritating that nature insists on introducing weird


orbitals when we get to larger elements, but that’s why the
table is an awkward shape. It’s because nature is.
Now, if you read from left to right across a row (period)
you’re reading in ascending proton number, while the column
(group) tells you what shape the atoms are going to have.
Reach the end of one period and you just go on down to the
next one.
When Alfred Werner included all the known elements and
orbital shapes, the table ended up like this:
Suppose you wanted to know about iodine. By counting
from left to right you learn that it is element number 53,
meaning it will have fifty-three protons and fifty-three
electrons. You can see it’s in the right-hand block (dumbbell
shaped) so you also know what angles it will make with other
atoms.
Directly above it are chlorine and fluorine, both colorful
non-metals. Iodine is in the same column so it will probably be
a colorful non-metal too, but with a higher density as it’s on a
lower period. Sure enough, we discover that these are exactly
the properties of iodine.
You can even use the periodic table to predict the properties
of elements nobody has ever seen. Directly below iodine is
astatine, the rarest element in the Earth’s crust (less than 1 g
exists on the whole planet), but if we had a sample it would
probably behave like a denser version of iodine. God bless
quantum mechanics.

ARCHITECTURAL SIMPLICITY

You probably know from your periodic table T-shirt, mouse


pad, shower curtain, pencil case, and notebook (I own all of
these and assume everyone else does too) that the above
pictures aren’t quite there yet.
This fully expanded version of the table is rather
cumbersome, so for simplicity we take one of the blocks,
scooch it down to the bottom, and slide the others in to meet
each other.
This form of the periodic table was proposed by Glenn
Seaborg in

1945 and soon became the standard thanks to its simplicity


and the fact that Seaborg did a lot of work to popularize
science.2 But obviously we’ve missed out elements 1 and 2.
Hydrogen and helium are both spherical atoms meaning
they belong in groups 1 and 2 respectively:
That’s what Harvey White did with his periodic table design in
1934 and it’s what Schrödinger would have wanted.3
Unfortunately, due to their small size, H and He don’t behave
quite the same as the other elements in that block.
They actually have more in common with elements on the
other side of the table so if we placed them according to
reactivity we would end up with things looking like this:

This is what Ernst Riesenfeld did with his periodic table in


1928 (and what Mendeleev would have wanted).4
Glenn Seaborg couldn’t make his mind up about where to
put these two fiddly elements, sometimes drawing them on
one side and sometimes the other (and briefly putting
hydrogen in two groups in 1945).5
Eventually, the general agreement was to acknowledge
both the electron-orbital work of Schrödinger and the
chemical-properties work of Mendeleev.
So we split them up and put them at each end of the table,
one for each scientist. It’s not logical but it’s a nice tribute to
the two men who built it. And voilà, we have ourselves a
periodic table.
CHAPTER SEVEN

Things that Go Boom


THE MOST EXPLOSIVE EXPLOSIVE

With the exception of nuclear weapons, most explosives work


in the same way. First, a material is synthesized, which is
highly unstable. In chemical terms this means it will fall apart
given the chance. Second, the material is provoked, giving it
the chance to break up and rearrange into stable substances.
During this rearrangement a whack-ton of energy gets released
(whack-ton being the technical term) in the form of light and
heat.
In addition, a small amount of solid or liquid explosive will
expand rapidly to become a large volume of gas. This sudden
expansion combined with lots of heat and light is what we call
an explosion.
Some substances are so unstable that even a little bit of
agitation will cause the reaction to begin. Gunpowder only
needs a candle flame to decompose, while TNT requires
nothing more than a spark. You can even buy bang snaps,
children’s toys consisting of small paper parcels filled with
silver fulminate, a chemical that explodes when struck. The
parcel is thrown at the ground and the impact causes a loud
snap.
Fireworks work on a similar principle. A powdered metal is
launched into the air and once the fuse within has burned, a
detonator gel will trigger, converting it into a gas. Each fleck
of powder is sprayed outward as the gas expands, becoming so
hot they start reacting with the atmospheric oxygen, giving us
sparks.
We’ve already seen in Chapter 4 that different elements
give out different types of light, so by picking particular
metals we get particular spark colors. Sodium will turn yellow,
barium goes green, copper blue, and strontium brick red.
Purple fireworks are famously difficult to achieve and usually
involve a mixture of copper and strontium together.
All explosives rely on the chemicals within being unstable,
and the most unstable chemical ever created is called
azidoazide azide, synthesized in 2011 by Thomas Klapötke.
Although why anyone would want to make this chemical is
beyond me.
It contains fourteen nitrogen atoms and two carbon atoms
clumped into branches around a tight ring, all squished
together without much room. The bonds between the atoms
are so strained that they spring apart under any circumstance.
When Klapötke tried to dissolve the chemical in water, it
exploded. When he tried to move it across his lab, it exploded.
When he breathed in its general direction, it exploded. It even
exploded when infrared light (the kind emitted from a TV
remote control) was shone at it.1
The best explosives, of course, are those that detonate on
cue. They have to be stable enough to be moved, but unstable
enough to still detonate. Azidoazide azide would be a poor
choice, for the same reasons chlorine trifluoride was a terrible
choice of rocket fuel. You’re better off with good old
dynamite, invented by none other than Alfred Nobel.

THE MERCHANT OF DEATH IS DEAD

When somebody dies, we tend to say nice things about them


because it’s taboo to speak ill of the dead. Such was not the
case on February 12, 1888, when Alfred Nobel’s obituary was
published. A French newspaper allegedly ran the headline
“The Merchant of Death is Dead” and went on to say, “Dr.
Alfred Nobel, who became rich by finding ways to kill more
people faster than ever before, died yesterday.”2
A lot of people were not happy with a newspaper
remembering the great scientist in such a way. That included
Alfred Nobel himself, who managed to read his obituary
owing to the fact that he was not really dead. The story goes
that Alfred’s brother Ludvig had died and the newspaper
mistook the two brothers, prompting them to publish their
powerful attack.
Nobel was a very talented chemist who had invented
dynamite twenty-one years earlier. Initially he intended it for
use in mining, but it had obvious military applications.
Apparently, realizing what his legacy was, Nobel decided to
alter his will and left his considerable fortune (then over thirty-
one million Swedish kronor, equivalent to nearly eighty
million US dollars today) as prize money to people who did
things “for the greatest benefit of mankind.” The prizes were
to be awarded for achievements in the three sciences,
literature, and the promotion of peace: the Nobel Prizes.3
The newspaper that ran the obituary is reported to be the
Idiotie Quotidienne and I have desperately tried to find
documentation to confirm its existence, but sadly I can find
none.4 The fact the newspaper’s title translates to the Daily
Idiocy might be a clue that the story is a hoax and indeed some
of Nobel’s biographers have dismissed it as a persistent
rumor.5
It may be an apocryphal morality tale or perhaps an
embellishment of Nobel’s reaction to his brother’s death.
Whether it’s true or not, the Nobel Prizes are still considered
the most prestigious awards it is possible to receive in science.
The prize money is substantial, numbering in the millions, and
it comes from Nobel’s estate built entirely on dynamite. Not
literally, obviously. That would be stupid.
The way dynamite works is simple. You take a large
amount of silicon-based rock powder and soak it in a chemical
called nitroglycerine. Pack this into a tube and stick a fuse in
the end. As the fuse burns, the heat is transferred to the
nitroglycerine-soaked powder and—boom!
Nitroglycerine is one of the unstable compounds I
mentioned earlier. Composed of carbon, nitrogen, oxygen, and
hydrogen atoms, it is a chemical that will self-react, i.e. one
nitroglycerine particle will react with another to produce a
bunch of gases, mostly carbon dioxide and water.
These gases expand to over twelve hundred times their
original volume and reach a temperature of 5,000°C. The
reaction is also fast with the expansion and heating taking
place in under a microsecond.
All of this comes down to the question we’re going to
answer in this chapter—why do chemical reactions happen at
all? What do we mean when we say a chemical is unstable and
how do atoms bond in the first place? To make sense of it all,
we’ll need to dive a little deeper into the quantum ocean.
GIVING CHEMISTRY A BAD NAME

The word chemistry comes from alchemy, but a better name


for the subject would be electronics, because chemical
reactions are all about electrons. The nucleus of an atom is
tiny compared to the overall radius so it’s the electrons on the
outside that are interacting with everything.
And electrons are always on the move. Were an electron to
cease moving it would simultaneously cease to exist because
movement, like the charge it possesses, is part of an electron’s
identity. A stationary electron exists no more than a four-sided
triangle.
So, if we take movement as a given, there are only two
things an atomic electron can really do. It can move outward
away from the nucleus or it can move inward toward it. These
two behaviors underpin almost every chemical reaction you’ll
meet.
Let’s revisit the concept of orbitals that we got from the
Schrödinger equation. They’re the regions around a nucleus
where electrons spend their time.
Orbitals are the permitted electron territories, but electrons
aren’t confined to live their entire lives in the same one. They
can hop around. When an electron jumps from one orbital to
another it’s called a “quantum leap” and it can happen between
any two orbitals, even ones that are empty.
Obviously, electrons prefer to occupy an orbital near the
nucleus because it carries the opposite charge but they don’t
always get their way. If the innermost orbitals are inhabited,
other electrons have to make do with being further out.
An atom is really a hustling, bustling place where the
orbitals nearest the nucleus are considered prime real estate
and every electron wants to move in. If one of the inner
electrons happens to vacate their orbital for some reason, an
outer electron will quantum leap to replace it.
These quantum leaps don’t happen at random, though. The
rule that ultimately accounts for chemistry is as follows: if an
electron absorbs a beam of light it gets bumped to an outer
orbital and if it emits a beam of light it drops to an inner one.
Some types of light have more ability to promote electrons
while others have less. A blue beam can promote an electron
to a far-out orbital while red light might only nudge it up by
one level. In the same way, electrons from far-out orbitals have
the ability to release blue light when they drop, while electrons
already near the nucleus might only release red.
This is how the fireworks and spectroscopy we’ve already
mentioned work. Every atom has a unique orbital arrangement
so every atom emits or absorbs a unique light spectrum. When
the electrons start jumping from orbital to orbital, the distance
they jump determines what kind of light gets emitted or
absorbed depending on which way they’re traveling.
The question everyone usually asks at this point is why
electrons absorb or emit light in the first place. I’m afraid the
answer is because that’s just the way nature is. It’s just one of
the fundamental laws that were established during the big-
bang expansion. The same way balls roll downhill as they
obey the laws of gravity, electrons release and absorb light as
they obey the laws of quantum mechanics.

ABILITY AND STABILITY

We’ve talked about some beams of light having more ability to


promote electrons than others, and in science we sometimes
replace the word ability with the word energy. I’ve mostly
avoided using it up until now because it’s a word fraught with
difficulty and misconception.
People talk about energy as if it were a thing being
transferred from place to place, but it isn’t really. You can’t
hold a lump of energy but a lump of matter can possess the
ability to bash into things or the ability to explode, i.e. it can
possess energy.
In the context of quantum chemistry, energy means “how
capable a beam of light is to push an electron into a higher
orbital.” You’ll sometimes hear scientists say that electrons in
outer orbitals have “absorbed energy” and that this energy gets
released when they drop. This is convenient shorthand but we
have to be clear: it is light that gets absorbed and emitted.
Light has the ability to promote electrons and therefore
possesses energy, but energy is not an actual thing.
The opposite of ability is what we mean by “stability” and
it’s a measure of how much energy an electron has lost when it
drops down, or how reluctant it is to shift upward from its
present orbital.
An electron from an inner orbital, close to the nucleus, is
less willing to change because it is happy where it is. We
describe it as being chemically “stable.” An electron in a
higher orbital with a lot of energy (ability to release light) is
very unstable, however, because it is not happy and will
change given the chance.
The diagram below shows what happens when an electron
absorbs a beam of light. It jumps from a low-energy orbital to
a high-energy orbital, becoming unstable.

The next diagram shows the reverse process. This is a high-


energy electron dropping down to a more stable orbital. The
only difference is that light is being emitted here rather than
absorbed.
Ability and stability are always at odds with each other and
govern an electron’s reactive behavior. Gaining energy means
losing stability and vice versa. This trade-off between ability
and stability is what determines whether a reaction will happen
or not.

SHAKE, RATTLE, AND ROLL

Different kinds of light will produce different kinds of effect


on an atom. Infrared light, which is too low in energy to
interact with the electrons in our eyes so we can’t see it, will
cause the orbitals themselves to stretch and twist rather than
shunting electrons between them. Microwaves do something
similar except they cause the atom to spin, rather than twist
and bend.
If you beam atoms with infrared or microwave light the
result is that the atoms start dancing around and bashing into
each other, exchanging energy. Ultimately this happens via the
same light-transfer mechanism (electrons on one atom release
light to electrons on the other, promoting them to a higher
orbital/making the atom twist or spin more) but it’s quicker
and more convenient to talk about atoms colliding and
transferring energy.
These twists and spins of the atoms are what we call heat
and it’s why you feel warm when infrared or microwave light
hits your skin. It also provides the basis of microwave ovens
by causing the water inside a piece of food to jiggle.
Obviously, the hotter a sample of chemical, the more likely
it is the atoms will bump into each other and trigger orbital
rearrangements/twists/spins. Or, put another way, heating most
reactions tends to make them happen faster.

UNITED WE FALL

Imagine being an electron tethered to an atom’s nucleus. If


another atom approaches, its nucleus can draw you toward it at
the same time. If the pull is strong enough you can be dragged
into a position halfway between both nuclei and you are no
longer occupying an atomic orbital but a “molecular orbital.”
A molecular orbital is known by a more common name: a
chemical bond.
If the molecular orbitals are at lower energy than the
atomic orbitals with which we started, then electrons on two
approaching atoms can drop into a molecular orbital together,
releasing light as they go. A bond between the atoms is formed
and we have carried out a chemical reaction.
In the case of the hydrogen/oxygen reaction, the hydrogen
and oxygen units are floating freely, but when they react the
electrons on each atom slot into molecular orbitals, linking
things together and forming a bonded H2O molecule.
All the energy from these dropping electrons is released as
light both visible and infrared (heat), creating the explosions
first seen by Henry Cavendish.
It doesn’t have to be atoms with which we start either, it
can be molecules. The bonds of a nitroglycerine molecule are
at very high energy so they gladly break down into molecules
with more stable orbitals, like carbon dioxide and water. A lot
of energy gets kicked out in the process as all the electrons
drop and we see the result as an explosion.

LET’S GET IT STARTED

The basis of chemistry is simple: start with one set of orbitals


and finish with another. Prizing your original molecules apart,
however, can be difficult. Electrons in a molecular orbital
don’t necessarily know there’s a better deal to be had, so we
have to give them a kick of energy in order to let them fall into
the arrangement we want.
An analogy would be to picture a coat hanging on a coat
hook. The coat will sit there until the end of time, even though
it would achieve greater stability by dropping to the floor. That
won’t happen because you have to put energy into the system
first. It’s only when you lift the coat up a few centimeters,
freeing it from the hook, that you give it the option of falling
into a more stable configuration.
Electrons are exactly the same. We need to excite them first
and get them out of their orbitals before they can drop into
new ones.
A stable molecule like water can be thought of as having a
coat hook several meters long. You’d have to get on a ladder
and lift the coat all that distance to get it free. And once you let
go, it would probably just fall right back onto the hook again.
That’s why water reacts with hardly anything.
Nitroglycerine, on the other hand, is like a coat hook a few
millimeters long, positioned over a cliff. A tiny nudge (say,
from a burning fuse) is enough to get the electrons out of their
orbitals, and the subsequent energy drop is enormous.
Or you could think of it like a LEGO® model. If you want
to make something new, you have to put energy in and
separate the blocks. It’s only when everything is broken down
into its constituent parts that you can form something else.
Whatever the reaction is, chemistry is about persuading the
electrons to jump out of their starting orbitals and into the ones
you want. How hot do you need to get it? What shape does
your starting molecule need to be? What by-products do you
get? What do you do if your reaction doesn’t work? How
many molecules will rearrange and how many will fall back
into their original positions?
Although a myriad of complexities can arise in the lab, the
overall premise is simple. Push the electrons up and let them
drop down.
CHAPTER EIGHT

The Alchemist’s Dream


THE MOST EXPENSIVE ELEMENT YOU’VE NEVER HEARD OF

On April 3, 2017, the Pink Star diamond was sold at auction to


Chow Tai Fook Enterprises for a cool seventy-one million
dollars.1 At the time of writing this is the largest sum of
money ever paid for a gemstone.
For perspective, the Hope diamond was sold in 1908 to
Selim Habib on behalf of the Sultan of Turkey for $200,000,
then resold in 1911 to Evalyn McLean for $154,000. In 1958,
it was gifted to the Smithsonian Institution in Washington, DC,
insured for one million dollars and rumored to be worth even
more today.2
Diamonds are pure carbon so perhaps it would be fair to
call carbon one of the most expensive elements on the table.
Then again, charcoal, which is also made from pure carbon,
retails for a few dollars at any supermarket. So perhaps it’s one
of the cheapest.
We treat gold as a more valuable metal than silver but in
the 1890s the winner of an Olympic event was presented with
a silver medal rather than a gold one. Record companies
reward artists with a platinum album as their highest accolade,
but platinum sells for fifteen dollars less per troy ounce than
gold on the open market.
Rhodium and palladium, used to make catalytic converters
in cars, currently have a similar value to platinum but enjoyed
a brief spike in 2008 when their value increased ten-fold,
making them more valuable than gold for a month. Things are
only worth as much as someone is willing to pay for them, and
the elements are no different.
Plutonium is one of the most expensive materials on Earth
for obvious reasons, with a value of $11,000 per gram
(according to the US Department of Energy), and it’s often
reported as being the most expensive element.3 But there is
one other, rarely discussed, that outranks it. Californium,
element number 98, is used as a starting agent in nuclear
reactors and sells for a titanic twenty-seven million dollars per
gram.4 The Pink Star diamond weighs about 12 g, meaning
californium is over five times more expensive gram for gram.
What makes it so pricey is that californium does not occur
in nature. It’s an element we have to make for ourselves.

ALL OF THEM WITCHES

Before the discovery of phosphorus and the fire experiments


of the eighteenth century, chemical research was a mess.
Armed with a mixture of Judaeo-Christian symbolism, ancient
fairy tales, and the works of a Persian writer named Jabir ibn
Hayyan, rigorous testing of chemicals was ignored and fact
was mixed with superstition.
The resulting field was called alchemy, an Arabic-derived
term which comes from the Greek chemia, meaning black
magic. Nobody was trying to find substances that were
elemental during that period. Instead, they were trying to find
substances they more or less made up.
“Alkahest” was one, thought to be the ultimate acid capable
of dissolving anything. The “elixir of life” was another,
thought to prevent the onset of death, and the “panacea” was
yet another, thought to be a medicine capable of curing all
illnesses.5
Above all, though, the goal of the alchemists was to
generate a material called “philosopher’s stone,” which could
turn other metals into gold. Nobody knows who came up with
the idea of philosopher’s stone but rumors of its existence had
been circulating since the thirteenth century.
The author of a medieval encyclopedia, Vincent of
Beauvais, claimed that God had imparted knowledge of
“transmutation” to Adam, who passed it on to Noah and so
forth. His source for this seems to have been his own
imagination although The Book of Sydrac, an anonymous
thirteenth-century text, tells a similar story, so it was obviously
a common idea at the time.6
One of the earliest recorded references to the phrase
“philosopher’s stone” is in a 1610 play called The Alchemist
by Ben Jonson, which suggests that Adam was told how to
make the fabled substance.7 After he got kicked out of Eden,
presumably he forgot the recipe. Nice going, Adam! First you
lose a rib, then you lose the philosopher’s stone recipe. What’s
next? Your second-born son?
Alchemy did give us knowledge about various chemical
reactions, not to mention Brandt’s discovery of phosphorus,
but it had no structure to it and there was more guesswork than
anything else.
The problem with trying to turn one element into another is
that an element’s identity is determined by the number of
protons in its nucleus and changing that isn’t a simple matter
of mixing things in a test tube.
As we saw in the previous chapter, chemistry is all about
manipulating electrons. The nucleus is too small and hidden
for us to have any impact on it. Simply put, electrons can
dance to any tune you want but if the nucleus remains
untouched the element remains the same.
And yet suns are constantly transmuting hydrogen into
helium so there is obviously no law of science that forbids it
from happening. To mimic the technique here on Earth would
take superhuman powers. Speaking of which …

THE ORIGIN OF SUPERHEROES

Peter Parker got his Spider-Man powers when he was bitten by


a radioactive spider and his DNA became irreparably altered.
Bruce Banner got caught in the blast of an atom bomb and was
belted by radioactive gamma rays, turning him into the Hulk.
The Fantastic Four were caught in a storm of radioactive
cosmic rays, Daredevil was splattered with radioactive waste,
and Jean Grey of the X-Men (in the original storyline) released
her telekinetic potential from flying a shuttle through a
radioactive solar storm.8
Radioactivity has obviously given us much to be thankful
for, but it also created Godzilla and an uncountable number of
giant insects during the 1950s so we should probably treat it
with caution.9 Nevertheless, it was through radioactivity that
humankind was finally able to transmute one element into
another, so we need to get acquainted with it.
The phenomenon was discovered by accident in 1896 by
the French physicist Henri Becquerel. Becquerel had been
planning to do some experiments with photographic plates but
on the day of his tests the sky was overcast, so he put them in
his drawer.
Two days later when he got them back out, the plates had
somehow been impregnated with the image of a copper cross
lying next to them. Apparently, something in the drawer had
taken a photograph. The only other object present was a jar of
potassium uranyl sulfate solution on the other side of the cross
so Becquerel decided it had to be the culprit.
While a photographic plate is best activated by sunlight,
any high-energy beam will cause it to undergo a change.
Potassium and sulfate particles don’t emit beams, so logically
it was coming from the other element in the liquid, uranium.
Invisible to the human eye, the uranium was apparently
emitting something that altered the surface of the photographic
plates. The cross had got in between them and, voilà, the first
radiogram taken by a jar.
Soon after Becquerel’s discovery, Marie Curie, the only
person to win Nobel Prizes in two sciences, named the
phenomenon radioactivity from the Greek radius (wheel
spoke) and the Latin aktinos (ray).
With her husband Pierre, Marie discovered two more
radioactive elements, which she named radium (for obvious
reasons) and polonium after her home country of Poland.
Sadly, both the Curies succumbed to illnesses caused by their
exposure to radioactivity, which taught us something else—it’s
damaging to cells.

INHERENT INSTABILITY

As we learned in Chapter 3, the nucleus of an atom is an


unstable design. While the protons hold electrons in place,
they also repel each other, which requires neutrons to glue
them together.
The Austrian-born scientist Lise Meitner figured out that
once you got up to elements around the high eighties, this
equilibrium would become unstable and the nucleus could fall
apart. For this important discovery, Meitner did not win the
Nobel Prize for physics. Her male lab partner did. But Element
109 was eventually named meitnerium after her, so she hasn’t
been snubbed completely.
As we ascend through the elements, proton numbers
increase so the neutron numbers have to follow suit to keep
things together. But there’s a complication (isn’t there
always?). The repulsive force between protons has an infinite
range but the glue force from the neutrons doesn’t.
This means that in large atoms it’s only a matter of time
before repulsion wins, making them precarious structures.
Larger atoms are fragile and left for long enough will break
apart.
Blue-glowing actinium has a colossal nucleus of eighty-
nine protons so, if you have a lump of it, around half will
decay into something else within twenty years. Rubidium by
contrast is much smaller, with only thirty-seven protons, and
takes forty-nine billion years to decay by the same amount.
The nuclei these elements turn into tend to have peculiar
numbers of neutrons, which the element doesn’t normally
have. These “daughter” particles can only be produced from
radioactive decay, so if you measure the amount of mother and
daughter nuclei in a rock, the ratio between them allows you to
work out how much you had to start with and subsequently
how long it’s been around for.
It was using this technique that the American chemist Clair
Patterson calculated the age of the Earth to be approximately
4.5 billion years old.10

BREAK IT DOWN
There are different ways a nucleus can decay. Sometimes the
whole thing will split in what we call fission, but for reasons
that aren’t understood the most common thing to get ejected
from a nucleus when it fractures is a bundle of two protons
and two neutrons moving at tremendous speed.

These packets come hurtling out of their atoms at 15


million meters per second, and turn out to be the very same
alpha particles Rutherford used in the gold-foil experiments.
When an alpha particle is emitted the nucleus left behind
has lost two protons, changing its identity. Rutherford decided
to use this to his advantage. Given the velocity of alpha
particles, he proposed that if you shot them at another atom
they could shatter its nucleus, turning it into something lighter.
By firing alpha particles through a highly pressurized
container of nitrogen gas to increase the chances of collision,
Rutherford was eventually able to bash nitrogen atoms apart,
turning them into carbon, in 1919. His experiment made
headlines because he had “split the atom” and achieved
transmutation between elements. The long sought-after dream
of the alchemist was not a mythical stone from the Garden of
Eden: it was a gas chamber and an alpha-emitter.11
Turning lead into gold might not be possible through sacred
incantation but, if you take an element like thallium, boil it to
a gas, pressurize it, and fire alpha particles through the sample,
one in every few thousand thallium atoms would be turned
into gold.
BUILD IT UP

Alpha decay makes a certain amount of sense to the human


mind because we can imagine something falling apart when
repulsion overcomes attraction. There’s another thing that can
happen inside the nucleus, though, which can’t be visualized
so easily. Neutrons can turn into protons and spit out an
electron as they do so.
There’s a detailed account of how this happens in
Appendix IV but it would take us way off track at this point.
The best thing to do is cheat and think of a neutron as being a
proton with an electron wrapped around it like a candy
wrapper. If the electron is peeled off and discarded, the
resulting particle will be a proton.
We call these streams of ejected electrons beta radiation
and, unlike alpha decay, which only happens to heavy nuclei,
the neutron/proton transformation can occur in any element.
Some are more susceptible than others (those with more
neutrons) but any atom is potentially beta radioactive.

If we could persuade an element to turn one of its neutrons


into a proton, we could obtain an element one number higher
than that with which we started: Rutherford’s process in
reverse. But first, bananas.

BANANAS

Radioactive particles are charged and move at high speed,


which means they destroy things in their path including the
chemicals of your body.
If you become exposed to enough radioactive beams the
DNA in your cells will fall apart and your body will
disintegrate from the inside out. Usually, the fastest-growing
parts (hair and nails) get affected first, which is why radiation
sickness causes them to fall out. Then all sorts of lovely things
happen like your skin peeling off, your teeth dropping out, and
your innards gradually dissolving into a disordered mush.
In order to monitor the radioactivity to which a person is
exposed, we measure the dosage in units called sieverts. A
sievert is how much energy a radioactive beam is carrying,
compared to the mass of the person it’s entering.
There aren’t clear figures on how many sieverts are
dangerous to a human but roughly five-hundredths of a sievert
per year is when things become problematic.12 The most
radioactive thing you’re ever likely to come across is a dental
X-ray or a mammogram scan, which delivers approximately
0.0004-hundredths of a sievert. A completely safe dose, in
other words.
There is another unit that can be used to measure
radioactive exposure: the banana.
The first thing to say here is that certain nuclei are more
stable than others. By using the Schrödinger equation for
protons and neutrons, we can obtain a list of especially stable
nucleus values, which are genuinely called “magic numbers.”
There isn’t an agreement on why this works—we just know
that certain numbers of protons and neutrons are good and
others are bad.
Potassium is a prime example. Most of the potassium
atoms in the Universe are stable with nineteen protons and
twenty neutrons, but around 0.012 percent have twenty-one
neutrons instead, making them potassium-40, and this
configuration happens to be unstable.
Potassium-40 will undergo beta decay readily so any
sample of potassium will be emitting a very faint trickle of
radioactivity, and the fruit that contains the most potassium is
the humble banana.
Originally created as a joke in 1995 by Gary Mansfield at
the Lawrence Livermore National Laboratory, the Banana
Equivalent Dose (BED for short) calculates the amount of
radioactivity one is likely to experience from eating a single
banana and can be used to calculate the radioactivity of your
food.13
Don’t be alarmed, though: one BED comes to just under
one-millionth of a sievert, so before you boycott bananas let’s
run the math. If we assume five-hundredths of a sievert per
year is lethal, you’d have to consume five thousand bananas
fast enough for it to be dangerous. That’s fourteen bananas a
day. For a year.
If you really want to attempt this experiment, then I suggest
you consult a doctor first. And probably a psychiatrist.

BACK TO PLAYING GOD

In 1940, the American chemist Dale Corson isolated element


85, astatine.14 Predicted by Mendeleev’s table, it was the last
natural element to be discovered.
Slotting it into place gave us a periodic table that went all
the way from 1 to 92 without any gaps. From hydrogen
formed in the big bang to uranium formed in supernovae,
every element was finally identified. But could we go further
and generate our own with higher numbers?
In Iron Man 2 Tony Stark is looking for an element to
power his suit before it kills him from palladium poisoning.
No appropriate metal exists, however, so in order to save the
movie and defeat Mickey Rourke, he manufactures a new
element using a UV-laser and his raw charismatic charm.15
We already know there are no elements missing from the
periodic table so Stark’s unnamed one is going to be made
from huge atoms, and therefore massively radioactive. My
screenplay for Iron Man 3 centered around Tony Stark
vomiting into a hospital bucket for two hours as radioactivity
slowly destroyed his internal organs. For some reason, the
script they eventually chose went a different way. Their loss.
In 1940, Edwin McMillan decided to pre-empt Tony Stark
and make a new element for himself. He took a lump of
uranium and fired a stream of high-energy neutrons at it until
some were absorbed. A uranium nucleus can accept a neutron
but doing so makes it unstable.
In order to lose some energy, one of the neutrons has to
undergo a beta decay, kicking out an electron and converting
itself into a proton. The uranium atom now has an extra proton
in place of a neutron so it’s not really uranium anymore. It’s
element 93.
For fourteen billion years the ninety-third element didn’t
exist in the Universe, and then suddenly, on Earth in 1940, it
did.16
Uranium had been named after the planet Uranus, so
McMillan named his element neptunium after the next planet
in line. Later the same year, as part of the Manhattan Project,
Glenn Seaborg managed to synthesize element 94. This one is
a lot more stable than neptunium so, while neptunium was the
first artificial element, Seaborg’s could actually be made in
chunks big enough to hold. It’s a shiny metal with toxicity
comparable to nerve gas, and Seaborg named it plutonium to
keep with the planetary theme.17
Throughout the remainder of the Second World War and
after, Seaborg went on to synthesize americium (element 95,
named after America), curium (element 96, named after Marie
Curie) and berkelium (element 97, named after Berkeley,
California, where the research was carried out).
These experiments were highly classified as part of the war
effort, but once it was over Seaborg was given permission to
present his findings to the American Chemical Society on
November 16, 1945. However, he accidentally spilled the
beans five days earlier.
An avid popularizer of science, Seaborg was asked to
appear on the children’s radio show Quiz Kids and answer
questions about physics. When an eleven-year-old boy named
Richard Williams asked him if people would ever make new
elements (not realizing he was talking to the world’s leading
expert on that very topic) Seaborg was unable to contain his
excitement and blurted out the classified discoveries live on
air, much to the annoyance of his superiors.18
Really though, can we blame him? He was surrounded by
eager minds asking him everything he knew about his favorite
topic. One might say he was in his element. I’ve been waiting
for eight chapters to drop that joke.

COMPLETING THE TABLE

The periodic table is split into seven periods representing the


seven orbital shells, and eighteen groups representing how
many electrons occupy each one. As a result, the table has 118
spaces. With 92 occurring naturally, that gives us 26 blanks to
fill.
Seaborg got lucky with his elements because they were all
reasonably stable. Had he kept going he would have found
things got a lot more difficult. Forcing nuclei to put on weight
isn’t easy because the larger they get, the more repulsion there
is between protons.
The best approach is to take samples of an already large
element and bombard it with smaller nuclei in the hope that
they get absorbed. In 1950, californium was made by firing
alpha particles at curium, and einsteinium and fermium were
made in 1952 via a similar route.
We’ve also used this technique to create lower-numbered
elements, which are normally rare in nature. Francium is the
second scarcest element on the table (behind astatine) with
approximately 30 g available in the Earth’s crust. But if we
fire an oxygen atom at a piece of gold we can generate it.
We can also create supplies of technetium, element 43,
which has an unstable nucleus and doesn’t normally last. It’s
worth doing because it makes up 80 percent of the world’s
medical tracers, injected into the body to track blood flow.
Making artificial elements is a precision operation, of
course. Fire the nuclei too slow and they bounce off; go too
fast and everything shatters. But over the last half century we
have edged closer and closer to a full table.
We’ve fired carbon atoms at americium and curium to
make mendelevium and nobelium. We’ve fired neon at
einsteinium and made lawrencium. We’ve fired neon again at
plutonium and made rutherfordium.
By the early 2000s, we had dubnium, seaborgium,
bohrium, hassium, meitnerium, darmstadtium, roentgenium,
copernicium, flerovium, and livermorium, leaving only four to
be discovered. The missing numbers were 113, 115, 117, and
118.
At this point, the bottom-right of the periodic table looked
like a row of punched teeth. Then, in November 2016, the
International Union of Pure and Applied Chemistry announced
the successful synthesis of nihonium, moscovium, tennessine,
and finally oganesson. The periodic table was finally
complete.19
Some of this might seem like pointless playing around but
many of these artificial elements can be useful. You probably
have a sample of americium, element 95, in your home right
now. At least, I hope you do.
Americium emits alpha particles constantly so if you put it
in an open circuit, the charged particles can fly across a gap to
a receiver and complete the circuit without wires. When flecks
of smoke or dust float into this gap, the alpha stream gets
blocked and an alarm triggers. This is how your smoke-
detector works.

THE END OF EVERYTHING

Now that we’ve got all the way to element 118 and completed
the table, could we go further? The honest answer is that we
aren’t sure. Oganesson represents the filled seventh shell, but
there might be an eighth or even a ninth shell.
Seaborg suspected the periodic table might stop when we
reach element 126 because it’s a magic number and beyond
that the proton-repulsions may become too powerful, no
matter how many neutrons we include. It has even been called
unbihexium as a placeholder name.20
Other physicists speculate that we could go on to create a
ninth period or a tenth and an eleventh without limit. We don’t
know enough about the nucleus to say for sure, so the only
sensible thing to do is try. And that’s the whole point of
science: to see what might be possible.
CHAPTER NINE

Leftists
THE EASIEST NOBEL PRIZE EVER EARNED

Twenty years ago, if you’d asked a scientist what the most


electrically conducting element was, they would have said
silver. The only reason we don’t use it in electronics is that
copper is cheaper.
Then in 2004 two physicists won a Nobel Prize by making
another element conduct better with a piece of Scotch tape.
Russian physicists Kostya Novoselov and Andre Geim
(who you might remember as the scientist who made frogs
levitate in 19971) were working with graphite, the soft form of
carbon used to make pencil cores. Because graphite is a brittle
material it tends to become flaky and the scientists down the
hall were using Scotch tape to clean their samples. By sticking
tape to the graphite and peeling off the excess dust, the result
was a shiny new surface.2 It was watching this that gave
Novoselov and Geim an idea.
If you stick the tape to a lump of already cleaned graphite
you can extract a single layer of carbon, no more than one
atom thick. This peculiar substance, which they named
graphene, is arranged like a chicken-wire fence of carbon
atoms and it has many unusual properties. Not only is it two
hundred times stronger than steel, it is also transparent and can
be used as a sieve to filter the salt out of seawater.3
On top of all this, graphene has an electrical conductivity
better than silver. We measure electrical conductivity in units
called siemens (pronounced zeemuns) per meter. Silver has a
conductivity of 60 million siemens per meter and graphene
clocks in even faster, although nobody has been able to agree
on a definitive reading yet.4 What makes this surprising is that
carbon isn’t a metal and it’s usually only metals that conduct.
Something very weird is going on.

WHAT IS A METAL?

When we hear the word metal we all picture the same thing:
Ian “Lemmy” Kilmister, the bassist/vocalist of English rock
band Motörhead. May he rest in peace.
After that we tend to think of grayish solids that are hard
and shiny. What we’re really thinking of when we do so are
steel, titanium, aluminum, and chromium, the four metals that
dominate our everyday experience, but metals have all sorts of
other appearances and properties.
Bismuth forms labyrinthine square crystals, which glisten
like oil on a puddle, while lutetium and thulium are found in
fibrous clumps that look like pieces of torn beef. Niobium is a
dull silver when first isolated but pass an electric current
through it and it becomes rainbow-colored.
Some metals show magnetism (iron, cobalt, nickel,
terbium, and gadolinium) while some are not magnetic
themselves, but will reinforce the property in those five
(neodymium). Some metals will remain solid when heated to
over 3,000°C (tungsten) while others will melt in the palm of
your hand (gallium). Their reactivity also ranges from gold,
which won’t even corrode in acid, to erbium, which explodes
if you warm it gently.
With such a broad spectrum of behavior, what is it that
unites them all? The answer is that a metal is an element that
will always conduct electricity. Sure, carbon will conduct in
the graphene state but metals will conduct no matter what state
they’re in.
In order to understand metal chemistry, we need to
understand electricity and that story starts in ancient Egypt.

THE FIRST PHARAOH

In 3100 BCE the kingdom of Egypt was united for the first time
under the rule of Narmer, the original pharaoh. There’s a lot of
debate around Narmer’s true identity but we know the
meaning of his name with some confidence. Narmer,
translated into English, means “angry catfish.”5
It may seem odd that a pharaoh would adopt the name of a
river fish, but in Egyptian culture catfishes were the lord-
protectors of the Nile and one of the most revered creatures in
the world.
It’s true that most catfish are useless monstrosities but the
breed found in Egypt is special. Its Latin name is
Malapterurus electricus, which means “electric catfish.”
Like the electric eel of South America, this creature harbors
a special organ that gives it the ability to deliver 400-volt
shocks to anyone touching its skin. Records of the electric
catfish are the earliest examples we have of electricity and it
was five thousand years before humans could boast a similar
control of the phenomenon.

SHOCKING

It is a crying tragedy that the man who discovered electricity is


usually forgotten. The Greek scientist Thales (the one who fell
down the pit) had already made the discovery that rubbing
pieces of amber with wool caused them to gain a crackly
property, which sparked under the right circumstances, but the
discovery of what we think of as electric current goes to an
English experimenter named Stephen Gray.
One of the reasons why Gray’s work was overlooked is that
he made the mistake of asking another scientist to help him
develop it. That scientist was John Flamsteed, who happened
to be a mortal enemy of Sir Isaac Newton.
Newton was a socially cruel, even malicious character who
used his position as head of the Royal Society to discredit and
bury the work of people he disliked, including Flamsteed.6
Consequently, much of Flamsteed and Gray’s achievements
were ignored. It has to be said that while Newton was one of
the greatest minds in history, he was also a jackass sometimes.
So, let’s redress the balance and give Stephen Gray his due.
Born in 1666, Gray worked as a dyer for most of his life
and only indulged science as a hobby. He discovered
electricity one night in his bedroom at the age of forty-two
while playing with a crude instrument used to generate static
—a tube of glass.
Static generators had been around since 1661, invented by
the German politician Otto von Guericke, but Gray didn’t have
the money for such lavish equipment. He had to make do with
rubbing a glass rod on rabbit fur and tapping it on whatever
was around in the hopes of creating a shock.
Gray was curious about the fact that if you put the rod on
the ground after rubbing it, it seemed to lose its electricity and
wouldn’t shock anything again until recharged.
On this particular night he decided to jam the end of the rod
into a piece of cork and discovered that when he tapped the
cork-tip against a pile of feathers, it sparked. The glass had
been rubbed but the cork was somehow able to transfer the
electricity through itself. Whatever electricity was, it could
flow.
Excited by this result, Gray built a silk harness from his
ceiling so objects wouldn’t touch the ground, and began
testing things to see if they would transfer electricity. After
trying vegetables, string, coins, and anything else he could
find, Gray began dividing everything into two categories:
insulators, which wouldn’t transfer electricity, and conductors,
which would.7
The best conductors turned out to be metals, located on the
left side of the periodic table. These were so good at electric
transfer that Gray was able to pass a shock down nearly 250
meters of wire suspended from his bedroom window.8
Metals even conducted when pointing upward, which
meant that whatever electricity was, it wasn’t influenced by
gravity. Electricity would still go into the ground, of course,
but it’s obviously not because of gravitational attraction.
Instead, the planet itself was a conductor, which electricity
will flow through given the chance.
Even more surprising among Gray’s results was the fact
that humans conducted electricity. By suspending a young boy
from his silk harness, Gray was able to charge him and
generate sparks from his face. This became the basis of a
popular sideshow exhibit called “The Flying Boy” in which
spectators could tap the floating youngster’s fingertips and
receive a shock.9 All in the name of science.
The secret to this exhibition is that human skin is usually
coated in a fine layer of saltwater in the form of sweat,
allowing electricity to zap across its surface. When the
spectators, who were connected to the ground, touched the
charged boy, the electricity would flow over their skin and into
the earth, creating the shock effect.
We know from Chapter 3 that electricity is made from
electrons, so to explain all these behaviors we must turn once
again to the Schrödinger equation.

STATIC

As we know, electrons occupy orbitals around their nucleus


and atoms brought together can mix orbitals to form
molecules.
Static electricity happens because this orbital mixing is not
a rare occurrence. In fact, it happens when any two surfaces
meet. As you sit on your chair right now, a few of the chair’s
electrons are forming temporary bonds with the electrons in
your clothes (at least I hope so; please don’t read my book
naked).
When you stand up, most of the electrons return to their
original atoms and the bonds are severed. Chair electrons go
back to the chair and clothes electrons return to you. We refer
to this as the triboelectric effect and it’s a weak form of
chemical bonding.
The thing is that some molecules are better at holding
electrons than others and when it’s time for the bonds to break
they don’t always return to their original configuration.
The molecules that make up human hair, for instance, are
poor electron-holders whereas rubber is very good at it. If you
put a piece of rubber such as balloon against your hair some of
your hair electrons realize they’re happier sticking with the
rubber and they transfer across.
There’s no limit to how many electrons you can cram onto
a molecule so the rubber is happy to accept these travelers.
When you separate from the rubber, some of your hair
electrons stay on the surface of their new home and a charge
imbalance arises.
The rubber and hair originally had no overall charge
because the electrons and protons canceled each other but, if
we transfer electrons from hair to rubber, things look different.
The balloon finds itself holding an electron surplus while your
hair has an electron deficit.
The surprising thing is that transferring electrons this way
leads to greater stability. It sounds wrong because the rubber
has stolen something from your hair, but remember that
stability in quantum terms means “things have already lost
energy to get to this state.” Two molecules can be a lot more
stable if they split, the same way a house of cards is a lot
happier falling to pieces.
The overall result is that when you rub a balloon on your
hair it steals somewhere in the region of two hundred billion
electrons. That sounds like a lot but it’s less than a trillionth of
a percent of the electrons your body has in total.
If the balloon is now brought near a good conductor (like a
piece of metal or the ground), the electrons are offered an even
better deal and will flow into it, spreading as far away from
each other as they can. Except this time we’re not talking
about little bonds being formed, we’re talking about all the
electrons jumping at the same time, creating the infamous
static shock.
When Stephen Gray rubbed the glass rod, he was
depositing electrons on its surface from the rabbit fur. Glass is
an insulator so it can store electrons on its surface and won’t
allow them to flow from one end to another. The cork was a
conductor, however, so the electrons were able to travel
through it and into the ground. His experiments with wires
were an extension of the same principle.

WHY DO METALS CONDUCT AT ALL?

As you read from left to right across the periodic table you’re
gradually increasing the number of protons in the nucleus. The
more proton charge you have, the more electrons will be
pulled inward and the smaller your atom becomes, meaning
we see a decrease in atom size along each row.
Atoms on the left are therefore big and diffuse with great,
floppy orbitals. Their electrons are also a long way from the
nucleus with nothing much keeping them in place. This makes
them ideal for sharing electrons with other atoms since the
electrons have very little incentive to stay put.
When you get these bulky atoms together, their orbitals
start mixing not just on a one-to-one basis but over the entire
population. The atoms are so happy to share that when you
solve the Schrödinger equation to describe millions of metal
atoms, the result is a kind of mega-orbital—a turbulent free-
for-all, which physicists call “the electron sea.” This network
of overlapping orbitals means electrons can easily slosh from
one side of the structure to the other.
Touch any piece of metal and beneath your fingertips
you’ve got a swarm of electrons flitting back and forth at will.
These movements are random but if we can persuade the
electrons to travel in one direction at the same time we have an
electric current.
In smaller molecules, formed by elements on the right,
gaps between the orbitals make it hard for electrons to move,
so they won’t conduct. That doesn’t mean, of course, that it’s
impossible to force an electron through an insulator. Teflon,
the most insulating material on Earth, can still be made to
conduct but you need a fierce amount of energy to persuade
the electrons to hop across the orbital gaps.
A substance with a conductance over 1 million siemens per
meter is classified as a conductor while a substance below 0.01
is an insulator. Admittedly there’s a huge gap between 0.01
and 1 million siemens per meter, but very few substances fall
in this region. Those that do are deemed “semi-conductors.”

THE WEIRDO
Whether a substance is a solid, liquid, or gas depends on how
much the particles are attracted to each other. Oxygen
molecules have little interaction because they’re stable,
making oxygen a gas at room temperature. It can be turned
into a liquid by cooling it down (fun fact: liquid oxygen is
blue) but under standard conditions it tends to spread out.
By contrast, metals are good electron sharers, meaning
their orbitals overlap and they clump together forming a solid,
with the obvious exception of mercury, the metallic liquid. A
full explanation for mercury’s liquidity requires knowledge of
Einstein’s theory of special relativity, but we can get the gist
without worrying about that.
Like other metals, mercury’s orbitals stick out in many
directions like petals on a flower so it can conduct, but it’s in a
funny position on the table. It sits on the bottom row, making
it huge, but over on the right-hand side meaning it has a lot of
protons pulling the orbitals inward. The result is that the
orbitals are extended enough to overlap but not quite enough
to hold the atoms together.
Move to the right and you increase the proton number,
causing the atoms to pack together better, resulting in a solid.
Move to the left and the orbitals overlap better, also resulting
in a solid.
Mercury atoms are just too weakly attracted to stick
together, but just attracted enough to allow electrons to hop
from atom to atom. The result is that mercury is a conducting
element and therefore a metal but it’s unquestionably the worst
metal on the table.

THE OTHER WEIRDO

When electrons travel through a piece of metal they don’t


move in perfect lines. The nuclei vibrate and the inner orbitals
interfere with the outer ones. The result is that conductivity
never happens perfectly and we call the collection of things
that slow it down “resistance,” measured in ohms. The energy
electrons are given as they are pushed through the metal is
called voltage (measured in volts). These things together give
rise to the overall electron flow.
If we think of voltage as a fist squeezing the end of a
toothpaste tube, the resistance is the diameter of the tube and
the actual amount of toothpaste that comes splurging out is
what we call current, measured in amperes (amps for short).
A watch battery delivers electrons into the watch with an
energy of about 1.5 volts. The resistance of the circuit slows it
down and we end up with a current in the region of five-
millionths of an amp (0.000005 A).
For perspective, a bolt of lightning packs around 100
million volts. This electricity is forced through the air,
however, and the overall current ends up at around 5,000 amps
by the time it reaches the ground. Passing electricity through
non-metals like air involves a lot of energy being lost.
Graphene’s conductivity is therefore very strange. Carbon
is a non-metal most of the time but when it is arranged in the
thin wafers of graphene it starts to conduct.
It happens because the atoms in graphene are arranged in
flat hexagons with each atom bonded to three others. Since
carbon has an available four electrons in its outer orbitals, each
atom has a spare one that isn’t involved in bonding. This
electron can move from atom to atom with hardly any
obstruction, so even a small voltage will produce a lot of
current.
Where graphene differs from metals is that it is almost two-
dimensional. In a metal, electrons can change route and go
exploring in all directions but in graphene there are less places
to go. It is practically a flat plane, meaning electrons have no
possibility of moving up or down, making them more likely to
stay on track.

ELECTRICITY AND YOU

In 1886 an American human rights committee decided that


execution of criminals by hanging was inhumane and a new
method of capital punishment was needed. One of the people
on the committee was Alfred Southwick, a dentist from New
York who had already designed an electrocuting chair several
years previously. Southwick’s idea was approved and testing
began, endorsed by none other than Thomas Edison himself.10
At the time, there was a battle going on over which type of
electricity the United States should adopt. Edison had put a lot
of money into battery-based electricity and needed to find a
way of tarnishing the reputation of magnetically generated
electricity, favored by his rival George Westinghouse. His
solution was simple, if a little gruesome.
In the most morbid marketing strategy ever employed,
Edison insisted that the newly designed electric chair be
configured to run on Westinghouse’s electricity, so people
would associate it with death.
He tested the chair on stray animals in his workshop and is
on record as having killed dogs, cats, birds, a horse, and a
circus elephant named Topsy (he was considerate enough to
film that last one, and you can watch it for free online if you’re
into that sort of thing).11
Soon after, the electric chair was rigged-up for its first
victim, William Kemmler, in 1890.12 Kemmler took over four
minutes of continual electrocution to die, with the procedure
stopping halfway through until someone screamed, “Great
God, he is alive!”13 Humane indeed.
The key to the electric chair is making sure the human
body is part of a circuit, which is actually quite difficult to do.
Despite what Saturday morning cartoons claim, you’re not
very easy to electrocute.
If you’re ever unfortunate enough to be hanging from a
power line you may feel a tingling in your fingers, but you’re
in no real danger. Once the electricity has filled all the
available orbitals on your surface there is nothing else it can
do.
If, on the other hand, you somehow connect to the ground
then you’re not a cul-de-sac anymore: you’re a pathway and
the electricity will use it. If electricity goes onto you, you’re
fine, but if electricity goes through you, you’re in trouble.
The human body is a fairly decent conductor (you’re a bag
of salty water) but to complicate matters your skin is an
excellent insulator. Dry skin has a resistance of about 100,000
ohms, although wet skin absorbs water into its pores and the
resistance drops to around 1,000 ohms.
It’s also worth pointing out that once electricity enters your
body it will travel along the easiest path available. A tiny
amount may go exploring, but you could pass thousands of
amps through your hand without dying. It would still hurt, so
don’t do it, but you wouldn’t be in mortal danger.
The only time electricity becomes lethal is if it passes
through your heart, lungs, or brain for a sustained period of
time.
The way your heart works is that the muscular outer layer
is given a short electric shock of around 0.0000012 amps
every second, which causes it to contract, squeezing blood to
your body. Afterward it is allowed to relax and reopen, taking
in more blood, before the whole thing is repeated.
If a current is pushed through the heart for a long time,
however, it squeezes tight and doesn’t reopen, meaning it can’t
take in a fresh load of blood. That’s why people can survive
lightning strikes but not the electric chair. The electricity of
lightning may pass through your heart, but it does so for a
short time only and your heart is able to return to normal. If
you keep the current flowing, you essentially give the person
an artificial heart attack.
Surprisingly (or perhaps not) very little research has been
done on how much current is needed to make the heart do this.
The approximate guidance, based largely on anecdotal
evidence and a bit of bioelectrical theory, suggests that around
0.05 amps is required to kill a person.
The electric chair worked by passing a current of between 1
and 7 amps through the body depending on the state
legislation. That’s over twenty times the lethal dosage.
Typically, the two live ends of the circuit would be
connected to the scalp and ankle so the current would pass
through the brain, heart, and lungs together, guaranteeing the
malfunction of at least one of them and ensuring a warm
death. Have a nice day.
CHAPTER TEN

Acids, Crystals, and Light


A BARREL OF HORRORS

In March 1949, English newspapers reported one of the most


gruesome crimes to occur in British history since those of Jack
the Ripper. John George Haigh, who the Daily Mirror had
referred to as the Vampire Murderer on March 3, was taken to
court and charged with six counts of pre-meditated homicide.
What made them particularly ooky wasn’t the murders
themselves but the way he disposed of the bodies.
After drinking glasses of their blood, Haigh loaded each
body into a 40-gallon drum, which he topped with
concentrated sulfuric acid and left for two days. The remaining
sludge was poured into a drain behind his workshop, earning
him his other colorful nickname, the Acid Bath Murderer.
Acids capture people’s imagination because they are the
standard “nasty” chemicals, capable of chewing through a
human body and destroying any evidence the person was ever
there. The only reason Haigh was caught was because the
solution of his final victim, Olive Durand-Deacon, still
contained part of her plastic denture, which her dentist
identified.
Haigh was executed by hanging on August 10, 1949. He
claimed to have other victims but they are unidentified to this
day because the bodies were disposed of so perfectly.1

IT BURNS, IT BURNS!

An acid is a substance whose molecules fall apart in water to


produce free-floating protons. Protons are the charged particles
inside a nucleus, shielded most of the time by their electron
orbitals, but if they get released when their parent molecule
dissolves they can cause untold damage.
A rogue proton is a concentrated lump of charge and will
pull electrons toward itself at any cost. Things like glass or
plastic have strong bonds between atoms so acids aren’t
usually able to react with them, but any chemical with loose
bonds, including the ones in your body, will be pulled apart.
An acid can be thought of as a proton juice and the easiest
way to generate a solution of protons is to make sure your
starting molecule contains hydrogen. Hydrogen is the simplest
element, consisting of one proton and one electron, so if its
electron is more interested in the other atoms of the parent
molecule, the proton will drift away.
Take hydrogen chloride. Each molecule consists of one H
atom and one Cl atom, giving it the formula HCl. The chlorine
atom is very good at holding electrons, better than the
hydrogen, so the bond between them isn’t a fifty/fifty share—
it’s lopsided like so:

Put the whole thing in water and the two atoms will
separate with chlorine keeping all the electrons and hydrogen
being left essentially naked.
This lonely hydrogen proton drifts away, waiting until
some other molecule with which it can react comes along. We
have generated hydrochloric acid, the one in your stomach,
capable of dissolving bone.

THE STRONGEST ACID

We measure how strong an acid is by how willing it is to let go


of a proton. The numbers involved spread over an enormous
range and we use something called the pKa scale to measure it.
The scale works the same way as the earthquake Richter scale
where each number is ten times greater than the one before it.
The scale also works backward for reasons we don’t need to
worry about (see Appendix V if you’re curious). So, the lower
the number, the stronger the acid.
Household vinegar has a pKa of 5 whereas oxalic acid, the
one in rhubarb, is closer to 4, making it ten times more potent.
Then there’s chromic acid, a powerful industrial agent, with a
pKa of 1—three places lower down the scale than oxalic and
therefore one thousand times stronger. For context, you can eat
oxalic acid and feel fine but chromic acid will set fire to living
tissue.
The concentrated sulfuric acid Haigh used to dispose of his
victims scores a −3 on the pKa scale, seven numbers lower
than vinegar and therefore ten million times stronger.2 That’s
another way of saying that sulfuric acid is ten million times
better at releasing its proton than vinegar. But if we can create
molecules with absolutely no interest in holding their
hydrogen, we end up with a class of chemicals on our hands
called superacids (well, hopefully not on our actual hands).
Perchloric acid has a pKa of −10, which is ten million times
stronger than concentrated sulfuric and triflic acid has a pKa of
−14, one hundred billion times stronger.3 And that’s not even
touching on magic acid (actual name), which will dissolve
even candle wax.4
If you scan around the internet, most popular science
websites tend to report the strongest acid in the world as
something called fluoroantimonic acid, boasting a pKa of −19.
That’s ten quadrillion times stronger than sulfuric. It’s
occasionally used in the electronics industry to etch
equipment, but it doesn’t really deserve the gold medal. That
goes to an acid so strong it has only been synthesized once in
recorded history.5
An acid’s job is to kick hydrogen away so the best acid will
be one where the other atoms don’t want to bond with it in the
first place. And there’s no better atom for that than helium, the
least reactive element on the table. If you can force hydrogen
to bind with helium, you’ve created the weakest bond it’s
possible to get and it will fall apart instantly.
In 1925, the chemist Thorfin Hogness managed
successfully to brew a microscopic quantity of helium hydride
that possesses a pKa of, brace yourself, −69.6 That’s so strong
there isn’t even a word to describe how much better it is than
sulfuric acid.
The non-reactivity of helium is also responsible for another
record-breaking property it possesses: liquid helium is the
most fluid liquid in the Universe. When a sample of helium is
cooled to around –269°C, the atoms lose their movement
energy and settle into liquid form. In most liquids the atoms
still interact with each other a little, but in helium they keep to
themselves.
If you take a cup of liquid helium and stir it once, it will
keep spinning forever. Any other liquid would interact with
the container and be slowed down but liquid helium doesn’t
feel friction and will keep spinning until the end of time.7
Wouldn’t that constitute a perpetual motion machine,
though? The answer is that if we tried to put something like a
propeller into the swirling vortex, the helium would just flow
around it. The only way to get the liquid helium to work on
something would be to warm it up, and as soon as you do that
the superfluidity is lost.
Liquid helium also happens to defy gravity. Air pushes
down on everything at atmospheric level and near the edges of
a container some liquids are able to creep up the sides because
they’re being pushed by air on one side but not the other.
Most liquids are self-attracting enough to stay together and
not begin climbing the walls but liquid helium isn’t most
liquids. Helium will move up the sides of an open container
and creep its way out, emptying the vessel as if it had a desire
to escape.
In order to understand the surreal properties of liquid
helium and helium hydride, we’re going to travel to the right
side of the periodic table. The realm of the non-metals.
SELFISH CREATURES

Most of the spectacular and violent reactions in chemistry take


place in the non-metals because they’re so greedy. As we’ve
already seen, metals are big and have friendly, overlapping
orbitals, but atoms on the right are small and grip their
electrons tightly.
The most reactive element is fluorine, which we met in
Chapter 1 when it was setting fire to water. A sparse yellow
gas, fluorine needs to be transported in dense steel and
bulletproof glass because it will rip the electrons out of
anything else it touches.
Because it’s so electron-hungry, a molecule of two fluorine
atoms will be perfectly symmetrical as the electrons are shared
between them. If you bond it with a metal such as cesium,
however, the bond is uneven, with fluorine getting the lion’s
share of electron density. It’s similar to the way hydrochloric
acid molecules are arranged—non-metals always win because
they don’t like to share.
This electronic exchange means cesium atoms become
electron deficient while the fluorine atoms become electron
rich. It’s not really correct to call them atoms anymore since
they aren’t neutral units so we refer to them as “ions” instead.
Ions are still sharing electrons but it’s such an uneven bond
we usually just imagine cesium losing electrons and fluorine
gaining them.
You’ll see diagrams of ionic bonds in which the particles
are drawn like balls packed together, such as in the diagram on
page 122 (top). That’s not strictly correct, but it helps keep
track of where the ions are and how they’re arranged. The
diagram at the bottom gives a slightly more accurate picture.
This type of bonding where things are arranged in a lattice
framework gives rise to crystal properties. Starting at group 13
with boron, non-metals tend to keep their electrons in very
specific places, forming grids with sharp edges. This brings us
to …

SPARKLY CREATURES

Boron is the second-hardest element after carbon. Used in


making glues and glass, it’s usually found bonded to oxygen
and sodium in the form of borax crystals exported from Death
Valley, California, the hottest place on earth.
Borax crystals have a ghostly white appearance and are
mostly transparent, something you don’t get with metals.
Because of the sea of electrons, light will bounce off the
surface of a metal, making it opaque. Non-metals, on the other
hand, hold their electrons in fixed orbitals with gaps, meaning
beams of light travel through rather than getting reflected.
Depending on the angles between ions and the sizes of their
orbitals, a beam of light can emerge from a non-metal looking
very different to when it entered. As the light is bounced
around inside the crystal matrix, it can lose or gain energy,
changing color and giving the crystal a different appearance.
The most common crystals on Earth are based on silicon
and oxygen in the form of SiO2. It’s the other elements mixed
with them that give rise to the different minerals we find in the
ground. A single hunk of rock (a conglomerate of mineral
crystals packed together) can contain dozens of different
elements and we have to extract them with acids or electricity.
In fact, most elements on the table were discovered by
grinding up rocks and seeing what was inside them. The
elements yttrium, ytterbium, erbium, and terbium, for instance,
were all discovered at the same Swedish mine from a single
type of rock.
The most prized crystals, though, tend to be based on
oxygen bonded to aluminum rather than silicon. On its own,
aluminum oxide is a white crystal called corundum with an
appearance similar to table salt. But if a few chromium atoms
get mixed in, you’ve got ruby. Replace the chromium with
titanium or iron and you get sapphire.
Then the most precious crystals of all, diamonds, are made
from carbon atoms forming a tetrahedral array, with each atom
linked to four around it. And again, it’s the impurities that give
the colors. A bit of boron and your diamond turns blue while a
touch of nitrogen will give you yellow. Change the atoms and
you change the color.

HIGHFALUTIN ELEMENTS

As we go across any row on the periodic table we’re dealing


with atoms that house more and more protons. The electron
orbitals get sucked in and, as a result, everything on the right
is smaller and greedier.
Group 17 is where we get things like fluorine (sets fire to
cotton wool), chlorine (a chemical weapon), and bromine (a
toxic disinfectant). But when we get to group 18 something
strange happens. The elements of this column—helium, neon,
argon, krypton, xenon, and radon—are the least reactive on the
table.
They are so reluctant to get involved with bonding that they
were originally named inert gases. We have since learned that
group 18 elements will mingle with others a little, but not if
they can help it. As a result, these snooty substances are
referred to as “noble” gases (other groups have names too, see
Appendix VI).
We saw earlier that helium’s refusal to bond is what makes
helium hydride the strongest acid in the world, so the obvious
question is what are the most lifeless elements doing next to
ones like fluorine and chlorine?
The answer comes from how electrons are distributed
around the nucleus. Orbitals are fixed in certain shapes
according to the quantum rulebook but they are also grouped
at specific distances.
The first set of orbitals are huddled around the nucleus, but
the second set are a great distance away. This outer set is
repelled by the inner set and there is a no-man’s-land between
them.
The diagram opposite shows the energy levels of the first
and second orbital sets. For simplicity, we’re ignoring the
orbital shapes because then our diagram would look like a
plate of panda entrails.
We call these orbital groups “shells” and they are the
reason for the periodic trend Newlands identified. As we go
from one side of a row to the other, we are filling the orbitals
of a particular shell. When a shell is full, we jump to a higher
one and start filling that instead, starting a new line on the
table.
The noble gases are the elements we get when we have
completely filled a shell. Because every orbital of these atoms
is full, there’s nowhere to put an incoming electron. The atoms
are also small (they’re on the right-hand side), which means
they hold their own electrons tightly and won’t donate to
anything else.
Noble gases are therefore unlikely to accept electrons or
donate them, making them bad at bonding. A few dozen
noble-gas compounds have been created over the last few
decades but it’s not a common occurrence.
It might seem that these elements are pointless and boring,
but their refusal to react makes them useful. Take a light bulb.
The filament inside is made from tungsten, which glows when
electrified. The problem is that the tungsten gets so hot it
would begin reacting with oxygen. To avoid this problem, we
flood lightbulbs with argon instead of air so nothing reacts and
the bulb can continue to do its thing.
We can also use noble gases to produce vibrant colors of
their own. If you trap a sample of noble gas inside a glass tube
and pass a current from one end to the other, the atoms will
start vibrating. The electrons get pushed outward by the
electrical energy, but they fall back immediately, kicking out
specific beams of light.
Any other gases would start reacting and everything would
rearrange to become stable. Since stable means no more
energy is available, your light would switch off moments after
you switched it on. But noble gases are so reluctant to bond
they just keep jumping back and forth, emitting a steady
stream of light. Neon makes the tube glow red, helium glows
orange, argon glows blue, krypton glows green, and xenon
glows turquoise. Neon was the first to be discovered so we
refer to these harsh buzzing tubes of gas as “neon lights,” the
kind you see outside store windows.
CHAPTER ELEVEN

It’s Alive, It’s Alive!


THE MOST TOXIC POISON

In 2006, the world’s media reported on the agonizing death of


Alexander Litvinenko as he succumbed to polonium
poisoning. What made the story so chilling, aside from the
political overtones, was the minuscule amount of polonium
needed to cause death. It was estimated that Litvinenko
consumed less than one-hundredth of a gram and was dead
within three weeks.1 Is polonium the worst thing you can have
in your body?
Judging toxicity is not as straightforward as you might
imagine. For starters, everyone metabolizes things differently.
Nicotine alone turns into seven different chemicals depending
on the person, which might explain why some people find it
harder to quit smoking. They literally turn it into more
addictive substances.
This means that if you poison a large group of people some
will die and some will survive, purely by chance. In order to
get around this, biologists use something called the LD50
value, the lethal dose guaranteed to kill 50 percent of a group.
The number is given in mg/kg (how many milligrams needed
to kill every kilogram of creature) and the lower the LD50, the
more toxic the substance.
The LD50 of pure caffeine is 367 mg/kg.2 A baby duck,
which typically weighs around 1 kg, could therefore ingest
367 mg of caffeine and have a 50 percent chance of living. An
African bull elephant, on the other hand, weighs 5000 kg so
you’d need about 2 kg of caffeine to be 50 percent confident of
killing it.
It’s also difficult to quote accurate LD50 values for humans
because the only way of obtaining them would be to poison a
bunch of people and see how many died. Sadly, there have
been cases of experimentation on unwitting suspects, but
usually such studies aren’t common.3
Some animals can be seen as close approximations to
humans but you run into the same problems. Different species
metabolize things differently. Glucuronic acid is harmless to
humans and used in cooking sauces but it’s lethal to a cat.
Arsenic is toxic to us but when added to chicken feed it causes
them to gain muscle mass. Plus, there’s the well-known fact
that theobromine in chocolate can kill a small dog, but all it
does to humans is leave them with a sense of self-loathing.
The animals biologically closest to us aside from
chimpanzees, which are not tested on, are rats. Whatever your
ethical stance on animal testing, the fact remains that trialing a
chemical on rats is the closest thing we can get to human data.
It’s also worth remembering that chemicals get processed
differently depending on how they are absorbed. Some
elements like holmium are toxic no matter how you take them,
but something like indium is only dangerous if inhaled. (NB:
probably best you don’t ingest either.)
All of these factors make it very hard to say what is the
most poisonous chemical in the world. That’s probably a good
thing, but since we’re on the subject we might as well look at
some of the candidates.
Lead has an LD50 of 600 mg/kg while thallium has one of
32 mg/kg, making it twenty times more dangerous. Arsenic,
the preferred poison of nineteenth-century novelists, has an
LD50 of 20 mg/kg while phosphorus comes in at close to 3
mg/kg.4
If we go on toxicity alone, this makes phosphorus the most
poisonous element, but if we include the effects of
radioactivity polonium outranks it by a clear mile. Radioactive
elements don’t just kill by interfering with the functioning of
the body: they spit alpha particles (see Chapter 8), which
essentially rip your cells apart.
Because of this additional mode of killing, polonium
probably is the deadliest element. In fact, nobody knows what
the LD50 of polonium is because experimenters are reluctant to
work with it. Even its dust can kill you. But given the amount
needed to kill Litvinenko, the LD50 is going to be very small.
If we start including compounds as well as elements,
though, polonium is no longer that bad. Dimethylcadmium is
often cited as the most toxic compound in the world, so toxic
that a thousandth of a gram dissolved in a ton of water is
lethal.5 But the crown really belongs to botulinum toxin, a
chemical produced by the bacteria Clostridium botulinum.
There are several varieties given the names A to H, and it’s
botulinum toxin H that is the worst. Only two-billionths of a
gram are needed to kill a fully grown adult.6 Assuming the
population of Earth is around seven billion, you would
therefore need only 14 g (a teaspoon’s worth) to wipe out the
entire species. And it kills you in a pretty nasty way,
paralyzing you to death.
You can also dilute it down to low concentration and inject
it into your forehead, paralyzing the muscles and preventing
wrinkles. Botulinum toxin A (not quite as deadly) is used for
exactly this purpose and is marketed under the trade-name of
Botox®.7

THE ELEMENTS OF LIFE

In 1924, the head of the American Medical Association,


Charles Mayo, published a tongue-in-cheek calculation
showing that if you split a human body into piles of its
constituent elements the total value would be around eighty-
four cents.8 The iron from your blood would make a single
household nail while the carbon in your proteins would make a
small bag of charcoal, etc.
We did something similar in the introduction when we
looked at the chemical formula for a person. It’s a powerful
reminder that the atoms that make up our bodies are no
different to the atoms that make up the contents of our kitchen.
A lot of people seem uncomfortable with this notion. I once
saw a magazine ad in which a worried customer is reassured
by a scientist that their ice cream “contains no 4-hydroxy-3-
methoxybenzaldehyde, only natural vanillin.” What the writers
of the advertisement didn’t seem to realize is that 4-hydroxy-
3-methoxybenzaldehyde is just the chemical name for vanillin.
It would be like saying “this drink contains absolutely no H2O,
only water.”
During the Middle Ages, everyone thought living creatures
were made from magical “essences” different to non-living
things. It was a belief called vitalism, but like most ancient
quackery the cracks were beginning to show by the
Renaissance.
In 1745, Vincenzo Menghini burned human organs to ash
and discovered that you could extract iron powder from the
remains with a magnetized knife.9 He concluded that humans
had to contain the base metal iron and that perhaps we weren’t
made of magical ingredients after all.
In 1828, Friedrich Wöhler went even further by
manufacturing urea from cheap lab chemicals.10 Urea is the
main component of urine and therefore was assumed to be
beyond human understanding. Wöhler showed it to be a bog-
standard molecule with the formula CH4N2O.
Whether you like it or not, the elements used in living
biology are no different to those used in sterile chemistry. A
strand of human DNA contains 204 billion atoms, all of them
carbon, hydrogen, oxygen, nitrogen, or phosphorus. There’s no
additional “essence” to make it special.
The iron Menghini discovered is used in blood to bind
oxygen molecules and transport them to the various organs.
When the oxygen gets to where it’s needed, enzymes and
proteins containing chromium, molybdenum, copper, and zinc
help store it, while manganese holds harmful atoms in place
before they cause damage.
When a woman is pregnant she spends nine months
breaking down food and reconstituting the atoms into a baby.
The calcium in milk becomes the calcium in your bones, the
nitrogen in potatoes becomes the nitrogen in your skin, and the
sodium in salt becomes the sodium in your brain. In a very
literal sense, we are what we eat.
It’s not just animals either. Plants use magnesium to absorb
sunlight and vanadium or molybdenum to bind nitrogen from
the soil, a crucial nutrient in growth. It doesn’t matter what the
biological system is, you’ll find every bit of it on the periodic
table.
I’ve occasionally heard people referring to biology as
applied chemistry because of this deep connection, but this
isn’t fair at all. Biology is just chemistry at its most
wonderfully elaborate.
But it comes at a price. Since we are made from the same
stuff as the world around us, that makes us vulnerable to the
same malfunctions.

STRIKING A BALANCE

During the 1500s, Germany was going through a scientific


renaissance and one of its most prominent figures was the
great Swiss physician Paracelsus. His real name was
Theophrastus Bombastus von Hohenheim and he was the first
person to investigate medicine as a science rather than a
superstition (although he did believe in gnomes—nobody’s
perfect).
His most famous dictum is named the Paracelsus principle
in his honor and it’s simple: “the dose makes the poison.” In
other words, whether something is beneficial or harmful is all
about the quantity.
Even something like cyanide is only harmful above a
certain level. In fact, apple seeds contain amygdalin, which
your body converts to cyanide, but you’d need to eat the seeds
from about eighteen apples in order to get sick (assuming
radioactive bananas don’t kill you first).
The metals in your body are the same. If you don’t have
enough copper then your immune system can’t function, but if
you get too much your eyes turn reddish-golden. Beautiful for
sure, but you won’t appreciate it since you’ll be vomiting
blood at the same time.
The element arsenic is famous for its use as a poison, but in
small doses it can treat leukemia.11 It was also the central atom
in Salvarsan, the world’s first wonder-drug and the main
reason we don’t hear much from syphilis these days.12
Antimony can be administered as an anti-bacterial agent but
too much starts killing the host, and a small amount of cerium
can treat tuberculosis but too much gives you a heart attack.13
The Paracelsus principle is why your medication has a
recommended dosage. Get the amount of chemical right and
you save lives, get it wrong and you end them.

WHY ARE THINGS POISONOUS IN THE FIRST PLACE?

The honest-to-God truth is that we don’t know why some


things are bad for you and others are good. Given the number
of chemical compounds that exist, it would be impossible to
catalog the effect of each one. We’ve only known about
molecular bonding since the late 1920s so it’s no surprise that
much of biology is still out of our reach. It’s been doing its
thing for over three billion years so there’s no way we’ll have
it figured out in a century.
Humans are a delicate balance of reactions. If we alter one
of them we can trigger a chain reaction and the final outcome
can be unpredictable.
For example, if you get too much of the element tellurium
in your body it causes horrendous breath and elemental silver
will turn your skin blue, a condition known as argyria.14 Even
nitroglycerine, which we met as the active ingredient in
dynamite, is used to treat angina and nobody’s sure why it
works.15
One of the few poisons the actions of which we do have a
good understanding is cyanide. It works because cyanide
molecules bond strongly to iron. If they happen to bond to the
iron at the center of a molecule called cytochrome c oxidase,
the iron can no longer be used and the whole thing shuts down.
This is bad news because cytochrome c oxidase is the
molecule we need to extract energy from food. Switching it off
means we essentially starve to death in a matter of minutes
rather than weeks.
We also know that some elements, particularly heavy
metals, are poisonous because they’re similar to elements your
body needs and enzymes can accidentally incorporate them.
Zinc is needed for growth, and the element cadmium has a
similar size so if you ingest it the body starts building enzymes
with cadmium instead. Cadmium doesn’t have the right
orbitals to interact with the chemicals in your body, however,
and the result is that you suffer from cadmium poisoning. Your
body stops growing.
Lead poisoning occurs because lead is a similar size to
calcium, needed to manufacture red blood cells, so if your
body absorbs too much lead you can’t make blood. Mercury is
even worse because it’s the right size to fit through membranes
surrounding your brain. Once it gets inside, it can affect your
nervous system, not to mention your thought patterns.
Most people avoid mercury for this very reason but, during
the nineteenth century, warm mercury nitrate was used as a
key ingredient in preparing hat felt. Sure enough, people in the
hat industry soon got a reputation for being a few electrons
short of an atom, hence the term “mad as a hatter.”16

THE FIRE WITHIN

Running all these reactions is exhausting for your body,


requiring a constant supply of energy to stay alive, so you
obtain it by taking in sugar and setting fire to it.
In a chemical context, sugar doesn’t refer to one chemical
but a collection of them. They’re all made from carbon,
oxygen, and hydrogen atoms looped into hexagons or
pentagons, and the stuff in your kitchen is a mixture of two
kinds called sucrose and fructose. The different types of sugar
you can buy, such as granulated, powdered, icing, etc., refer to
the size of crystals rather than the chemicals themselves.
Most of the food we eat contains sugars, which the body
breaks into the smallest type, glucose (C6H12O6). The glucose
molecules then enter a sequence of reactions that convert them
into water and carbon dioxide. The water is lost through sweat
and the carbon dioxide dissolves into your blood where it is
carried to the lungs and breathed out. The air you’re exhaling
now is made from the food you ate this morning.
The original C, H, and O atoms are then repackaged into a
highly unstable molecule called adenosine triphosphate, or
ATP for short. ATP has a chain of phosphorus oxides hanging
off it, which will detach at any given moment, releasing light
and heat as they do. This energy can be absorbed by other
molecules and gets used to drive all the reactions in a cell.
The whole procedure is controlled by molecular machinery
swinging in and out to ensure the correct reactions happen at
the correct time, and its discoverer, Hans Krebs, scooped the
Nobel Prize for mapping the whole carnival.
This is the reason we need food in the first place. Without
sugars we couldn’t supply energy to drive all the other
chemical reactions that make us a living thing. With the
exception of one species (Spinoloricus cinziae, which seems to
have evolved a different way of getting energy), every creature
on Earth carries out Krebs’s reaction.
It’s called respiration from the Latin spirare (to breathe)
and it’s the same thing, chemically speaking, as fire. Some
chemical reacts with oxygen, producing carbon dioxide, water,
heat, and light in the process. We are all walking fire factories.
The only reason we’re not in danger is because it happens
in several stages and on a very small scale. Just as well,
because otherwise we’d burst into flames spontaneously.
Speaking of which …

THE FURNACE WITHIN

The earliest record of spontaneous human combustion was the


death of an unnamed Polish knight during the early sixteenth
century under the reign of Queen Bona Sforza. The account
appears in a 1654 book written by Thomas Bartholin who
heard it as a secondhand account from Adolphus Vorstius, who
heard it from his father, who claimed he once saw a document
where it was written.17 Originally in Latin, the brief
description translates as “he drank two cups of warm wine,
then belched flames and was toasted.”
Spontaneous human combustion (SHC) is a controversial
subject because nobody agrees on whether it happens. The
idea that a person can catch fire without external ignition is
very dramatic but apparently so rare that solid research is
impossible to find. It’s not like you can study a group of
people and see which one of them spontaneously combusts.
It’s spontaneous.
Most SHC reports are like those of the Polish knight above;
spurious secondhand descriptions and probably nothing more
than ghost stories. Plus, the accounts that do give details are
usually easy to explain. But it’s an interesting topic that
captures the imagination so it’s worth looking into.
In most cases of SHC, the remains of a human body are
found charred or melted with the exception of the feet and
hands. The bones are turned to ash and, in most instances, the
surrounding furniture is untouched.
Let’s address the bones turning to ash first. Many people
argue that the temperature of such fires must be fierce in order
to have such an effect. After all, the furnaces in crematoriums
typically run to over 980°C.
However, the need for these high temperatures is because
crematoriums have to burn a body quickly. A flame of a few
hundred degrees is still enough to turn bones to ash provided
it’s left for several hours. If you have a fuel source that lasts
that long, there’s no mystery.
Explaining the fuel is the next task and in 1998 a scientist
named John de Haan conducted a series of experiments in
which he wrapped a pig carcass in cloth and set fire to one
corner. Once the ignition had been provided, the water content
of the pig boiled away and the dry carcass continued burning
for five hours, destroying everything apart from the trotters.18
The explanation for this gruesome demonstration is “the wick
effect.”
The subcutaneous fat of most mammals is flammable so, if
the skin is broken, it can melt and leak into the surrounding
cloth. The fabric is now doused in liquid fat and will burn like
a candle wick for hours, using the full supply of body fat as
fuel. This also explains why feet and hands are the only things
left over; they have very little fat content so the fire leaves
them unscathed.
So how come the rest of the room is always left alone?
We’re used to hearing about fires getting out of control and
buildings burning to the ground because fire will supposedly
spread and destroy everything in its path. But if we really
think about it, we know that isn’t true.
Most fires stay put and combust upward, not outward.
Unless the ceiling is very low a fire will usually have nothing
else to burn once the fuel is exhausted. Think of how you’re
able to stand right beside a bonfire or hold a flaming match
without your skin catching fire. Or think of all the exercise
books sitting in chemistry labs the world over, inches away
from Bunsen burners, none of them burning.
You can hold a piece of tissue paper an inch from a flame
and it still won’t catch. Even if you waft it through the fire
itself, it will only warm up.
Fires that do spread and make the news are usually the
result of direct contact. A forest fire proliferates because the
trees are touching each other or the wind is blowing flames
from one place to the next. Contrary to gut feeling, fires do not
spread through air with ease—otherwise we’d set the
atmosphere ablaze every time we switched on an oven or lit a
cigarette.
Provided there is something to start off the fire, the
discovery of an SHC victim is not suspicious at all and
actually goes along with straightforward science. And it turns
out that in most detailed reports of SHC there is an obvious
source of ignition.
For example, the death of Nicole Millet (February 20,
1725, Rheims, France) is often cited as spontaneous human
combustion since she was found on the floor charred to a crisp
with little damage to the surroundings. What has to be factored
in is that Millet was a heavy drinker and had gone to “warm
herself by the fire” with a bottle of alcohol.19 Hmm.
Similarly, Mary Reeser (July 2, 1951, St. Petersburg,
Florida) was found torched in an armchair, again with little
damage to the room apart from the chair she was sitting on.20
After investigation, however, the FBI concluded that Reeser
was taking sleeping pills, which caused her to fall asleep while
smoking.21 Hmm.
As scientists, we have to be skeptical, particularly of
strange claims. In most cases, it turns out that while human
combustion can happen there is nothing spontaneous about it.
And yet …
I don’t know whether spontaneous human combustion
happens. Almost all the claims turn out to have obvious
causes, but I cannot ignore the fact that one or two do not. Of
the few hundred documented cases of SHC in history, there are
a handful that seem to defy explanation.
The case of Robert Francis Bailey (September 13, 1967,
Lambeth, London) is one such instance. A group of people
walking outside a vacant house in London reported a bright
flickering light inside and called the fire department, who
arrived within minutes. When they entered the house, Brigade
Commander John Stacey reported, Bailey’s body was curled
on the floor with a four-inch slit in his stomach from which a
roaring flame emanated. The house’s electric and gas supply
had been disconnected and there was no sign of matches
anywhere.22 So how did the fire start and why was it bursting
from his gut?
Then there’s the account of Raymond Reed, who was with
the Ninth Battalion of the Royal Welsh Fusiliers during the
Second World War. Reed didn’t combust himself, but recounts
one night in Dorset when he was crossing a field and a nearby
sheep exploded.23 Presumably the sheep wasn’t smoking in
bed.
There’s also the 1867 case of Mr. Watt of Garston whose
corpse suddenly began to burn in a church crypt long after his
death from typhoid fever.24 Not only is it unlikely he was
smoking in bed, he was encased in a coffin.
Accounts like these, if they are to be believed (and that’s a
big if), are difficult to rationalize. The wick effect would
explain the remains, but there doesn’t seem to be a source of
ignition.
We must be careful, though. Just because we don’t have an
explanation for something doesn’t mean we have to accept a
fanciful one. These accounts can’t be explained, but the
sensible thing to do is say we don’t know the explanation, not
put in any hypothesis that we like. There’s no reason to
assume SHC unless we can find evidence for it directly.
Otherwise, we might claim that every unexplained fire is the
result of spontaneous combustion.
There is, however, one detail that pervades every account
of witnessed spontaneous combustion and might just qualify
as potential evidence. The flames are always reported to be
bright blue and originating in the gut.
In 1993, Gunter Gassmann and Dieter Glindemann showed
that the interior of the human gut is capable of forming a
chemical called phosphane (PH3).25 By itself, phosphane isn’t
flammable, but if two phosphane molecules are linked together
they form diphosphane (P2H4), which is. Diphosphane can
spontaneously ignite in the presence of oxygen and burn the
other gases in the vicinity. The main gas within the human
body is methane (CH4), mostly found in the gut and famous
for its blue flame.
Diphosphane often forms in marshland conditions, which is
why people occasionally report blue flames around swamps
and graveyards. So-called will-o’-the-wisp ghosts are actually
methane fires triggered by phosphorus chemistry.
At present, there is no known mechanism that causes
diphosphane to form inside the intestines, but if there was and
if it came into contact with oxygen and if there was enough
methane present, there is a slim chance a fire could
conceivably start.
The scientifically honest answer to whether spontaneous
human combustion can occur is still “we don’t know.”
Diphosphane offers a tantalizing possibility, but speculations
are not proofs. What we can say is that if spontaneous human
combustion really does happen, it’s a one-in-a-billion chance.
I have made it clear to my friends that if I happen to be one
of the few people who dies from spontaneous human
combustion, they need to film the entire episode so that other
scientists can learn something. So, if you ever meet me and
I’m complaining of a stomach problem, cameras at the ready
please.
CHAPTER TWELVE

Nine Elements that Changed the


World (and One that Didn’t)
THE LONGEST EXPERIMENT IN HISTORY

Classifying something as a solid, a liquid, or a gas is usually


straightforward. Solids don’t flow, liquids do but can’t be
compressed, and gases are both compressible and capable of
flowing. These definitions work for most materials but there
are some that aren’t what they first appear, bringing us to our
final record-breaking chemical: pitch.
Also called asphalt, pitch is the sticky black residue left
over when crude oil is distilled. We use it to make our roads
and what makes it interesting is that while it appears solid, it
isn’t. The roads you drive on are made of liquid.
In 1902, an unnamed scientist at the Royal Scottish
Museum in Edinburgh poured a sample of hot pitch into a
glass funnel and left it to cool. For over a hundred years the
pitch has oozed through the funnel and two drops have fallen
onto a dish below.1 To the naked eye it looks like solid black
gunk, but what you’re looking at is the most viscous liquid
known to humankind.
A similar version with a slightly runnier pitch was set up in
1927 at the University of Queensland in Brisbane. That one
has dripped nine times since the experiment began, with the
most recent one falling in 2014.
Time-lapse cameras have captured the slow creeping of
these liquids, but nobody has ever witnessed the precise
instant when a drop falls. Don’t despair, though. If you go to
http://www.thetenthwatch.com/feed you can watch a live
broadcast of the Brisbane experiment as the tenth blob of
liquid slowly forms. You’re welcome.
These two experiments have been running through both
world wars, the rise and fall of the Soviet Union, and the
release of every single Fast and Furious movie, making them
the longest running experiments in history. But if we wanted to
get philosophical for a moment, we could argue that one
experiment has been running for even longer and we are right
in the middle of it.
What happens if you take a planet’s worth of elements,
clump them into a ball orbiting a backwater star, and leave the
whole thing for 4.5 billion years? What will happen within the
planet’s core and what will happen on its surface?
Humans are a latecomer in a long line of chemical
reactions carried out with elements that have been around
since before the dinosaurs roamed. The story of the elements is
also the story of us and the periodic table has been there for
every step, whether we knew it or not.
So, in the final chapter, I want to examine which of the
elements have been crucial to our development and which
ones have had the biggest impact on this experiment called
humanity.

COME BACK ZINC!

There’s an episode of The Simpsons where Bart is forced to


watch a video about a kid called Jimmy who wishes to live in
a world without zinc. He soon discovers his car battery no
longer exists, preventing him from picking up his girlfriend
Betty. Not only that, the rotary mechanism on his phone has
vanished, as has the firing pin in the gun with which he tries to
commit suicide. Jimmy suddenly wakes up screaming, “Come
back zinc!” and breathes a sigh of relief. It was all a terrifying
dream.2
It’s a perfect satire of the hokey educational videos popular
in the 1950s, because nobody has ever wished to live in a
world without zinc. I do know someone who considers zinc
her favorite element, but most people probably know little
about it.
And that’s true for most of the elements on the table. We
know they exist but don’t give much thought to what they do.
If you suffer from kidney problems then you should thank
zirconium because it’s used in dialysis machines for absorbing
ions. If you’re a smoker you owe your habit to cerium because
it’s one of the only metals that produces sparks, allowing your
lighter to function.
If you work in welding, your goggles are tinted with
praseodymium to block yellow light. Or perhaps you work in
the solar-panel industry; if so, ruthenium is the element to get
excited about because it absorbs sunlight better than anything
else.
The microwave you use to heat your meals wouldn’t
function without samarium. The fountain pen you used in
school had a nib made of iridium and if you live in mainland
Europe the dollar bills you spend are impregnated with
europium to detect forgery.
Everyone will have their own favorite element (and if it’s
not phosphorus, what’s wrong with you?) but we can make a
case for some having played a more important role than others.
We could argue that aluminum has been more important
than selenium, for instance. One is used in construction and
vehicle manufacture while the other is used to decolorize glass
and eliminate dandruff. (Having said that, I do enjoy looking
through my windows while running my fingers through a fine
crop of healthy hair.)
If we ignore obvious and boring choices like the oxygen we
breathe or the iron in our planetary core, which elements have
played the most crucial roles in our cultural, political, and
technological evolution? Which ones have made the world
what it is, and which ones are secretly influencing our daily
lives without us even noticing?
This has been a difficult list to compose because as soon as
I settled on one selection I immediately felt I was leaving an
important element out. The problem is that every element is
special. Well, all except for one.

AN HONORARY MENTION

Originally I intended this chapter to be a conventional top-ten


list, but in the end I went for nine. The reason is that there’s
one element that deserves a very special mention but doesn’t
quite fit with the others.
In the process of researching this book I learned the stories
and characteristics of all 118 known elements. Every one is
unique either because it played an important role in the history
of chemistry or because it has a distinct property making it
ideal for a particular use.
I have succeeded in namechecking every element
somewhere in the book at least once, with the exception of
element number 66—dysprosium. The most pointless element
in the world.
Dysprosium was isolated by Paul-Émile Lecoq on his
mantelpiece in 1886 and that seems appropriate.3 It’s a
mantelpiece element if ever there was one. It exists and it
probably has a purpose, but nobody knows what it is.
Dysprosium is neither especially rare nor especially
common. It reacts with water but not as well as group 1
metals. It can be used to make lasers but they’re not as good as
those made from helium or neon. It’s occasionally used in
nuclear control rods, which stop things getting too hot, but you
can achieve the same effect with indium or cadmium.
Dysprosium is beaten at every turn by something else.
There will definitely be a dysprosium scientist out there
who’s currently foaming at the mouth as she reads this. But
dysprosium doesn’t seem to be exclusive in any way, which
makes it quite interesting.
I hereby declare dysprosium to be the only element you
could remove from human history and absolutely nothing
much would change. We salute you, dysprosium, the most
boring element on the periodic table.
Right, on with the list.

ELEMENT OF AGES

Carbon is an obvious choice to kick things off. It’s so vital to


our world it’s practically humdrum. Look around the room and
probably 90 percent of things you’re looking at are either
made from, extracted with, or powered by carbon. It is the
element that has defined the ages of humankind.
We’ve been around for hundreds of thousands of years, but
what we call civilization began with manipulating metals. The
Stone Age represented the primitive infancy of our species but
it was the Bronze and Iron Ages that were the turning points.
Before we mastered the art of metallurgy, the only metals
we knew about were gold and occasionally silver, so all our
construction materials, weapons, and tools came from bashing
rocks together. Then at some point between 8000 BCE and
3000 BCE everything changed.
Most metals in nature are bonded to oxygen, but oxygen
forms better bonds with carbon. This means if we mix enough
carbon together with our metal oxide (rock) and give the
whole thing some energy (heat it), everything rearranges to
carbon dioxide and pure metal. This technique, known as
smelting, was the most important chemical reaction since fire
itself.
The early technologists, whoever they were, discovered
that roasting rocks in the presence of charcoal produced metal.
First, we began extracting copper and tin, giving us bronze.
Then we learned how to get the fires hotter and started
extracting iron, previously only found in meteorites.
By the nineteenth century we were burning carbon itself as
a fuel source, using it to run our combustion engines. Carbon
has an advantage over other fuels because, rather than leaving
unpleasant residues, it burns away to an invisible gas. Where’s
the harm in that?
Today, we still use coal for our power stations, so the
electricity you use is most likely down to carbon too. It’s only
in the past sixty years we’ve realized that all that CO2 has the
awkward feature of absorbing infrared radiation, slowly
heating the atmosphere as the decades tick by.
On the plus side, carbon is also the basis of polymer
chemistry. Take a long chain of carbon atoms, use hydrogen to
make sure each one has the correct number of bonds, and if
you tangle the ropey chains together you end up with a plastic.
Imagine a world without plastic, metal, or widespread
electricity and you begin to see why carbon is so important.
Carbon’s versatility is a result of its location on the periodic
table. It sits on the top row, making it a small atom capable of
forming tight bonds, in the fourth column along, giving it four
available bonding electrons.
An element like fluorine is also on the top row but it is only
one electron away from a filled shell, meaning it will form one
bond and then stop. Carbon has four electron spaces, meaning
it can form four links to other atoms, all of them strong.
Other elements that form multiple bonds are usually too big
for the bonds to be robust, so carbon has the best of both
worlds, which is why we find it in everything from our cell
membranes to our cell phones.
It gave us the materials we use and the power to manipulate
them, and now its presence in the air is threatening to knock
our climate out of equilibrium. If there is one element that has
turned the course of human history more than any other, it’s
carbon.

FOOD FOR AN EMPIRE

At the start of the 1800s Britain’s armies were expanding


across the globe. The Napoleonic wars were finishing, slavery
was coming to an end, and the Empire was approaching its
“golden age.” But the admirals and generals of this ruthless
military machine were facing a problem. An empire is only as
strong as its food supply. How do you get food to thousands of
people, far away from where it’s being produced?
The answer was discovered by a French inventor named
Philippe de Girard who devised a method to vacuum-seal food
in a tin can. After testing his invention on several British
scientists, the idea was sold to the engineer Bryan Donkin who
set about improving the method.
Donkin was already a superb craftsman who consulted on
the manufacture of Babbage’s difference engine and Telford’s
suspension bridge, and was also the inventor of the humble
pen. While the inventor of the pen is sometimes misattributed
to John Loud in 1888, Donkin already had a patent in 1803.4
Let’s just get our pen history right, folks.
By 1813, Donkin had designed a method to mold tin cans
in such a way that food inside would be locked in without any
air, meaning it could last for years and be transported as far as
was needed.
After Queen Charlotte sampled a tin of his corned beef and
praised the taste, Donkin began manufacturing tin cans en
masse and sold them to the Navy. Tin cans allowed countries
to feed their armies during both world wars and today over
forty billion are sold around the world annually.5 While many
of these cans are made from steel today, it’s the tin plating that
prevents irreversible rust.
What also makes tin special is not what it does as a pure
metal, but how it can modify other metals when they are
mixed together, forming what’s called an “alloy.”
Its softness is one of the reasons it is mixed with copper to
make bronze. When alloyed with lead it forms pewter, the
material most cutlery was made of until very recently. When
mixed with a bit more lead you end up with solder, the “glue”
used in electronics to join wires.
Bell metal, used for making bells, is an alloy of tin with
copper. Gunmetal used for making, well, guns, is an alloy of
tin with copper and zinc. Terne, used to make roofing, is tin
mixed with lead again. Ball bearings are usually made from tin
with copper and iron. Galinstan, used for telescopes, is tin
mixed with gallium and indium, and the list goes on. Tin is the
great modifier of the periodic table.
It’s not quite as prevalent as iron but it has the clear
advantage of being rustproof and, because it’s easily extracted
and manipulated, anyone can work with it from the richest
monarch to the lowliest commoner. While armies and
politicians might have prized elements like gold, tin has
always been the element of the people. Not that gold hasn’t
been important, too, mind you.
ALL THAT GLITTERS

The color of gold has led many cultures throughout history to


worship it, often associating it with the sun (silver being linked
to the moon). It arises because gold has large gaps between the
atomic orbitals, so visible light loses a lot of energy when it
strikes. The highest energy colors like violet, blue, and green
are absorbed into the metallic surface while the yellows and
oranges are bounced back out. Cesium and copper also have
yellow/orange hues, but nothing compares to gold.
As we saw in Chapter 3, gold was essential to discovering
the nucleus and therefore modern chemistry itself. It was used
because it’s the most malleable metal available, so soft that 28
g would be enough to make a wire stretching nine times the
height of Everest.6
This ease in molding, as well as its shine, has also led to its
use in jewelry since prehistory, not to mention the fact it
doesn’t tarnish. While other metals will gradually react with
oxygen, gold will gleam forever.
It’s also a very rare metal. If you were to collect all the gold
deposits in the world it would total around 170,000 tons. That
would barely fill three Olympic-size swimming pools.7
This combination of malleability, rarity, permanence, and
beauty are what make it so precious. Gold can be traded
anywhere in the world, regardless of local custom, because
everyone values it.
In Finland the skins of squirrels used to be acceptable as
money, and up until the twentieth century Ethiopia used blocks
of salt.8 Money is different wherever you go, but gold is
revered everywhere and always has been, making it the only
true international currency.
Alexander the Great led the Greek army to conquer the
Persian Empire—the largest in the world—in order to steal
their gold. Julius Caesar did the same thing to western Europe.
So did King Ferdinand of Spain, sending his conquistadors to
rip gold from the Americas (and we all know how that story
turned out).
The first gold coins were used in China during the sixth
century BCE, but by the 1800s every large country in the world
(apart from China, ironically) was using a gold standard for
international as well as domestic business.
Owing to its rarity and weight, though, gold coins are far
from practical so banks began printing contracts that
corresponded to a certain amount of solid gold. This was the
invention of modern money itself.

KNOWLEDGE AND POWER

Some of the elements have a split personality. The same


substance can be of great benefit to the world, but also the
cause of endless pain. No other element can lay claim to
having enlightened so many or killed so many as lead.
Once extracted from its ore, lead is a dull metal with three
important properties: density, meaning it’s hard to break,
malleability, meaning it can be bent, and corrosion resistance,
meaning you can have it in contact with water.
The Romans carried out lead mining on a grand scale
because they used it for pipes and waterworks. Iron is no good
because it rusts so lead was used at a rate of thousands of tons
per year. The very notion of water straight to people’s homes
hadn’t been explored properly until the Roman plumbing
system. Even the word plumber comes from the Latin word for
lead, plumbum, because piping specialists were plumbum
experts. That’s why it has a silent “b.”
Because of its toxicity, some people have speculated that
leadpoisoning contributed to the decline and ultimate defeat of
the Roman Empire.9 This seems unlikely, however, as lead
poisoning was already a known malady and water doesn’t
usually dissolve enough to reach dangerous levels.10
It’s possible that boiling grape juice in huge lead vats may
have caused lead poisoning in some of the aristocracy, but this
is speculation at best. It’s unlikely that lead caused the collapse
of Roman civilization, but don’t worry, it’s still responsible for
millions of deaths every year.
In thirteenth-century China, it was realized that a small
tube of gunpowder could launch a projectile at high velocity
when it exploded—the invention of the gun. The technology
spread to European armies and the best metal for making
bullets turned out to be lead—not only because it’s readily
available and easy to manipulate, but because it is so dense
that once it is fired from the barrel it keeps going in a straight
line. No other metal allows us to shape it so well while being
dense enough to hold its trajectory.
Nobody knows how many bullets are manufactured in the
world today but the number is probably in excess of ten billion
a year: enough for one bullet per person. It’s hard to think of a
weapon that has caused more death than guns firing lead
bullets.
But lead has also done wonders for us. In 1440 Johannes
Gutenberg was looking to find a way of quickly conveying
information to people. Up until then, every text and book had
to be copied by hand. If a machine could be rigged to do the
job, books could be produced in a matter of days rather than
months.
The result was his printing press, only achievable thanks to
lead (alloyed with a little tin). Because lead was so malleable,
it could be carved into the precise shapes of block letters.
Other metals could be molded too, but lead’s density meant
hammering it repeatedly onto a page wouldn’t cause it to wear
away.11 The same characteristics that help lead kill are those
that help it educate.

DRINK IT DOWN

People are living longer these days. Obviously a good thing.


The only downside is that we’re more prone to age-related
disease. This has led to a lot of hoo-ha and fear-mongering
about the apparent rise of cancer and heart disease. I’ve heard
everything blamed from GMO foods to (perversely)
chemotherapy drugs themselves, but it really comes down to
cold numbers.
Humans die. Sorry to break that to you. Our bodies are
fragile and they aren’t built to last. The older you get, the less
you tend to function and the more likely you are to die from
something like cancer or heart disease. The only reason we’ve
seen an apparent rise in these deaths is because people are
lasting long enough to die from them. Age-related illness has
existed as long as the human body; it’s just that most people
tended to extinguish before they got that far.
Death is always unpleasant but I would say age-related
illness is a fair price to pay for a life expectancy in the
eighties. During the mid-1800s, life expectancy was forty-two,
mainly because people died in childhood, bringing the average
down.12 The only reason we enjoy a higher number today is
simple. It has little to do with a gluten-free diet or a Pilates
class. It’s because we have defeated the world’s number-one
killers. We don’t die of infection anymore.
In the 1340s, hundreds of millions of people died from
bubonic plague. Between 1817 and 1917, an estimated thirty-
eight million died from cholera.13 Measles and smallpox have
been responsible for more deaths worldwide than any war you
care to mention and don’t get me started on polio or malaria.14
In many parts of the globe, these diseases are still rampant, but
in the West we are fortunate because we have eradicated them.
Quite frankly, dying from old age is something for which we
should be grateful. Many are not so lucky.
The reason we aren’t seeing epidemics breaking out every
year is down to two things: vaccination and element number
17, chlorine.
The first widespread use of chlorine was during the First
World War when the German chemist Fritz Haber introduced it
as a chemical weapon. In 1915, he oversaw the installation of
five thousand canisters of it along a 7-km distance at the
western front and, when the wind began blowing the right
way, Haber ordered the canisters be opened.
Chlorine is a thick green gas that rolls along the ground
like a liquid. Carried by the wind, the chlorine was dragged
toward the British Army and flooded their trenches,
asphyxiating and blinding thousands of men.
According to Hermann Lutke, on May 1, 1915, a party was
being held in Haber’s honor to praise his simple but effective
use of chlorine chemistry. A few hours after the party, his wife
Clara (a noted pacifist) took Haber’s service revolver into the
garden and shot herself in the chest, dying moments later in
her son’s arms.15 In this context, chlorine has a similar
reputation to lead but, just like lead, it can be put to a far better
use.
Because it is lethal to biological organisms, if handled
correctly it can be used to kill the pathogens that would
otherwise be lurking in our water supplies.
The average person in the US and the UK uses around 340
liters of water a day and it has to be clean in order to stop the
spread of disease.16 Even toilet water has to be potable
because if it contained anything harmful it could get airborne
during flushes.
There are a few alternatives to chlorination, such as
bubbling ozone through the water, but chlorine is the main
choice for every European country and the whole of the
United States.
It works because chlorine dissolves to form hypochlorous
acid, HOCl, which is lethal. That’s what kills you if you are
unfortunate enough to inhale it as a gas. It’s easy to remove
from water, though, so if we pump it into our drinking supply
it will kill everything, after which we remove the excess with
carbon charcoal.
While adding fluoride to the water has caused controversy
(largely because it was implemented before long-term studies
were finished), nobody objects to chlorine. It’s the main reason
you aren’t currently dead.

THE SILVER SCREEN

Any writer of non-fiction, no matter how objective they claim


to be, will be writing with bias, putting their own personal
views into things, often without realizing it. And by the way,
don’t you just hate celery? Almost all of recorded history has
been the result of eyewitness accounts and people’s memories,
which makes things hard to verify.
That changed in 1717 when a German chemist named
Johann Schulze left a bottle of silver nitrate and chalk on his
windowsill. Schulze placed the bottle down absentmindedly
and, when he picked it up a few minutes later, was shocked to
find it had turned brown. Except for a thin white line
suspended in the liquid.17
He looked out the window to see what could have reacted
with his solution and noticed a piece of thread hanging across
the window in exactly the same shape as the white line inside
the bottle.
Where the sunlight hit the silver nitrate it made it go dark,
but where something had obscured the sun, the liquid
remained white. Schulze had taken the very first photograph
and it was a liquid. Considering Henri Becquerel’s
radioactivity discovery, it’s strange how many times leaving a
jar lying around has led to a monumental realization.
Silver atoms can be bonded to nitrate molecules in solution,
but given a bit of energy they can separate and form solid
metal. A lump of solid silver glistens brilliantly but powdered
silver is a dark brown, showing exactly where the light has hit
it.
It was a French inventor named Joseph “Nicéphore”
Niépce who realized that putting silver compounds onto a
piece of paper and focusing images with a pinhole camera
would create a black-and-white copy of what was being
projected. In 1829 he used this technique to capture the
world’s first proper photograph from his bedroom window,
View from the Window at Le Gras, taking eight hours of
exposure time.
At least, that’s the official story. In 1777, another scientist
had already discovered that you could trap silver-solution
images on a piece of card using ammonia. This scientist also
figured out the cause of the phenomenon but never pursued the
research, denying himself the title of being the inventor of
photography. That scientist (and I’m not making this up) was
none other than Carl Scheele.
Over the following century we discovered other silver
chemicals that reacted faster than nitrate and, by using lenses,
we were able to intensify the light, creating instantaneous
images of particular moments. We didn’t have to rely on word-
of-mouth or written accounts to store information anymore:
silver allowed us to capture pictures of things as they truly
were.
Historians disagree over who took the idea of photographs
and strung them into movie reels, but the patent for the first
film camera seems to have been filed by one Wordsworth
Donisthorpe in 1876.18 He used it to film a few seconds of
Trafalgar Square and started the movie industry, which we still
call “the silver screen” after the element involved.
Color photography relies on silver as well but with
additional chemicals that respond to different frequencies of
light. When red light hits the layer of film containing the red-
sensitive chemical, it causes silver powder to form. The same
happens in the blue layer with blue light, and the green layer
with green light.
You end up with a black-and-white image of where the reds
should be, one for the blues, and one for the greens. By
developing each layer with the correct dye in the right order,
you can recreate the color image with which you started.
It’s hard to imagine a world where photographs or film
didn’t exist because they are so ubiquitous. We rely on them to
give us reliable information and silver is what made it
possible. In recent years the invention of computer editing
technology and digital cameras has changed things somewhat,
but there was originally truth to the phrase “the camera doesn’t
lie.”

DESTROYER OF WORLDS
For a long time, the science of nuclear bombs was highly
classified. When Nobel Prize–winning chemist Linus Pauling
gave a public talk on the subject, an FBI agent showed up at
his office to interrogate him on how he knew the workings of a
bomb so perfectly. Pauling responded, rather coolly, by saying,
“Nobody told me, I figured it out.”19 Nowadays, the design of
an atom bomb is well known and it’s all about uranium.
Uranium atoms have 92 protons in their nucleus and
usually 146 neutrons, giving a total of 238 particles. But
roughly 0.7 percent of uranium atoms have 143 neutrons
instead, making uranium-235. This combination of protons
and neutrons is volatile, and when the nucleus fractures it spits
out neutrons. These neutrons go flying off and get absorbed by
other uranium nuclei, making them unstable and causing
another fission event.
If you’ve got 1 kg of uranium, most of your neutrons can
escape through the surface of the metal, but once you get to
around 47 kg, what’s called critical mass, the neutrons in the
center don’t escape. The energy builds, fissions multiply, and
the result is a nuclear blast.
It might be fair to say that gold dominated global politics
up until 1945, but uranium certainly dominated it afterward.
With 47 kg, you can end one war and start another.
On August 6, 1945, a uranium bomb was detonated over
Hiroshima, causing the deaths of over eighty thousand people.
Three days later, a plutonium bomb (made from uranium as a
starting agent) was dropped on Nagasaki, killing forty
thousand people and bringing the Second World War to a
close.
With the ability to manipulate uranium, the United States
became the most powerful nation on Earth. That kind of
strength invites challenge. Four years after Hiroshima and
Nagasaki, the USSR demonstrated its own nuclear capability
and the Cold War began, shaping the technological, cultural,
and economic landscape of the twentieth century.
Most nuclear weapons today are based on plutonium but
uranium is still the starting ingredient. Getting hold of it isn’t
difficult, mind you. Uranium was used for Fiesta dinnerware
glazes (hilariously the US government confiscated the
products during the Cold War). The tricky bit is extracting the
0.7 percent of atoms that are fissile.
At the time of writing, nine nations have the technology to
do this, with the United States and Russia owning the largest
stockpiles. The precise number of warheads is unknown but
it’s estimated to be well over five thousand each.20 With that
arsenal, you could eliminate all life on Earth many hundreds of
times over.
The physicist who coordinated the invention of nuclear
weapons was Robert Oppenheimer. He was once asked in an
interview what it was like to witness the first nuclear-bomb
test, codenamed Trinity. His response was chilling:
We knew the world would not be the same. A few people laughed, a few people
cried. Most people were silent. I remembered the line from the Hindu scripture
the Bhagavad Gita. Vishnu is trying to persuade the prince that he should do his
duty and, to impress him, takes on his multi-armed form and says “Now I am
become Death, the destroyer of worlds.” I suppose we all thought that. One way
or another.21

WE WANT INFORMATION … INFORMATION … INFORMATION

Silicon sits right below carbon on the periodic table and has a
similar electronic structure. The only difference is that it’s
bigger, so its bonds are not as strong.
It can form crystals similar to diamond with comparable
strength, and can also be strung into plastic chains, the most
famous of which is the silicone gel responsible for some
people’s career success in Hollywood. But silicon’s primary
use is at the nerve center of every electrical device you own.
If the nineteenth century is remembered for the Industrial
Revolution and the internal combustion engine, the twentieth
century should be remembered for the silicon revolution and
the transistor, an invention of which many have never heard.
Invented in 1947 by Walter Brattain, William Shockley,
and John Bardeen (the only person to win two Nobel Prizes for
physics), transistors are to computers what bricks are to
houses. Your smartphone contains about three billion of them
and your laptop contains seventy times that.
A transistor’s job is to let electrical current pass through it
sometimes and block it at other times. On its own, this sounds
mundane, but get enough transistors hooked up in an intricate
pattern and you’ve got a microchip. By programming a series
of instructions for these transistors as 1s and 0s, we can tell
transistors to switch currents on and off, allowing us to control
circuits and store information.
The problem with making a transistor out of metal is that
metals always conduct. Similarly, non-metals are always
insulators. In order to create something capable of switching
on and off at different times you need an element that is
halfway between a metal and a non-metal. Enter silicon.
Silicon atoms are large so they’re vaguely metallic in
nature, but their shape has more in common with non-metals
like carbon and boron. These hybrid properties make silicon a
semi-conductor and its crystals form the backbone of
transistors.
Not only that, silicon is also the key ingredient in glass,
giving us the optical fibers for the internet. Not to mention
making windows.
Most optical fibers are made by one company called 3M
and their glass is so transparent that, if you were to make the
ocean out of it instead of saltwater, you would be able to see to
the bottom with perfect clarity.
During the 1950s, after inventing the transistor, William
Shockley set up a business in California doing research with
the computer science department of Stanford University.22
Prior to his invention, all computers were mechanically
based and occupied whole rooms. Silicon offered the
possibility of computers you could have on your desktop.
Once interest in silicon began to boom so did the local
economy, and today Shockley’s neighborhood is the
headquarters of Apple, eBay, Facebook, Google, Intel, Netflix,
Yahoo, and Visa. It’s a region of southern San Francisco called
the Santa Clara Valley, more commonly known by a name
inspired by the element that built it: Silicon Valley.
Silicon enables us to perform calculations that previously
took a library of people days to complete and runs everything
from our digital watches to our mobile phones, although that
technology comes with a moral dilemma tied to a different
element—tantalum.
Tantalum vibrates when electrified, making its importance
in mobile phones obvious. Seventy percent of the world’s
tantalum deposits come from the Democratic Republic of
Congo, a country whose economy is based on its mining and
export. The civil war that raged there from 1994 to 2002, the
bloodiest conflict since the Second World War, was funded
through the sale of tantalum.23 Sometimes our relationship
with the elements is ethically quite dark.

SAVIOR OF WORLDS

In the 1930s, hydrogen was set to be the element of the future.


It’s easy to get ahold of, easy to transport, and when it burns
the only by-product is water. It’s the cleanest, greenest fuel
imaginable.
Not only that, its low density makes it perfect for
generating lift. An airplane needs to build a lot of speed for its
wings to bite the air, but a hydrogen dirigible will float without
assistance or persuasion. Helium is less reactive, which makes
it safer, but when the United States began stockpiling it in
1925 at the National Helium Reserve in Amarillo, European
agencies turned to hydrogen as the obvious alternative.
The German government was particularly eager to harness
hydrogen technology and in 1931 began constructing the
world’s largest zeppelin, LZ-129 Hindenburg, a marvel of
chemical and aeronautical engineering.
But on May 6, 1937, as it was being tied to the ground of
Lakehurst Naval Air Station, the Hindenburg caught fire.
Nobody knows how it started (spontaneous zeppelin
combustion?) but in under thirty seconds all two hundred
thousand cubic meters of hydrogen had combusted.24
The crash was caught on camera and the accompanying
audio by Herbert Morrison shouting “Oh the humanity!” has
become iconic. People saw what a hydrogen fire looked like
and the age of the zeppelin was over before it began.
The world didn’t hear much from hydrogen for a few
decades, until the USSR detonated the Tsar Bomba in 1961
(see Chapter 4). The Tsar wasn’t any old uranium bomb: it was
a hydrogen bomb and the difference was obvious. With a
mushroom cloud reaching 64 km into the sky, it made the
bombs dropped at the end of the Second World War look like
fire crackers.
The exact details of how a hydrogen bomb works are still
classified and, bearing in mind what happened to Linus
Pauling, I’m reluctant to do extensive research on it. While
writing this book, I’ve investigated the price of plutonium and
how much thallium is needed to kill someone. I should
probably exercise caution before I start asking people how to
build an H-bomb.
The basic premise, though, is fairly well understood.
Einstein’s E = mc2 equation tells us that we can obtain energy
from an atom by splitting it. What’s surprising is that reversing
the process and fusing nuclei together releases even more
energy (because quantum mechanics, that’s why).
The bomb works in two stages (I think). First, a
conventional uranium bomb is triggered and the heat from that
blast causes a capsule of hydrogen atoms to fuse, generating a
miniature sun as they convert to helium. That’s the awesome
power being demonstrated in images of the Tsar Bomba
explosion.
Combined with the terrifying sight of the Hindenburg,
hydrogen has been an element of terror in the public eye. But
we shouldn’t give up on it. In fact, as the future creeps toward
us, we may find ourselves becoming entirely reliant on it.
The energy released from fusing hydrogen doesn’t
necessarily have to be done in one go. In the same way
uranium rods can be brought near each other to generate heat
rather than explosions, it should be possible to bring hydrogen
nuclei together in controlled conditions.
Fusion-based nuclear plants would produce no toxic
products, would end our dependency on fossil fuels, would
end all conflicts fought over fossil fuels while providing
limitless energy for the planet, as well as bringing an end to
human-made climate change. Fusing hydrogen could really be
the ticket humanity needs to solve all of its problems. There’s
only one slight snag—we haven’t been able to do it.
In order to fuse hydrogen atoms, you have to heat them fast
enough to collide. That takes energy, and all our current fusion
reactors take more power to get going than we can usefully
extract.
We have only achieved one positive fusion reaction so far,
at the National Ignition Facility in California in 2013. There, a
group of researchers led by a man whose name is genuinely
Omar Hurricane, blasted samples of hydrogen with laser
beams and excited them into fusing. Mr. Hurricane and his
team are the first and, to date, only people to have successfully
got more out of a fusion reaction than they put in.25 It’s not
perfect and it’s not enough to power the world, but it’s a
promising step.
And there’s something that may be even more important.
Because hydrogen burns beautifully with oxygen, it’s a perfect
rocket fuel. Those enormous tanks you see on the sides of
ships being launched into space aren’t full of petrol, they’re
full of chemicals that generate hydrogen and oxygen.
Hydrogen isn’t just the element which may save the world,
it’s the element which may help us leave the world altogether.
And sooner or later, we’re going to have to.
At the moment we’re living in a golden age of pulling
elements out of the ground with abandon, but it can’t last.
Assuming our planet doesn’t get obliterated by an asteroid
(we’re overdue), we will eventually consume all the resources
Earth has kindly given us.
If our species wants to survive, we’re going to have to do it
somewhere else, which means we need to get out there and
exploring. For that, we’re going to need hydrogen, our ticket
to ride the universal express.

A FINAL THOUGHT

Each element on the periodic table has a story to tell but what
that story is depends on us. It is our duty not to abuse such
power. And I don’t think we will.
When I look at the periodic table, I see a monument to how
far we have come and how much we have learned in such a
short time. Through science we are capable of understanding
the Universe and using its resources to do amazing things. I
truly believe that science will save our species.
APPENDIX I

Sulfur with an “f”


Naming an element is usually an honor given to the person
who isolates it. Unfortunately, it can cause disagreements
when scientists give elements unpopular names.
In 1875 the French chemist Paul-Émile Lecoq named a
new element gallium from the Latin gallia, meaning France.
However, it was soon suspected he had been a bit sneaky.
Gallus is also the Latin for the rooster, which translates into
Lecoq, his own name. Perhaps he had immortalized the Lecoq
name by subtly naming the element after himself.
To try and solve these problems, the International Union of
Pure and Applied Chemistry, IUPAC, has stringent rules for
the naming of a new element. Elements can be named after:
1. a character from mythology (e.g. thorium, after the Norse
god Thor);
2. a place (e.g. rhenium, from Rhenus, the Latin name for the
Rhine river);
3. a property of the element (e.g. bromine, from the Greek
bromos, meaning foul stench);
4. the mineral from which it was extracted (e.g. samarium,
after the mineral samarskite);
5. a scientist (e.g. roentgenium, after Wilhelm Röntgen the
discoverer of X-rays).
IUPAC will deliberate a proposed name for up to five
months before giving the thumbs up and then, once they have
spoken, the name is internationally recognized and periodic
tables are adjusted to incorporate it.
Many British chemists were horrified in 1990 when IUPAC
endorsed the American spelling of sulfur with an “f” as
opposed to the British spelling of sulphur with a “ph.” And
throughout this book, I have gone along with their decision.
To be clear, IUPAC is well within its right to favor sulfur
over sulphur. The etymology of the word is unknown and
anyone who claims otherwise is misinformed. The first
recorded usage is to be found in the writing of the second-
century BCE poet Ennius, who called it sulpureus. That word
itself (whose etymology has been lost), however, may come
from the word swefel (whose etymology has also been lost).
As we don’t know where the word comes from, there is no
reason to prefer one spelling over another. Spelling it with a
“ph” isn’t just a matter of British pride: it’s refusing to accept
the agreed international standard.
Personally, I find it infuriating I have to write the name of
an element with a lower case and my own name with a capital,
but I have to play ball (in this book, at least; on my website I
capitalize whatever I feel like!).
I have heard some people suggest IUPAC accept a trade
where sulfur is spelled with an “f” in exchange for aluminum
being spelled with two “i”s (the British way, aluminium). It’s
maybe worth pointing out that Humphry Davy, the English
scientist who named it, did originally choose aluminum so the
American spelling is more authentic after all.
APPENDIX II

Half a Proton?
The closer we look at particles the more substructure we find.
Atoms are made of electrons and a nucleus. The nucleus is
split into protons and neutrons. How do we know when we’ve
really got to the bottom of it?
During the 1960s, theoretical physicists decided it was time
to take a bottom-up approach rather than a top-down one.
Starting with the basic laws of nature, what fundamental
particles should we see arising? The resulting framework,
called quantum field theory, predicts a buffet of particles, all of
which have been found, so the approach is definitely along the
right track.
Electrons turn out to be fundamental, as do photons, the
particles that make up light. There are lots of others with
names like neutrinos, gluons, and Higgs bosons, but protons
and neutrons are not on the list.
It turns out that protons and neutrons themselves are not
fundamental but can be thought of as three particles that are. A
proton can’t be split in half but it can be described as being
made up of thirds. Murray Gell-Mann named these particles
quarks (pronounced “kworks” not “kwarks”).
It still wouldn’t be right to say you can chop a proton into
thirds, however, because they don’t actually let you do that.
Quarks do not exist as individual things, but in little pairs and
trios.
If you were to take a proton, made from three quarks, and
break it apart you wouldn’t end up with the three individual
quarks, you’d end up with six … that’s quantum field theory,
folks.
So, while in one sense you can describe a third of a proton,
you could never actually have it. The quarks never leave their
proton so it’s fine to talk about protons as if they were
fundamental particles. They might as well be!
APPENDIX III

Schrödinger’s Equation
Schrödinger’s equation is a full description of everything we
can know about what a particle is doing. We might be
interested in looking at how a particle is going to behave at a
specific point in time or at a specific point in space. We might
not be interested in either and only want to know what
energies are involved or what rotations a particle can have.
This means there are lots of different forms of the
Schrödinger equation and different ways of writing it. The
most straightforward one is called the generalized time-
dependent Schrödinger equation and it looks like this:

i represents the square root of minus one. Any normal number,


either positive or negative, always generates an answer that is
positive when multiplied by itself. −2 × −2 isn’t −4, it’s +4.
But that means −1 would have no square root, so there must be
another type of number which multiplies by itself to generate
negatives. These numbers are called i numbers. The reason the
letter i is chosen will muddle things at this point so don’t
worry about it: it’s a historical convention and it’s
meaningless. It may seem strange that we have to use freaky
numbers in our equation because it seems like a cheat but
that’s not what’s happening. Nature does things outside of our
normal experience so we have to use numbers outside our
normal experience in order to make sense of it. When we try
using regular numbers, the equation gives answers that don’t
match reality. It would seem that nature uses i numbers so we
have to as well.
H is called the Hamiltonian and it refers to the total energy
of the thing we’re examining. We’ve written it here with one
letter but that’s a shorthand. Written in full, the Hamiltonian is
a long-winded term that takes into account the particle’s mass,
its kinetic energy, how far it is from the nucleus (called its
potential), and so on, but it still just means how much energy
the particle has.
Ψ: This symbol (psi) represents something called the
wavefunction. It can mean a great many things, but in the
context of chemistry it refers to the fact that the probable
location of a particle ripple. Rather than having a specific
coordinate in space, an electron’s location has a wavelike
character when it’s not being interfered with. The
wavefunction takes this into account.
|> is called a ket vector. What it refers to, generally
speaking, is the state that something is in. In this case the state
of the particle’s wave-function. The left side of the equation,
read in full, is now telling us that if we calculate the total
energy of the wavefunction’s state and multiply the answer by
negative i, we’ll get something useful.
ħ is called Planck’s constant and represents 1.055 × 10−34
Joule-seconds. This number is a property of the Universe that
relates the energy of a particle to its frequency. Frequency is
how many times something will ripple per second and since all
particles have ripply movement, we need a term that relates
the two. Specifically, if we divide the energy of a particle by
its frequency, we get a number called h, which is 6.626 × 10−34
Js in SI units. Planck’s constant is arrived at by dividing h by
2π. We do this because 2π is often used when measuring
frequencies, so we include it as part of our constant to make
the equations neater.
∂ is called a partial differential. It’s a symbol that tells us to
measure how one property changes when you’ve got lots of
other stuff going on and you only want to focus on one thing.
In this case is telling us to compare the change of something
compared to t (time).
So, the whole equation is telling us that if we can work out
the total energy a particle has in a particular state (left-hand
side), we can work out how its behavior will change with time
(right-hand side).
If you know what energy an electron has, you can predict
where it’s likely to be at any moment. Do this for all three
spatial dimensions and you’ll end up with a description of
where an electron is likely to be around its nucleus—the
orbital.
APPENDIX IV

Neutrons into Protons


We’ve already met quarks in Appendix II and they can be
tricky things. They come in many varieties but they all have an
electric charge that adds up to the charge of a proton or a
neutron.
An “up” quark has a charge of +2/3 while a “down” quark
has a charge of −1/3. When two up quarks and a down quark
occur in a trio, the charges combine to create an overall +1, a
proton. If an up quark occurs with two down quarks, however,
the charges cancel and the result is a neutron.
But quarks don’t stay as one type. An up quark can turn
into a down quark and vice versa. A neutron is an udd (up
down down) combination, but if one of the downs turns into
an up we get an uud (up up down)—a proton. It’s the quark
inside that flips and turns a neutron into a proton.
When this happens the amount of overall charge has
changed and for some reason this is a big no-no for the
Universe. Rather than creating a charge imbalance, the
Universe prefers to keep things neutral so it does a particle
shuffle.
When the −1/3 down quark changes character, it emits a
particle called a W− boson, which carries away a −1 charge,
leaving a +2/3 charge behind.
The W− quickly splits into an electron, which retains the
charge, and another particle called an anti-neutrino, which has
none. And there’s no simpler way of describing the whole
process.
APPENDIX V

The pH and pKa Scales


You may have met the pH scale in school. The more acidic
something is, the lower its number. Acids tend to have values
below 7 while non-acidic things tend to be 8 and upward. The
reason the scale was introduced was because the numbers
involved in acid chemistry are often extremely small.
Suppose we had 1 × 105 hydrogens in a 1-liter bottle and 1
× 104 in another. The first contains 100,000 and the second
contains 10,000. Clearly, the first is ten times more
concentrated than the latter, but they are both extreme
numbers. So we invoke the laws of logarithms.
A logarithm is the number of times you have to multiply
something by itself in order to get a particular result. Say you
had the number three and multiplied it by itself four times.
That would be written as 3 × 3 × 3 × 3 or, more simply, 34.
The answer is 81.
But suppose you want to do your calculation the other way
around. You want to know how many times you had to
multiply three by itself to reach 81. You would write it like
this:
Log3 81 = 4
In other words, what number do I have to raise three to in
order to reach 81. The answer would be 4.
In our earlier example, we had one solution containing
100,000 hydrogens. If we express this logarithmically we
would write:
Log10 100,000 = 5
Five is a much easier number to work with so we might
describe this acid as a “5 solution.” The number isn’t telling us
the concentration directly but it’s telling us the order of
magnitude with which we’re dealing.
Likewise, a solution ten times more dilute would be called
a “4 solution.” This is useful because, when we’re dealing
with something huge like a concentration of
1,000,000,000,000,000,000, it’s easier to call it an “18
solution” rather than write the whole thing out.
So why does the scale run backward? The answer is that
most acids, even very concentrated ones, only contain a small
number of hydrogens per liter.
Most of the acids you’re likely to encounter fall somewhere
in the region of 0.00001 to 1. Writing this in standard form, we
use negative powers, i.e. 10−6 to 10−1. In this case, the 10−1 is
the more concentrated solution.
It was a Danish chemist named Søren Sørensen who
suggested we use negative logarithms when writing our acid
concentrations, purely because it looks neater. The less
concentrated acid gets a value of:
−Log10 0.00001 = 6
While the more concentrated acid ends up as:
−Log10 1 = 1
He referred to the negative logarithm of a number as the
“potenz” of it, which really means “the power you have to
raise the number ten to.” And so we define the pH scale as:
pH = −Log10 (concentration of hydrogen ions in a liter)
On this scale, most acids fall between 1 and 6 but the scale can
go in either direction. An acid with a concentration of 10
would have a pH of 0, while an acid with a concentration of
100 would be pH −1.
The pKa scale works in exactly the same way and uses the
same system. Only this time, rather than measuring the
concentration of hydrogens, we’re measuring the strength of
an acid, i.e. how willing it is to release a proton into solution.
The best way to express this is to say what fraction of original
acid ends up dissociating.
Suppose you have 100 acid molecules and only one of
them splits. We would say the strength for this acid is 1
percent. For reasons that we won’t get into (an appendix of an
appendix is a bit silly), we express the strength of an acid as a
fraction called the Ka. And, once again, these numbers are
typically extremely small.
Only a tiny amount of hydrogens have the gumption to
dissociate, so we get Ka values with high negative numbers.
Using the p method, we take the negative logarithm of the Ka
(how strong it is) and voilà, the result is our pKa scale.
APPENDIX VI

Groups of the Periodic Table


Going from left to right across the periodic table, several of
the columns (groups) have names that are largely historical.
Groups 3 to 10 are named after the top element in the group
e.g. group 10 is called “the nickel group,” but all other
columns have informal monikers.
Group 1 Alkali metals
Group 2 Alkaline earth metals
Group 3 Scandium group
Group 4 Titanium group
Group 5 Vanadium group
Group 6 Chromium group
Group 7 Manganese group
Group 8 Iron group
Group 9 Cobalt group
Group 10 Nickel group
Group 11 Coinage metals
Group 12 Volatile metals
Group 13 Icosagens
Group 14 Crystallogens
Group 15 Pnictogens
Group 16 Chalcogens
Group 17 Halogens
Group 18 Noble gases
Acknowledgments
One thing I’ve learned during the course of writing my first
book is that while my name appears on the cover, what you
read is a collaboration of many minds. I’d like to thank several
of them.
First and foremost, the students of Northgate High School,
for giving me a reason to get up in the morning and helping
me get the whole thing off the ground. This book wouldn’t
exist without them. Rock and roll, guys. Rock and roll.
Huge and heartfelt thanks to my agent Jen Christie for
taking a chance on an awkward writer, being patient when I
was difficult, and for guiding me through the insane world of
publishing and marketing—something far more complex than
science.
Duncan Proudfoot from Little, Brown Book Group is
wonderful and I want to thank him for immediately “getting”
what I was trying to do. I really hope the book doesn’t let him
down.
In terms of the book’s content, the person I need to thank
most is Ella Catherall who edited every page, fact-checked the
science, and corrected all 385 errors. If the book is any good,
it’s down to her.
Thank you to the Science department at Northgate (there’s
about twenty staff so I can’t namecheck everyone) for all the
tolerance they have shown me while I’ve learned to be a
teacher. It’s a privilege working with them. And in particular, a
very special thanks to Hazel and David—not just for advice,
but for their inspiration and friendship.
I need to thank the great Seishi Shimizu for teaching me
how to think and write like a scientist. He was a great mentor
and it was an honor being his student.
Thank you to Karl Dixon for reading and critiquing the
manuscript, making me laugh when I stopped believing it was
any good, and for being the Sherlock Holmes to my Watson
these many, many years.
A huge thank you to Mandalyn King for always giving it to
me straight and being pretty much amazing. I respect her
opinion more than she knows.
From an earlier time in my life, I want to thank Mr. Evans
for being that teacher, and John Miller for persuading me to
become one myself.
I obviously need to thank my wife who is probably the
most patient person in the world. Thanks for letting me spend
so much time doing this.
And finally, I want to thank my father, who taught me the
importance of asking questions and turned me into a scientist.
Notes
INTRODUCTION: A RECIPE FOR REALITY
1. R. W. Sterner, J. J. Elser, Ecological Stoichiometry: The Biology of Elements
from Molecules to the Biosphere (Princeton, NJ: Princeton University Press,
2002).

CHAPTER 1: FLAME CHASERS


1. H. Krug, O. Ruff, “Uber ein neues chlorfuorid ClF3,” Zeitschrift für
anorganische und allgemeine Chemie, vol. 190, no. 1 (1930), pp. 270–76.1.
2. “Compound summary for CID 24627,” Open Chemistry Database. Available
from:
https://pubchem.ncbi.nlm.nih.gov/compound/chlorine_trifluoride#section=Top
(accessed August 18, 2017).
3. J. D. Clark, Ignition! An Informal History of Rocket Propellants (New
Brunswick, NJ: Rutgers University Press, 1972).
4. “Eastern Germany 2004,” Bunker Tours. Available from:
http://www.bunkertours.co.uk/germany_2004.htm (accessed August 18, 2017).
5. Diogenes Laertius, The Lives and Opinions of Eminent Philosophers, Vol. II,
Books 6–10, trans. R. D. Hicks (Cambridge, MA: Harvard University Press,
1925).
6. “Protactinium,” Encyclopedia. Available from: http://www.encyclo-
pedia.com/science-and-technology/chemistry/compounds-and-
elements/protactinium (accessed August 18, 2017).
7. J. Emsley, The Shocking History of Phosphorus: A Biography of the Devil’s
Element (London: Macmillan, 2000).
8. H. M. Leicester, H. S. Klickstein, A Source Book in Chemistry 1400–1900
(Cambridge, MA: Harvard University Press, 1952).
9. H. Muir, Eureka: Science’s Greatest Thinkers and Their Key Breakthroughs
(London: Quercus, 2012).
10. Muir, Eureka.
11. M. Sędziwój, “Letters of Michael Sendivogius to the RoseyCrusian Society,”
Epistle 54 (January 12, 1647), The Masonic High Council the Mother High
Council. Available from: http://rgle.org.uk/Letters_Sendivogius.htm (accessed
October 8, 2017).
12. I. Asimov, Breakthroughs in Science (Boston, MA: Houghton Mifflin, 1960).
13. R. Harré, Great Scientific Experiments: Twenty Experiments that Changed Our
View of the World (Oxford: Phaidon, 1981).
14. I. Asimov, Words of Science (London: Harrap, 1974).
15. Isaiah 54:11.
16. “Periodic table—lithium,” Royal Society of Chemistry. Available from:
http://www.rsc.org/periodic-table/element/3/lithium (accessed August 18,
2017).
17. B. C. Gibb, “Hard-luck Scheele,” Nature Chemistry, vol. 7 (2015), pp. 855–6.
18. Leicester and Klickstein, A Source Book in Chemistry.

CHAPTER 2: UNCUTTABLE
1. The Core (2003), dir. Jon Amiel, Paramount Pictures.
2. T. Irifune et al., “Ultrahard polycrystalline diamond from graphite,” Nature, vol.
421 (2003), pp. 599–600.
3. David Robson, “How to make a diamond from scratch with peanut butter,” BBC
(November 7, 2014). Available from:
http://www.bbc.com/future/story/20141106-the-man-who-makes-diamonds
(accessed August 18, 2017).
4. B. Russell, History of Western Philosophy (Oxford: Routledge Classics, 2004).
5. D. Hurd, J. Kipling, The Origins and Growth of Physical Science (London:
Penguin, 1958).
6. J. Dalton, A New System of Chemical Philosophy (London: R. Bickerstaff,
1808).
7. W. L. Masterson, C. N. Hurley, Chemistry: Principles and Reactions (Boston,
MA: Cengage Learning, 2012).
8. R. Harré, Great Scientific Experiments: Twenty Experiments that Changed Our
View of the World (Oxford: Phaidon, 1981).
9. A. Einstein, “Über die von der molekularkinetischen Theorie der Wärme
geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten Teilchen,”
Annalen der Physik, vol. 322 (1905), pp. 549–60.

CHAPTER 3: THE MACHINE GUN AND THE PUDDING


1. “A Boy and His Atom: The World’s Smallest Movie,” IBM Research. Available
from: http://www.research.ibm.com/articles/madewithatoms.shtml (accessed
August 18, 2017).
2. E. T. Whittaker, A History of Theories of the Aether and Electricity (Harlow:
Longman, Green & Co, 1951).
3. E. Rutherford, Nobel Lectures: Chemistry 1901–1921 (Amsterdam: Elsevier
Publishing, 1966).
4. H. C. von Bayer, Taming the Atom: The Emergence of the Visible Microworld
(New York: Random House, 1992).
5. R. W. Chabay, B. A. Sherwood, Matter & Interactions, third edition (Hoboken,
NJ: Wiley, 2002).
6. Man of Steel (2013), dir. Zak Snyder, Warner Bros.
7. H. P. Lovecraft, The Dunwich Horror and Other Stories (London: Pocket
Penguin Classics, 2010).
8. Superman Returns (2006), dir. Bryan Singer, Warner Bros; P. S. Whitfield et al.,
“LiNaSiB3O7(OH)—novel structure of the new borosilicate mineral jadarite
determined from laboratory powder diffraction data,” Acta Crystallographica
Section B, vol. 63, no. 3 (2007), pp. 396–401.

CHAPTER 4: WHERE DO ATOMS COME FROM?


1. “The coldest place in the world,” NASA (December 10, 2013). Available from:
https://science.nasa.gov/science-news/science-at-nasa/2013/09dec_coldspot
(accessed August 18, 2017).
2. R. Sahai et al., “The coldest place in the Universe: Probing the ultra-cold
outflow and dusty disk in the Boomerang Nebula,” The Astrophysical Journal,
vol. 841, no. 2 (2017).
3. J. W. Park et al., “Ultracold dipolar gas of fermionic Na23K40 molecules in
their absolute ground state,” Physical Review Letters, vol. 114 (2015).
4. Plato, Theaetetus, trans. J. McDowell (Oxford: Oxford University Press, 1999).
5. B. Russell, History of Western Philosophy (Oxford: Routledge Classics, 2004).
6. G. Dixon, P. Parsons, The Periodic Table: A Field Guide to the Elements
(London: Quercus, 2013).
7. H. Aldersey-Williams, Periodic Tales: The Curious Lives of the Elements
(London: Viking, 2011).
8. C. Payne-Gaposchkin, Cecilia Payne-Gaposchkin: An Autobiography and
Other Recollections (Cambridge: Cambridge University Press, 1996).
9. “Cecilia Payne-Gaposchkin,” Encylopædia Britannica. Available from:
https://www.britannica.com/biography/Cecilia-Payne-Gaposchkin (accessed
August 18, 2017).
10. “The early universe,” CERN. Available from:
https://home.cern/about/physics/early-universe (accessed August 18, 2017).

CHAPTER 5: BLOCK BY BLOCK


1. “The Scoville Unit,” Jalapeño Madness. Available from:
http://www.jalapenomadness.com/jalapeno_scoville_units.html (accessed
August 18, 2017).
2. “‘World’s hottest’ chilli pepper grown in St Asaph,” BBC News (May 17, 2017).
Available from: http://www.bbc.com/news/uk-wales-north-east-wales-
39946962 (accessed August 18, 2017).
3. A. Szallasi, P. M. Blumberg, “Resiniferatoxin, a phorbol-related diterpene, acts
as an ultrapotent analog of capsaicin, the irritant constituent in red pepper,”
Neuroscience, vol. 30, no. 2 (1989), pp. 515–20.
4. “How we taste,” Technology Review (April 2004). Available from:
https://www.heise.de/tr/artikel/Wie-wir-schmecken-404206.html (accessed
August 18, 2017).
5. “Vantablack,” Surrey Nanosystems. Available from:
https://www.surreynanosystems.com/vantablack (accessed August 18, 2017).
6. J. Clayden, N. Greeves, S. Warren, Organic Chemistry, second edition (Oxford:
Oxford University Press, 2012); “4 workers killed at DuPont Chemical plant,”
Scientific American (November 18, 2014). Available from:
https://www.scientificamerican.com/article/4-7.workers-killed-at-dupont-
chemical-plant (accessed August 18, 2017).
7. B. Russell, History of Western Philosophy (Oxford: Routledge Classics, 2004).
8. B. Pennington, “The death of Pythagoras,” Philosophy Now, no. 121 (2017).
9. Russell, History of Western Philosophy.
10. A. Lavoisier, Traite Elementaire de Chemie (Paris: Cuchet, 1789).
11. E. Scerri, The Periodic Table: Its Story and Its Significance (Oxford: Oxford
University Press, 2006).
12. E. Scerri, The Periodic Table: A Very Short Introduction (Oxford: Oxford
University Press, 2011).
13. J. E. Jorpes, Jac. Berzelius: His Life and Work (Stockholm: Royal Swedish
Academy of Science, 1966).
14. J. A. R. Newlands, On the Discovery of the Periodic Law: and On Relations of
the Atomic Weights (London: E. & F. N. Spon, 1884).
15. M. D. Gordin, A Well-Ordered Thing: Dmitrii Mendeleev and the Shadow of the
Periodic Table (New York: Basic Books, 2004).
16. “Periodic Law,” Mendeleev. Available from:
http://www.mendeleev.nw.ru/period_law/ver_trif.html (accessed August 18,
2017).

CHAPTER 6: QUANTUM MECHANICS SAVES THE DAY


1. A. Werner, “Beitrag zum Ausbau des periodischen systems,” Berichte der
deutschen chemischen Geselkchaft, vol. 38 (1905), pp. 914–21.
2. G. Seaborg, “Priestley Medal Address—The Periodic Table: Tortuous Path to
Man-Made Elements” (April 16, 1979), reprinted in G. Seaborg, Modern
Alchemy: Selected Papers of Glenn Seaborg Vol. 2 (Singapore: World Scientific
Publishing Co., 1994).
3. H. E. White, Introduction to Atomic Spectra (New York: McGraw-Hill, 1934).
4. E. H. Riesenfeld, Practical Inorganic Chemistry, reprint of the 1943 edition
(Barcelona: Labour, 1950).
5. Seaborg, “Priestley Medal Address.”

CHAPTER 7: THINGS THAT GO BOOM


1. T. M. Klapötke et al., “New azidotetrazoles: Structurally interesting and
extremely sensitive,” Chemistry—An Asian Journal, vol. 7, no. 1 (2012), pp.
214–24.
2. “Alfred Nobel,” Encylopædia Britannica. Available from:
https://www.britannica.com/biography/Alfred-Nobel (accessed August 18,
2017); E. J. Sirleaf, “Alfred Nobel’s legacy to women,” New York Times
(December 12, 2011).
3. “Alfred Nobel’s fortune,” Nobel Peace Prize. Available from:
https://www.nobelpeaceprize.org/History/Alfred-Nobel-s-fortune (accessed
August 18, 2017).
4. J. Janes, Documents which Changed the Way We Live (Lanham, MD: Rowman
& Littlefield, 2017).
5. K. Fant, Alfred Nobel: A Biography (New York: Arcade Publishing, 2014).

CHAPTER 8: THE ALCHEMIST’S DREAM


1. “Sotheby’s sells record $71 million diamond to Chow Tai Fook,” Bloomberg
(April 4, 2017). Available from:
https://www.bloomberg.com/news/articles/2017-04-04/sotheby-s-sets-world-
record-selling-71-million-pink-diamond (accessed August 18, 2017).
2. R. Kurin, Hope Diamond: The Legendary History of a Cursed Gem (New York:
HarperCollins, 2007).
3. “Plutonium certified reference materials price list,” US Department of Energy—
Office of Science. Available from: https://science.energy.gov/nbl/certified-
reference-materials/prices-and-certificates/plutonium-certified-reference-
materials-price-list (accessed August 18, 2017).
4. “Californium price,” Metalary. Available from: https://www.meta-
lary.com/californium-price (accessed August 18, 2017).
5. G. D. Hedesan, An Alchemical Quest for Universal Knowledge: The “Christian
Philosophy” of Jan Baptist Van Helmont 1579–1644 (Oxford: Routledge,
2016).
6. R. Patai, The Jewish Alchemists: A History and Source Book (Princeton, NJ:
Princeton University Press, 1994).
7. B. Jonson, The Alchemist (1610). Available from: http://www.public-
library.uk/ebooks/14/35.pdf (accessed August 18, 2017).
8. S. Lee, S. Ditko, Amazing Fantasy, no. 15 (August 15, 1962); S. Lee, J. Kirby,
The Incredible Hulk, no. 1 (May 1, 1962); S. Lee, J. Kirby, The Fantastic Four,
no. 1 (November 1, 1961); S. Lee, B. Everett, Daredevil, no. 1 (April 1, 1964);
C. Claremont, J. Byrne, X-Men, no. 137 (September 1, 1980), and Phoenix: The
Untold Story (April 1, 1984).
9. Godzilla (1954), dir. Ishiro Honda, Toho Co. Ltd.
10. C. Patterson, “Age of meteorites and the earth,” Geochimica et Cosmochimica
Acta, vol. 10, no. 4 (1956), pp. 230–7.
11. E. Rutherford, “The Collision of Alpha-particles with Light Atoms,”
Philosophical Magazine, vol. 37 (1919).
12. “Public ignorant about radiation dose of mammograph,” Medscape (May 12,
2014). Available from: http://www.medscape.com/viewarticle/824999 (accessed
August 18, 2017).
13. Gary Mansfield, “Banana equivalent dose” (March 7, 1995). Available from:
http://health.phys.iit.edu/extended_archive/9503/msg00074.html (accessed
August 18, 2017).
14. D. R. Corson, K. R. MacKenzie, E. Serge, “Artificially radioactive element
85,” Physical Review, vol. 58, no. 8 (1940), pp. 672–8.
15. Iron Man 2 (2010), dir. Jon Favreau, Paramount Pictures.
16. “Edwin M. McMillan—facts,” Nobel Prize. Available from:
http://www.nobelprize.org/nobel_prizes/chemistry/laureates/1951/mcmillan-
facts.html (accessed August 18, 2017).
17. R. M. Shoch, Case Studies in Environmental Science (Eagan, MN: West
Publishing Co., 1996).
18. “Americium,” ACS Publications. Available from:
http://pubs.acs.org/cen/80th/print/americiumprint.html (accessed August 18,
2017).
19. “IUPAC announces the names of the elements 113, 115, 117 and 118,”
International Union of Pure and Applied Chemistry (November 30, 2016.
Available from: https://iupac.org/iupac-announces-the-names-of-the-elements-
113-115-117-and-118 (accessed August 18, 2017).
20. J. Emsley, Nature’s Building Blocks: An A–Z Guide to the Elements (Oxford:
Oxford University Press, 2001).

CHAPTER 9: LEFTISTS
1. A. K. Geim, M. V. Berry, “Of flying frogs and levitrons,” European Journal of
Physics, vol. 18, no. 4 (1997), pp. 307–13.
2. K. S. Novoselov et al., “Electric firled effect in atomically thin carbon films,”
Science, vol. 306, no. 5696 (2004), pp. 666–9.
3. “How strong is graphene?,” University of Manchester. Available from:
http://www.graphene.manchester.ac.uk/discover/video-gallery/what-is-
graphene/how-strong-is-graphene (accessed August 18, 2017); J. Abraham et
al., “Tunable sieving of ions using graphene oxide membranes,” Nature
Nanotechnology, no. 12 (2017), pp. 546–50.
4. “Properties of stainless steel, metals and other conductive materials,” TibTech
Innovations. Available from: http://www.tibtech.com/conductivity.php
(accessed August 18, 2017); “Understanding graphene,” Graphenea. Available
from: https://www.graphenea.com/pages/graphene (accessed August 18, 2017).
5. J. Romer, A History of Ancient Egypt: From the First Farmers to the Great
Pyramid (New York: Thomas Dunne Books, 2013).
6. J. Levy, Scientific Feuds: From Galileo to the Human Genome Project
(London: New Holland Publishers, 2010).
7. S. Gray, “An account of some new electrical experiments,” Philosophical
Transactions of the Royal Society of London, vols 31–3 (1708).
8. D. S. Lemons, Drawing Physics: 2,600 Years of Discovery from Thales to Higgs
(Cambridge, MA: MIT Press, 2017).
9. P. Bertucci, “Sparks in the dark: The attraction of electricity in the eighteenth
century,” Endeavour, vol. 31, no. 3 (2007).
10. C. Brandon, The Electric Chair: An Unnatural American History (Jefferson,
NC: McFarland, 1999).
11. Levy, Scientific Feuds; Electrocuting an Elephant (1903)—WARNING: Viewer
Discretion—Disturbing footage—Thomas Edison, Change Before Going
Productions (January 16, 2014). Available from:
https://www.youtube.com/watch?v=NoKi4coyFw0 (accessed August 18, 2017).
12. C. S. Combs, Deathwatch: American Film, Technology and the End of Life
(New York: Columbia University Press, 2014).
13. M. S. Rosenwald, “‘Great God, he is alive!’ The first man executed by electric
chair died slower than Thomas Edison expected,” Washington Post (April 28,
2017).

CHAPTER 10: ACIDS, CRYSTALS, AND LIGHT


1. D. Wilson, A History of British Serial Killing (London: Sphere, 2011); M.
Whittington-Egan, R. Whittington-Egan, Murder on File: The World’s Most
Notorious Killers (Castle Douglas: Neil Wilson Publishing, 2005).
2. D. H. Ripin, D. A. Evans, “pKas of inorganic and oxo-acids,” The Evans Group.
Available from: http://evans.rc.fas.harvard.edu/pdf/evans_pKa_table.pdf
(accessed August 18, 2017).
3. Ripin, Evans, “pKas of inorganic and oxo-acids”; G. T. Cheek,
“Electrochemical studies of the Fries rearrangement in ionic liquids,”
Electrochemical Society Transactions, vol. 16, no. 49 (2009), pp. 541–4.
4. G. A. Olah, “My search for carbocatins and their role in chemistry,” Nobel
Lecture (December 8, 1994).
5. To illustrate this point, the author has taken the claim from the article on
superacids from Wikipedia. Available from:
https://en.wikipedia.org/wiki/Superacid (accessed August 18, 2017) to illustrate
this point. Wikipedia cites G. A. Olah, “Crossing conventional boundaries in
half a century of research,” Journal of Organic Chemistry, vol. 70, no. 7 (2005),
pp. 2413–29, for the claim that fluoroantimonic acid is 1016 times stronger than
sulfuric known to have a pKa of −3, giving a pKa of −19.
6. T. R. Hogness, E. G. Lunn, “The ionisation of hydrogen by electron impact as
interpreted by positive ray analysis,” Physical Review, vol. 21, no. 1 (1925), pp.
44–55.
7. This number is calculated from generating a Born-Haber cycle via: S. Lias et
al., “Evaluated gas phase basicities and proton affinities of molecules: Heats of
formation of protonated molecules,” Journal of Physical and Chemical
Reference Data, vol. 13, no. 3 (1984), p. 695, and assumes that the HHe+ ion
has a similar solubility to a lithium ion, which has comparable size. If we
assume a free energy change of dissociation to be −360 kJmol-1 then at standard
temperature and pressure we can invoke G = −RT lnKa. Taking −360/(0.008314
× 273) we obtain 158.6 = lnKa and therefore a Ka to have value of 4.15 x1068.
Taking the negative logarithm of this number yields −68.6, which the author has
rounded to −69.
8. “Strange but true: Superfluid helium can climb walls,” Scientific American
(February 20, 2009). Available from:
https://www.scientificamerican.com/article/superfluid-can-climb-walls
(accessed August 18, 2017).

CHAPTER 11: IT’S ALIVE, IT’S ALIVE!


1. A. C. Nathwani et al., “Polonium-210 poisoning: a first-hand account,” The
Lancet, vol. 388, no. 10049 (2016), pp. 1075–80.
2. R. H. Adamson, “The acute lethal dose 50 (LD50) of caffeine in albino rats,”
Regulatory Toxicology and Pharmacology, vol. 80 (2016), pp. 274–6.
3. E. Welsome, The Plutonium Files: America’s Secret Medical Experiments in the
Cold War (New York: The Dial Press, 1999).
4. Lead: K. Sujatha et al., “Lead acetate induced neurotoxicity in Wistar albino
rats: A pathological, immunological, and ultrastructural study,” Journal of
Pharma and Bio Science, no. 2 (2011), pp. 459–62. Note: this assumes lead
acetate. Thallium: Agency for Toxic Substances and Disease Registry,
Toxicological Profile for Thallium (Atlanta, GA: Agency for Toxic Substances
and Disease Registry, 1992). Available from:
https://www.atsdr.cdc.gov/ToxProfiles/tp.asp?id=309&tid=49 (accessed August
18, 2017). Note: this assumes thallium acetate for fair comparison with lead.
Arsenic: H. Marquardt et al., Toxicology (Cambridge, MA: Academic Press,
1999). Phosphorus: Agency for Toxic Substances and Disease Registry,
Toxicological Profile for White Phosphorus (Atlanta, GA: Agency for Toxic
Substances and Disease Registry, 1997) Available from:
https://www.atsdr.cdc.gov/toxprofiles/tp103-c2.pdf (accessed August 18, 2017).
Note: the value quoted seems to come from C. C. Lee, Mammalian Toxicity of
Munition compounds. Phase I: Acute Oral Toxicity, Primary Skin and Eye
Irritation, Dermal Sensitization, and Disposition and Metabolism, Report No. 1,
AD B011150 (Kansas City, MO: Midwest Research Institute, 1975).
5. S. Ela, “Experimental study of toxic properties of dimethylcadmium,” Gigiena
Truda i Professional’nye Zabolevaniya, no. 6 (1991), pp. 14–17.
6. J. R. Barash, S. S. Arnon, “A novel strain of clostridium botulinum that
produces Type B and Type H botulinum toxins,” The Journal of Infectious
Diseases, vol. 29, no. 2 (2014), pp. 183–91.
7. “Botox OnabotuliniumtoxinA,” Botox. Available from: http://www.botox.com
(accessed August 18, 2017).
8. C. H. Mayo, interview given in Northwestern Health Journal (December 1924).
9. V. Busacchi, “Vincenzo Menghini and the discovery of iron in the blood,”
Bullettino delle science mediche, vol. 130, no. 2 (1958), pp. 202–5.
10. E. Kinne-Saffran, R. K. Kinne, “Vitalism and synthesis of urea. From Friedrich
Wöhler to Hans A. Krebs,” American Journal of Nephrology, vol. 19, no. 2
(1999), pp. 290–4.
11. K. H. Antman, “Introduction: The history of arsenic trioxide in cancer therapy,”
The Oncologist, vol. 6, no. 2 (2001), pp. 1–2.
12. N. C. Lloyd, “The composition of Ehrlich’s salvarsan: Resolution of a century-
old debate’, Angewandte Chemie, vol. 44, no. 6 (2005), pp. 941–4.
13. H. P. Chauhan, “Synthesis, spectroscopic characterization and antibacterial
activity of antimony(III)bis(dialkyldithiocarbamato) alkyldithiocarbonates,”
Spectrochimica Acta. Part A, vol. 81, no. 1 (2011), pp. 417–23; “Education in
Chemistry—Cerium,” Royal Society of Chemistry. Available from:
https://eic.rsc.org/elements/cerium/2020005.article (accessed August 18, 2017).
14. “Getting a tiny bit of this element on your skin will make you reek of garlic for
weeks,” io9 (August 13, 2015). Available from: http://io9.gizmodo.com/getting-
a-tiny-bit-of-this-element-on-your-skin-will-ma-1723949124 (accessed August
18, 2017).
15. R. Hambrecht et al., “Managing your angina symptoms with nitroglycerin,”
Circulation, no. 127 (2013).
16. V. S. Ramachandran, Encyclopedia of the Human Brain (Cambridge, MA:
Academic Press, 2002).
17. T. Bartholin, Historiarum anatomicarum rariorum centuria I et II (1654).
Available from: https://books.google.nl/books?id=NT-
LAd44hZ4UC&printsec=frontcover&dq=%22Historiarum+an-
atomicarum+rariorum+centuria+I%22&hl=en&sa=X&ei=6T-
MLVagK09SgBJvGgaAH&redir_esc=y#v=onepage&q=%22Historiarum%20an
atomicarum%20rariorum%20centuria%20I%22&f=false (accessed August 18,
2017).
18. “New light on human torch mystery,” BBC News (31 August 1998). Available
from: http://news.bbc.co.uk/2/hi/uk_news/158853.stm (accessed August 18,
2017).
19. M. Harrison, Fire from Heaven: A Study of Spontaneous Combustion in Human
Beings (London: Skoob Books, 1990).
20. “Cause of fire killing woman still mystery,” St. Petersburg Times, Section 2
(July 4, 1951). Available from: https://news.google.com/newspapers?
nid=888&dat=19510704&id=rwRZAAAAIBAJ&sjid=lE8DAAAAIBAJ&pg=3
085,1265930&hl=en (accessed August 18, 2017).
21. Garth Haslam, “1951, July 1: Mary Reeser’s fiery death,” Anomalies: The
Strange and Unexplained. Available from:
http://anomalyinfo.com/Stories/1951-july-1-mary-reesers-strange-death
(accessed August 18, 2017).
22. L. E. Arnold, Ablaze! The Mysterious Fires of Spontaneous Human Combustion
(New York: M. Evans and Co., 1995).
23. J. Randles, P. Hough, Spontaneous Human Combustion (London: Robert Hale
Ltd, 2007).
24. G. Whitley, “Garston Church” (1867–74), Speke Archive Online. Available
from: http://spekearchiveonline.co.uk/garston_church.htm (accessed August 18,
2017).
25. G. Gassmann, D. Glindemann, “Phosphane (PH3) in the biosphere,”
Angewandte Chemie, vol. 32, no. 5 (1993), pp. 761–3.

CHAPTER 12: NINE ELEMENTS THAT CHANGED THE WORLD (AND


ONE THAT DIDN’T)
1. “Pitch Drop Demonstration,” National Museums Scotland. Available from:
https://www.nms.ac.uk/explore-our-collections/stories/science-and-
technology/made-in-scotland-changing-the-world/scottish-science-
innovations/pitch-drop-demonstration (accessed September 9, 2017).
2. “Bart the Lover,” The Simpsons, season 3, episode 16, dir. Carlos Baeza
(original airdate February 13, 1992).
3. J. Emsley, Nature’s Building Blocks: An A–Z Guide to the Elements (Oxford:
Oxford University Press, 2001).
4. E. Barrett, J. Mingo, Not Another Apple for the Teacher: Hundreds of
Fascinating Facts from the World of Education (Newburyport, MA: Conari
Press, 2002).
5. “The story of how the tin can nearly wasn’t,” BBC News (April 21, 2013).
Available from: http://www.bbc.com/news/magazine-21689069 (accessed
August 18, 2017).
6. Adapted from “Gold fun facts,” American Museum of Natural History.
Available from: http://www.amnh.org/exhibitions/gold/eureka/gold-fun-facts
(accessed August 18, 2017).
7. Adapted from R. O’Connell et al., GFMS Gold Survey 2016 (New York:
Thomson Reuters, 2016).
8. “The history of money,” The Mint of Finland. Available from:
https://www.suomenrahapaja.fi/eng/about_money/the_history_of_money
(accessed August 18, 2017); E. M. Green, Lady Midrash: Poems Reclaiming
the Voices of Biblical Women (Eugene, OR: Wipf and Stock, 2016).
9. J. O. Nriagu, “Saturnine gout among Roman aristocrats—did lead poisoning
contribute to the fall of the empire?,” New England Journal of Medicine, no.
308 (1983), pp. 660–3.
10. H. Needleman, “Low level lead exposure: History and discovery,” Annals of
Epidemiology, vol. 19, no. 4 (2009), pp. 235–8; H. Delile et al., “Lead in
ancient Rome’s city waters,” PNAS, vol. 11, no. 18 (2014), pp. 6594–9.
11. D. Childress, Johannes Gutenberg and the Printing Press (Minneapolis, MN:
Twenty First Century Books, 2008).
12. A. Gallop, “Mortality improvements and evolution of life expectancies,”
Actuary, Pensions Policy, Demography and Statistics (London: Government
Actuary’s Department, 2006).
13. G. W. Beardsley, “The 1832 cholera epidemic,” Early America Review, vol. 4,
no. 1 (2000).
14. “Measles” and “Frequently asked questions and answers on smallpox,” World
Health Organization. Available from:
http://www.who.int/mediacentre/factsheets/fs286/en/ and available from:
http://www.who.int/csr/disease/smallpox/faq/en (accessed August 18, 2017).
15. D. Charles, Between Genius and Genocide: The Tragedy of Fritz Haber, Father
of Chemical Warfare (London: Jonathan Cape, 2005).
16. “How much water does the average person use at home per day?,” United
States Geological Survey. Available from: https://water.usgs.gov/edu/qa-home-
percapita.html (accessed August 18, 2017).
17. T. P. Garrett, “The wonderful development of photography,” The Art World, vol.
2, no. 5 (1917), pp. 489–91.
18. Stephen Herbert, “Wordsworth Donisthorpe,” Who’s Who of Victorian Cinema
(2000). Available from: http://www.victorian-cinema.net/ donisthorpe (accessed
August 18, 2017).
19. J. Watson, DNA: The Secret of Life (London: Arrow Books, 2003).
20. “U.S. Nuclear Weapons Capability,” 2017 Index of U.S. Military Strength
(2017). Available from: http://index.heritage.org/military/2017/assessments/us-
military-power/u-s-nuclear-weapons-capability (accessed August 18, 2017).
21. J. Robert Oppenheimer: “I am become death, the destroyer of worlds,”
Plenilune pictures, (August 6, 2011). Available from:
https://www.youtube.com/watch?v=lb13ynu3Iac (accessed August 18, 2017).
22. J. N. Shurkin, Broken Genius: The Rise and Fall of William Shockley, Creator
of the Electronic Age (London: Macmillan, 2006).
23. A. Usanov et al., Coltan, Congo & Conflict: Polinares Case Study (The Hague:
The Hague Centre for Strategic Studies, no. 21.05.13, 2013); E. Sutherland,
“Coltan, the Congo and your cell phone: the connection between your mobile
phone and human rights abuses in Africa,” MIT (2016). Available from:
http://web.mit.edu/12.000/www/m2016/pdf/coltan.pdf (accessed August 18,
2017).
24. D. Grossmann, C. Ganz, P. Russell, Zeppelin Hindenburg: An Illustrated
History of LZ-129 (Stroud: The History Press, 2017).
25. O. A. Hurricane et al., “Fuel gain exceeding unity in an inertially confined
fusion implosion,” Nature, vol. 506 (2014), pp. 343–7.
Index
3M
ability
absolute zero
acids
actinium
adenosine triphosphate (ATP)
age-related diseases
air
al-Sufi, Abd al-Rahman
alchemy
alkahest
allotropes
alloys
alpha particles
alpha decay
aluminum
aluminum oxide
Amagat, Émile
American Chemical Society
americum
ammonia
amperes
amygdalin
ancient Egypt
ancient Greece
Andromeda constellation
anti-bacterials
anti-neutrinos
antimony
argon
argyria
Aristotle
arsenic
astatine
astrochemistry
atomic bonds
Atomic Energy Authority
atomic hypothesis
atomic sponge
atoms
electrical
and elemental differences
instability
interior
large
non-metals
nucleus
orbital arrangement
origins
and proton numbers
and the Schrödinger wave equation
splitting
and star formation
of unstable substances
uranium
ATP see adenosine triphosphate
azidoazide azide
Babbage’s difference engine
Banana Equivalent Dose (BED)
barium
Becquerel, Henri
bell metal
berkelium
beryllium
beta radiation
beta decay
big bang
biology
bismuth
black holes, super-massive
bohrium
bone
Boomerang Nebula
borax crystals
boron
bosons
bottom-up approaches
botulinum toxin
Boyle, Robert
brain
Brandt, Hennig
bridges
bromine
bronze
Bronze Age
Brownian motion
Bunsen, Robert
cadmium
cesium
caffeine
calcium
californium
calxes
cancer
carbon
conductivity
and diamonds
as fuel
and graphite
of human DNA
and making elements
and nitroglycerine
and sugar
carbon charcoal
carbon dioxide (CO2)

carbon monoxide
catfish
causation
Cavendish, Henry
Celsius scale
centigrade
cerium
Chadwick, James
chalk
charcoal
charge
Charles’s law
chemical bonds
chemical reactions
chemicals
classification
see also specific chemicals
chemotherapy
child mortality figures
China
chloride
chlorination
chlorine
chlorine gas
chlorine trifluoride (CIF3)
cholera
Chow Tai Fook Enterprises
chromic acid
chromium
climate change, human-made
Clostridium botulinum
coal
cobalt
Cold Atom Laboratory
Cold War
compounds
conductors
copernicium
copper
corrosive resistance
corundum
cosmic rays
crematoria
critical mass
crust
crystals
see also specific crystals
Curie, Marie and Pierre
curium
current
cyanide
cytochrome c oxidase
Dalton, John
darmstadtium
Davy, Humphry
de Haan, John
Democritus
density
deoxy ribonucleic acid (DNA)
“dephlogisticated air”
diamond
dimethyl cadmium
diphosphane (P2H4)
Döbereiner, Johann
Donkin, Bryan
dubnium
dynamite
dysprosium
E = mc2
Earth
earths
Edison, Thomas
Einstein, Albert
einsteinium
electric chair
electric current
electric shocks
electrical atoms
electrical conductivity
electrical fields
electrical resistance
electricity
electron orbitals (electron clouds)
and chemical reactions
effect of increasing numbers of protons on
and electricity
large
and mercury
and metals
predicting the position of
electron sea
electrons
and acids
and carbon
charge
and chemical reactions
and electricity
and fluorine
and molecular orbitals
movement
and new element formation
of non-metals
and the periodic table
predicting the position of
and the Schrödinger wave equation
and static electricity
substructure
and turning neutrons into protons
and W- bosons
wavy movement
elements
abundant
and atoms
changing one into another
classification
construction of artificial new
cyclical pattern
discovery
Empedocles on
formation in stars
hidden
most crucial to human development
most expensive
naming
poisonous
and the Schrödinger wave equation
and the shape of the periodic table
see also specific elements
elixir of life
Empedocles
energy
erbium
essences
ether
europium
execution
explosives
Fahrenheit scale
fat, body
fermium
Feshbach resonance
fire
fireworks
fission
“flammable air”
flammable substances
Flamsteed, John
flavor
flerovium
fluidity
fluoride
fluorine
fluoroantimonic acid
food
force fields
fossil fuels
francium
Fraunhofer, Joseph von
fructose
fundamental particles
fusion
gadolinium
galaxies
galinstan
gallium
gamma rays
gases
compression
inert
noble
Geiger, Hans
Geiger counter
Geim, Andre
germanium
glass
global warming
glucose
glucuronic acid
gluons
God
gods
Godzilla
gold
gold-foil experiments
graphene
graphite
gravity
Gray, Stephen
Group 17/Group 18
gunmetal
gunpowder
Gutenberg, Johannes
H (Hamiltonian)
Haber, Fritz
hafnium
Haigh, John George
hardness
hassium
hat industry
heart
heat
see also fire
heat sensors
helium
and lasers
liquid
non-reactivity
and the periodic table
helium hydride
Heraclitus
Hertz, Heinrich
Higgs bosons
holmium
Hubble telescope
human body
chemical formula
electrical conductivity
Hurricane, Omar
hydrochloric acid
hydrogen
and acid
of human DNA
and nitroglycerine
and the periodic table
and plastic
and sugar
and water
hydrogen bomb
hydrogen chloride (HCl)
hyperdiamond
hypochlorous acid (HOCL)
i numbers
IBM
indium
infectious diseases
insulators
International Space Station (ISS)
International Union of Pure and Applied Chemistry (IUPAC)
iodine
ions
iridium
Irifune, Tetsuo
iron
Iron Age
Ka
ket vector (|>)
Klapötke, Thomas
Krebs, Hans
krypton
lanthanum
lattices
Lavoisier, Antoine
lawrencium
lead
Lecoq, Paul-Émile
lethal dose (LD50)
life expectancy
LifeGem
light
blue
and electrons
infrared
red
splitting
lightbulbs
lightning
liquids
lithium
Litvinenko, Alexander
livermorium
living things
logarithms, laws of
lugduname
lutetium
McMillan, Edwin
magic acid
“magic numbers”
magnesium
magnetic fields
magnetism
malleability
mammograms
manganese
Manhattan Project
mantle
Marsden, Ernest
medicine
Meitner, Lise
meitnerium
Mendeleev, Dmitri
mendelevium
Menghini, Vincenzo
mercury
metal chemistry
metal oxides
metallurgy
metals
and calxes
conductivity
definition
heavy
Lavoisier on
opacity
reactivity
meteorites
methane (CH4)
microwaves
Milky Way
Mohs value
molecular orbitals
molecules
molybdenum
money
mortality rates
moscovium
musical theory, Western
National Helium Reserve
National Ignition Facility
Nazis
neodymium
neon
neptunium
neutrinos
neutrons
and forging new elements
and nuclear decay
and nuclear stability
and quarks
substructure
turning into protons
uranium
Newlands, John
Newton, Sir Isaac
nickel
Niépce, Joseph “Nicéphore”
nihonium
niobium
nitrogen
nitroglycerine
Nobel, Alfred
Nobel Prizes
nobelium
non-metals
non-reactivity
Novoselov, Kostya
nuclear explosions, of stars
nuclear reactors
nuclear weapons
nuclear-plants, fusion-based
nucleus (atomic)
and chemical reactions
decay
and electricity
and electron orbitals
and forging new elements
fusing
and the Schrödinger wave equation
stability
substructure
unstable nature
uranium
octaves
oganesson
ohms
Oppenheimer, Robert
osmium
oxalic acid
oxides
oxygen
and carbon
and crystals
discovery
and element generation
and fluorine
of human DNA
and metals
and nitroglycerine
and respiration
and sugar
and water
ozone
palladium
panacea
Paracelcus principle
partial differential (∂)
particle theory
particles, speed
pathogens
Pauling, Linus
Payne-Gaposchkin, Cecilia
perchloric acid
periodic table
architectural simplicity
and carbon
completion
and Empedocles
groups
and Lavoisier
and Mendeleev
and Newlands
and quantum mechanics
and Risenfeld
and Seaborg
shape
and silicon
and White
periods
perpetual motion machines
pH scale
philosopher’s stone
phlogiston
phosphane (PH3)

phosphorus
photography
photons
pitch (asphalt)
pKalow scale

Planck’s constant (ħ)


planetary formation
plant life
plastic
platinum
Plum-Pudding Hypothesis, The
plumbing
plutonium
poison
polonium
polymer chemistry
potassium
potassium uranyl sulfate
potassium-40
potenz
power stations
praseodynium
Priestly, John
printing press
prisms
promethium
protactinium
protons
and acids
charge
and making new elements
and nuclear decay
and nuclear stability
and quarks
repulsion
rogue
and the Schrödinger wave equation
substructure
uranium
psi (Ψ)
pyrolusite
Pythagorean order
quantum field theory
quantum leaps
quantum theory
quarks
radiation
exposure measurement
poisoning
sickness
radiograms
radium
radon
rainbows
rat studies
reactivity
relativy theory, special
resiniferatoxin
resistance
respiration
rhenium
rhodium
Richter scale
roentgenium
Romans
Royal Navy
Royal Society
rubber
rubidium
Russia
rust
ruthenium
Rutherford, Ernest
rutherfordium
salt (sodium chloride)
Salvarsan
samarium
sapphire
scandium
scanning tunneling microscope
Scheele, Carl
Scheelite (calcium tungstate)
Schrödinger, Erwin
Schrödinger wave equation
generalized time-dependent version
Schulze, Johann
science
Scoville Heat Units (SHUs)
Seaborg, Glenn
seaborgium
Second World War
selenium
semi-conductors
senses
shells
Shockley, William
siemens per meter
sieverts
silicon
silver
silver fulminate
silver nitrate
silver screen
smelting
Smith, Mike
smoke-detectors
smoking
Sörensen, Sören
sodium
sodium chloride
sodium-potassium
solar systems
solder
solids
space
spectroscopy
spiciness
spontaneous human combustion
stability
Stahl, Georg
stardust
stars
death
formation
super-giant
static electricity
steel
Stone Age
stop-motion films
strontium
sucrose
sugar
sulfate
sulfur
sulfuric acid
Sun
suns
superacids
supernovas
sweet
syphilis
tantalum
technetium
teflon
Telford’s suspension bridge
tellurium
tennessine
terbium
terne
tetrahedral arrays
Thales
thallium
theobromine
thermometers
Thomson, Joseph John “J. J.”
thorium
thulium
tin
titanium
TNT
Tokyo Institute of Technology
tooth enamel
transistors
transmutation
triboelectric effect
triflic acid
Trinity
TRPVI receptors
Tsar Bomba
tungsten
unbihexium
United States
Universe
beginnings of the
coldest place
composition
and gravity
unstable substances
uranium
uranium bombs
uranium-235
urea
urine
USSR
vaccinations
vanadium
vantablack
vinegar
viscosity
vitalism
voltage
volcanoes
von Helmholtz, Hermann
water
wavefunction
Werner, Alfred
Westinghouse, George
White, Harvey
will-o’-the-wisps
Wöhler, Friedrich
X-rays, dental
xenon
ytterbium
yttrium
zeppelins
zinc
zinc sulfide (ZnS)
zirconium
Zwierlein, Martin

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy