Ebook What S Eating The Universe and Other Cosmic Questions 1St Edition Paul Davies Online PDF All Chapter

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

What s Eating the Universe And Other

Cosmic Questions 1st Edition Paul


Davies
Visit to download the full and correct content document:
https://ebookmeta.com/product/what-s-eating-the-universe-and-other-cosmic-questio
ns-1st-edition-paul-davies/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

What s Eating the Universe And Other Cosmic Questions


First Edition Davies

https://ebookmeta.com/product/what-s-eating-the-universe-and-
other-cosmic-questions-first-edition-davies/

What If I m Wrong and Other Key Questions for Decisive


School Leadership 1st Edition Simon Rodberg

https://ebookmeta.com/product/what-if-i-m-wrong-and-other-key-
questions-for-decisive-school-leadership-1st-edition-simon-
rodberg/

The Universe A Biography 2022 Edition Paul Murdin

https://ebookmeta.com/product/the-universe-a-
biography-2022-edition-paul-murdin/

The Romance of Reality How the Universe Organizes


Itself to Create Life Consciousness and Cosmic
Complexity 1st Edition Bobby Azarian

https://ebookmeta.com/product/the-romance-of-reality-how-the-
universe-organizes-itself-to-create-life-consciousness-and-
cosmic-complexity-1st-edition-bobby-azarian/
What s Wrong with China 1st Edition Paul Midler

https://ebookmeta.com/product/what-s-wrong-with-china-1st-
edition-paul-midler/

Lenin s Brain and Other Tales from the Secret Soviet


Archives Paul R Gregory

https://ebookmeta.com/product/lenin-s-brain-and-other-tales-from-
the-secret-soviet-archives-paul-r-gregory/

Business Information Systems 3rd Edition Paul Beynon-


Davies

https://ebookmeta.com/product/business-information-systems-3rd-
edition-paul-beynon-davies/

Stone Heart Deep 9th Edition Paul Bassett Davies

https://ebookmeta.com/product/stone-heart-deep-9th-edition-paul-
bassett-davies/

Introduction to Company Law 3rd Edition Paul Davies

https://ebookmeta.com/product/introduction-to-company-law-3rd-
edition-paul-davies/
Paul Davies

W H AT ’ S E AT I N G T H E U N I V E R S E ?

And Other Cosmic Questions


Contents

Preface

1. Journey from the Edge of Time


2. The Search for the Key to the Universe
3. Why is It Dark at Night?
4. The Big Bang
5. Where is the Centre of the Universe?
6. Why the Cosmos is Actually Fairly Simple
7. What is the Speed of Space?
8. What is the Shape of Space?
9. Explaining the Cosmic Big Fix
10. Most of Our Universe is Missing
11. What is Dark Energy?
12. Where Does Matter Come From?
13. Gravity Conquers All
14. Warped Time and Black Holes
15. Is Time Travel Possible?
16. What is the Source of Time’s Puzzling Arrow?
17. The Black Hole Paradox
18. A Theory of Everything?
19. Fossils from the Cosmic Dawn
20. Can the Universe Come from Nothing?
21. How Many Universes Are There?
22. The Goldilocks Enigma
23. What’s Eating the Universe?
24. Is the Universe Actually a Botched Job?
25. Are We Alone?
26. Is ET in Our Backyard?
27. Why Am I Living Now?
28. The Fate of Our Universe
29. Is There a Meaning to It All?

30. What’s New on the Cosmic Horizon?

Index
About the Author

Paul Davies is a Regents’ Professor of Physics and Director of the


Beyond Center for Fundamental Concepts in Science at Arizona State
University. The bestselling author of some thirty books, his many
awards include the Templeton Prize and the Faraday Prize of the
Royal Society. He is a Member of the Order of Australia and has an
asteroid named after him.
Preface

What’s Eating the Universe? is a scientific detective story. It explains


how age-old cosmic puzzles have recently been solved, while
startling new discoveries are upending our understanding of physical
reality. To fully grasp the big picture, we must take a journey from
the very edge of time itself, through our own epoch, into the infinite
future. This is cosmology, the study of the origin, evolution and fate
of the entire universe. It weaves together the large and the small,
the vastness of space with the innermost recesses of subatomic
matter – a human endeavour of breathtaking audacity, tackling fields
of inquiry that for millennia lay exclusively in the provinces of
religion and philosophy.
There may be those who feel that, when the bright light of
scientific inquiry is shone into the hidden workings of nature, it strips
away the romance of the unknown. I have always believed, however,
that the deeper we probe, the more beautiful and awe-inspiring the
physical world appears. Nature demystified is nature revealed in all
its glorious subtlety and elegance. It is good that we should know.
In the chapters that follow, we will explore humankind’s
cosmological discoveries in detail, and examine the fundamental
philosophical questions that flow from them – questions such as why
the universe exists at all, why it has the form it does, why the laws
of nature are as they are, and how a system of mindless,
purposeless particles brought forth conscious, thinking beings who
can make some sort of sense of their world.
I have been preoccupied by these big questions of existence all
my life, and I am privileged to have been able to work on some of
them professionally in my career as a theoretical physicist,
cosmologist and astrobiologist. While my own contributions have
been modest, I have rubbed shoulders with some of the giants of
physics and astronomy, women and men of sparkling intellect and
insatiable curiosity, whose infectious zeal for understanding the
world from the perspective of science has been a constant source of
inspiration. I have lived through one of the most exciting times in
the entire scientific enterprise, when many big questions were
transformed from dreamy theorizing to hard-won discovery. As so
often in science, one question was answered only to raise a dozen
others. And so it is today. Much has been achieved but much
remains shrouded in mystery. I conceived of this book as a concise
summary of our current state of understanding.
So many people have helped me with my research over the years
that they are too numerous to list here, but I should like to
acknowledge Cecilia Lunardini and Rich Lebed of Arizona State
University, who clarified some aspects of particle physics for me,
Charley Lineweaver of the Australian National University for his
assistance with the characteristics of cosmological horizons, Glenn
Starkman for his inspirational lecture about the oddities in the
cosmic microwave background radiation, and Simon Mitton for
drawing my attention to some historical inaccuracies. Lucy Hawking
and Christopher McKay provided valuable help with the figures.
Special thanks must go to my wife, Pauline, for her encouragement
and critical reading of the text. Her communication skills far exceed
mine, and much of the credit for the end product must go to her
editorial refinements. Finally, I should like to thank my editor at
Penguin, Chloe Currens, for her suggestions, comments and
guidance, delivered in an unfailingly cheerful tone.

Paul Davies
Phoenix, March 2021
1. Journey from the Edge of Time

On 14 January 1990, the world’s press featured a mottled red-and-


blue image purporting to show nothing less than the birth of the
universe. ‘It was like looking at the face of God,’ proclaimed the
project’s lead scientist, George Smoot. In the words of Stephen
Hawking, the picture represented ‘one of the greatest scientific
discoveries of the century, if not all time’.
The subject of these superlatives was a colour-coded heat map of
the sky produced by a satellite called COBE, for Cosmic Background
Explorer. COBE had been tasked with surveying the fading afterglow
of the Big Bang, a sea of microwaves that suffuses space and travels
to us largely undisturbed from a time when the cosmos was a tiny
fraction of its current age. The amorphous-looking splodges
decorating the image indicated slightly hotter and cooler regions of
the universe. Etched into this kaleidoscopic pattern were important
clues about the birth pangs of the cosmos a split-second after its
origin, at the very edge of time itself.
Figure 1. All-sky map of the Big Bang afterglow obtained by the satellite COBE.

COBE ushered in a golden age of cosmology. In the three decades


since, the field has been transformed from a speculative backwater
to a precision science. Paradoxically, we now understand the history
of the universe in its overall outline better than we understand the
history of our own planet. Yet, to borrow from Churchill, this is not
the end of cosmology. It is not even the beginning of the end. But it
is, perhaps, the end of the beginning.
Cosmology might seem like a rarefied discipline, but in many
indirect ways it touches everyone. We all have a need to know why
the world is as it is and how we came to exist. Throughout history,
societies have sought to address this need by producing creation
myths: accounts which weren’t explanations in the scientific sense
but stories intended to place human beings in the context of a
grander scheme. When cosmology emerged as a scholarly discipline
among the ancient Greek philosophers two and a half millennia ago,
it was given a name that derived from the same root as ‘cosmetic’,
meaning beautiful, whole and complete; standing in opposition to
chaos. The word implied that there was such a thing as ‘a universe’,
a coherent and organized entity that can be understood by human
reason. Further progress, however, had to await the scientific age
two thousand years later, which unleashed a stream of dazzling
discoveries. When in 1543 Copernicus declared the Earth goes
around the sun, he shattered the anthropocentric model of the
cosmos that had prevailed for centuries. To be sure, the immediate
effect on daily life was minor; there were no riots, no wars, no
economic disruptions. Yet, over time, the knowledge that we are not
located at the centre of the universe fundamentally transformed the
context of all human existence. The impact was felt not only in
science, but in religion, sociology and economics too.
Today we are poised to undergo a shift in perspective even more
disruptive than that initiated by Copernicus. Future generations will
look back at our era and envy those privileged to witness it first-
hand. But beneath the catalogue of discovery lies a profound
mystery. For some reason, on an unexceptional planet orbiting a
run-of-the-mill star, a species of organism evolved that managed to
work out how the world is put together. That surely tells us
something deeply significant about our place in the natural order.
But what?
2. The Search for the Key to the Universe

It’s impossible to look up at the night sky and not be struck by the
grandeur and beauty of the vista – the sweeping arc of the Milky
Way, the myriad twinkling stars, the insistent, steadfast brightness of
the planets. The sheer vastness and complexity are overwhelming.
For millennia after millennia, our ancestors observed the same sky,
and struggled to make sense of what they saw. What was the key to
understanding it? How did the cosmos come to exist? What was the
place of human beings in the grand scheme of things?
For many ancient societies, making sense of the heavens was not
merely a philosophical or spiritual quest; it was also a practical
necessity. Knowing the movement of heavenly objects was critical
for human well-being: not just for navigation, but seasonal
migration, crop growing and timekeeping. Our distant ancestors’
preoccupation with the cycles of the sun, moon and planets is
obvious from the megalithic structures they built, some deliberately
designed to align with astronomical events – events often imbued
with divine significance and marked by elaborate ceremonies. The
sky was regarded as the realm of supernatural agents. In some
cultures, the sun, moon and planets themselves were treated as
gods.
But the regularities evident in the movement of astronomical
bodies suggested a very different concept of the heavens, not as the
playground of the gods but a mechanism, an elaborate system of
moving parts. Once this notion became established, precision
measurements were crucial to determine how the mechanism was
organized and regulated. Arithmetic and geometry were now
essential skills. Astronomers became powerful and important figures
in society, alongside priests and emperors. Their careful
measurements and analysis gradually revealed order and harmony,
number and form, in the celestial activity. Many theoretical models
were constructed over the centuries. A well-known distillation of
earlier ideas was compiled by the second-century CE Greek
astronomer Claudius Ptolemaeus. Ptolemaic cosmology featured a
complicated system of nested spheres, turning around the Earth at
different speeds.

Figure 2. The universe according to Ptolemy, with Earth at the centre.


The mechanistic models of the universe were on the right track,
but the theological dimension was never eliminated entirely. There
was ever the vexatious issue of origins. How did the great cosmic
contraption come to exist in the first place? Was there a Prime Mover
who set the complex mechanism in motion? A supernatural Creator
who conjured order out of chaos? A god who made the universe
from nothing? No attempt was made in these early models to link
the motion of astronomical objects with the motion of material
bodies on Earth. Heaven and Earth, each filled with movement,
remained separate domains. This view persisted through much of
the Medieval period, until the seventeenth century. Then, suddenly,
humankind’s understanding of the universe was transformed. A small
band of visionary ‘natural philosophers’ came to realize that the key
to the universe was not to be found in divine agency, nor in the
geometry of the cosmic architecture itself. Rather, it resides in laws
of nature that transcend the physical world and occupy an abstract
plane, invisible to the senses but nevertheless within the grasp of
human reason. Number and form, beloved of the ancient
philosophers, are manifested not just in specific physical objects and
systems, but interwoven into the very laws of nature themselves,
forming a mosaic of subtle patterns encrypted in a kind of cosmic
code. It was a stunning conceptual pivot, marking a transition from
mere description of the world to explanation. The quantum leap in
comprehension accompanying this transition was poetically
expressed by Galileo in 1632: ‘The great book of nature is written in
the language of mathematics,’ without which, ‘one wanders in vain
through a dark labyrinth.’ The key to the universe, said Galileo, lay
with mathematical decryption, a sentiment echoed three centuries
later by the astronomer Sir James Jeans, who declared ‘The universe
appears to have been designed by a pure mathematician!’ Galileo
himself began the task of unlocking nature’s hidden mathematical
order, but it was a generation later, especially with the work of Isaac
Newton and Gottfried Leibniz, that everything came together. Theirs
wasn’t a grand and highly organized cultural enterprise, as scientific
research is today. The founders of what we now call science were
more like members of an exclusive cult, almost a secret society,
unconventionally religious, fractious and egotistical, still steeped in
the mystical traditions of antiquity.
Galileo pioneered the use of the newly invented telescope to study
the heavens. This enabled more accurate measurements to be made
of the movement of planets and the shapes of their orbits. Like
Galileo, Newton set his sights on the solar system, seeking to
uncover mathematical laws of motion that applied equally on Earth
and in space, and which could be tested by observation and
measurement. To achieve this, he needed to find the correct
mathematical key, but it was nowhere to be found in the great
corpus of Greek arithmetic and geometry, nor in their Medieval
refinements. So, he fashioned it himself, calling it the theory of
fluxions – what is today termed differential calculus. Starting in his
early twenties, and no doubt assisted by the need to self-isolate at
home in Lincolnshire during the great plague of 1665–6, Newton
applied fluxions to the laws of motion, and discovered how gravity,
that enigmatic force that reaches across space, weakens in a precise
arithmetical way, in inverse proportion to the square of the distance
between objects.
Suddenly, humanity possessed a new window on the heavens. The
swooping parabolas of comets, the graceful ellipses of planets, the
scalloped gyrations of the moon – all the delicate tracery of celestial
orbits – fell into place, linked to each other by the immutable logic of
fixed mathematical relationships. I’ll never forget the thrill I felt as a
student when I first applied Newton’s laws to the motion of a planet
going around the sun and out popped the formula for an ellipse! It
was like magic. Imagine the sense of awe that Newton himself must
have experienced when his own hand-crafted equations produced
the very geometrical patterns so assiduously catalogued by
astronomers from years of observations.
Spectacular though such an advance was, Newton had a grander
vision. Having explained the solar system, he set about applying his
law of gravity to the entire cosmos. Since Galileo had turned his
telescope on the Milky Way it was obvious that the universe is
teeming with stars. But how were they arranged? Were they
clustered in a huge but finite cloud, or scattered endlessly through
infinite space? Newton envisaged the cosmos as a gigantic
clockwork, with gravity shaping its structure, literally holding it
together, a universal force of attraction tugging tenaciously on every
object in space. Where gravity is opposed by motion along a curved
path, as it is for the Earth going around the sun, stability is attained:
our planet is pulled by the sun but doesn’t fall into it. But what about
the universe as a whole? With nothing to prop it up, why, wondered
Newton, doesn’t the entire assemblage of stars fall together into one
great mass?
The solution he proposed was that the universe must be infinite.
With no boundary and no centre of gravity, the cosmos lacks any
privileged place to collapse towards. A given star would be pulled
equally in all directions, the forces balancing out: ‘by their contrary
attractions [they] destroy their mutual actions’ was how he quaintly
expressed it. Clever though this dodge may have been, it was
something of an intellectual sleight of hand. The delicate equilibrium
isn’t in fact stable, as Newton himself acknowledged: ‘For I reccon
this as hard as to make not one needle only but an infinite number
of them … stand accurately poised upon their points.’ Newton’s
infinite cosmos, it appeared, would teeter on the brink of collapse.
Yet as a religious man (albeit in his own eccentric way), Newton did
not baulk at invoking the hand of God to sustain His creation when
necessary.
And there the matter rested. It was to be another two centuries
before the correct solution was found, but through the lens of
history we can now see that the answer was hiding in plain sight –
right above our heads.
3. Why is It Dark at Night?

Unless you live in the Arctic, night follows day as reliably as, well,
night follows day. Most people never spare it a thought. But it turns
out that the familiar darkness of the night sky tells us something of
(literally) cosmic profundity.
The fact that the sun has set is only part of the night-time story.
There is also the matter of the stars. We don’t normally notice the
illumination from starlight because it’s so feeble. That’s simply
because the stars are extremely far away. Take Sirius, the brightest
star in the sky. It’s actually intrinsically about twenty-five times
brighter than the sun, but the way it appears to us, the sun
outshines it by a factor of 13 billion. Sirius is 80 trillion kilometres
away, whereas the sun is only 150 million kilometres away. Put the
sun next to Sirius and it would be dim by comparison.
The brightness of a glowing object diminishes with distance in a
precise manner: a star seen from twice as far away would appear
one quarter as bright, from three times as far it is one ninth as
bright, and so on. The farther out in space you look, the dimmer the
stars appear. But set against that is the fact that at a greater
distance there are more stars, and the growth in stellar population
compensates for the dimming effect of distance. It’s estimated there
are 400 billion stars in our galaxy alone, and there are billions of
galaxies within the scope of modern telescopes. So why don’t we
notice the combined light from all these glowing sources?
The same basic conundrum troubled a little-known Swiss
astronomer, Jean-Philippe Loys de Cheseaux, in the middle of the
eighteenth century. It occurred to de Cheseaux that if, as Newton
proposed, the universe is boundless, with stars scattered throughout
space to infinity, then the night sky should be ablaze with starlight –
and the Earth fried to a crisp. The same conclusion was drawn by
the German astronomer Heinrich Olbers in 1823 and became
dignified with the term ‘Olbers’ paradox’.
The paradox would go away if the stars were distributed out to
some particular finite distance, with nothing but a dark void beyond.
The total amount of starlight might then not add up to much. While
that’s true, it runs straight into the problem Newton fretted over: if
the number of stars is finite, then what’s to stop them all from falling
into the middle to form a huge messy agglomeration? So, it’s
between the proverbial rock and a hard place: either the universe
will collapse, or the sky should be a permanent incinerator.
Figure 3. Olbers’ paradox. If the universe is infinitely populated with
stars, every line of sight ought to intersect a star eventually, so there
would be no dark parts in the night sky.

However, there’s another escape route from Olbers’ paradox. Even


if the stars do indeed populate space without limit, could they
nevertheless be limited in time? Today it seems obvious that stars
can’t go on shining for all eternity. Whatever their source of power,
sooner or later they will run out of fuel and fade. This was not so
obvious in the nineteenth century, though; nobody knew what made
the stars shine. In fact, it wasn’t until the 1940s that astrophysicists
pieced the story together.
Granted that the stars can’t shine for ever, they are certainly
shining now, so if there really are an infinite number of them why
doesn’t their accumulated light turn night into day? To answer that
one, we have to go back to 1676, and a landmark observation made
by Ole Rømer, a Danish astronomer who studied the motion of
Jupiter’s moon Io, discovered by Galileo. Io goes around the giant
planet as regularly as hands go around a clock. Rømer spotted that
the Jovian clock seemed to be running slow when Jupiter was on the
far side of the sun from Earth. The mismatch could be explained by
the fact that light would take a few hours to travel from Jupiter to
Earth. Depending on where Jupiter and Earth were positioned in
their orbits, the journey time would vary. Rømer knew the sizes of
the planets’ orbits, so he used that information to work out the
speed of light based on the careful timing of Io’s movements. The
answer he gave was 214,000 kilometres per second. Considering the
crude methods then available, that’s impressively close to the actual
value of 299,792.458 kilometres per second.
The speed of light is so fast it might as well be infinite for
everyday purposes. But in astronomy it makes a big difference. The
light from Sirius, for example, takes 8.6 years to get here, so when
you see Sirius in the sky, you are seeing it as it was 8.6 years ago. If
it exploded today, we wouldn’t know about it for nearly a decade.
The farther away the star is, the deeper in the past we see it. The
Andromeda galaxy, for example, faintly visible to the naked eye as a
fuzzy patch of light, contains stars seen by us today as they were
two and a half million years ago. The distance light travels in a year,
known as a ‘light year’, is a convenient unit in astronomy: one light
year is about 10 trillion kilometres.
Back to Olbers’ paradox, then. The finite speed of light transforms
the entire argument. Suppose a particular star has been shining for
a billion years. If it’s more than a billion light years away we
wouldn’t be able to see it anyway, because its first light would not
have arrived on Earth yet. So even if the universe is infinite as
Newton suggested, we could still see only a limited number of stars
– those whose light has had time to reach us. The region beyond
that would appear dark.
Just in the last decade or so, the Hubble Space Telescope has
actually been able to penetrate almost as far as the dark zone,
beyond which there are no stars. It’s over 13 billion light years away
and encloses a volume of space around Earth that encompasses
about a trillion trillion stars. I just did a back-of-the-envelope
calculation to estimate how much the combined light of all those
stars adds up to, and it’s about the same in total as the light cast on
Earth by Jupiter but spread evenly across the sky. No wonder we
don’t notice it. At the end of 2020, astronomers announced that they
had actually managed to measure the integrated starlight of the
cosmos, using the New Horizons spacecraft which is zooming away
from us beyond the orbit of Pluto (where it is very dark). The night
sky turns out to be less than a ten-billionth as bright as the sun, but
apparently twice as bright as predicted, for reasons as yet unclear.
It’s remarkable that it took so long for astronomers to put two and
two together in this way. At any time since the seventeenth century
they could have drawn the obvious conclusion: the sky is dark at
night because the universe cannot always have been the way it is
now. Something very different – or perhaps nothing at all – must
have preceded it.
4. The Big Bang

In Flagstaff, Arizona, there is a famous observatory built in 1894 by


a rich businessman, Percival Lowell. His plan was to use the
telescopes to look for Martians. In the latter half of the nineteenth
century this wasn’t considered totally off the wall. Scientists openly
discussed the possibility that Mars was inhabited, and astronomers
searched eagerly for signs of life on the red planet. In 1877, an
Italian astronomer, Giovanni Schiaparelli, said he could see straight
lines, or ‘channels’, on the planet’s surface, and this stoked much
speculation about ‘canals of Mars’. Mars fever was brilliantly captured
in H. G. Wells’ 1898 science-fiction story The War of the Worlds.
Fixated by the notion of Martian engineers, Lowell embarked on his
own observations, producing elaborate maps of what turned out to
be an entirely fanciful network of canals.
While Lowell pursued his quixotic quest, rather more conventional
astronomy was also being conducted at his observatory. By the late
nineteenth century, telescopes had advanced to the point where
they could probe far beyond the confines of our Milky Way galaxy. A
big issue of the day concerned the large-scale organization of the
cosmos. In particular, what were all those nebulae – wispy patches
of light – seen scattered across the sky? Were they giant gas clouds
located within our galaxy, or were they entire galaxies in their own
right, too far away for the individual stars to be discerned?
In 1909, a Lowell Observatory astronomer named Vesto Slipher
set about examining the quality of the light from the mysterious
nebulae. The only telescope he had was a relatively modest 24-inch
instrument, so this was slow, painstaking, repetitive work. With no
fancy electronic gizmos of the sort astronomers have today, all
observations had to be done by hand and eye, often with on-the-fly
improvisations to coax the best out of the equipment. Slipher
analysed the faint nebulous glows with a device called a
spectroscope, designed to split light into its constituent colours. He
laboured away night after night, often in freezing conditions,
recording the results on film. In those days, an astronomer’s lot was
not a happy one. Yet the compulsion that drives the Sliphers of this
world is that dedicated toil in some small corner of a subject can
unexpectedly strike gold. And that’s just what happened at the
Lowell Observatory. By 1912 Slipher had assembled enough data to
conclude that most nebulae were measurably redder in colour than
the Milky Way. Why was that? An explanation was immediately
apparent. When a light source is receding at great speed, the
emitted light waves are stretched, shifting the wavelength towards
the red end of the colour spectrum. Slipher therefore concluded that
most nebulae are rushing away from us.
With hindsight, we can see that 1912 marked the true birth of
modern cosmology. But there was no fanfare, no press conference,
just a careful technical paper buried in the Observatory’s bulletin. It
took several years and many more observations before Slipher’s
discovery gained prominence, eventually coming to the attention of
a certain Edwin Hubble, a lawyer turned astronomer, rarely seen
without a pipe. In 1924, Hubble used the big 100-inch telescope at
Mount Wilson in California, then the world’s largest, to measure the
distance to the Andromeda nebula by managing to detect individual
stars, and in so doing proved that Andromeda is in fact another
entire galaxy like the Milky Way. Hubble went on to estimate the
distances to another twenty-three galaxies. Then he combined his
results with Slipher’s red-shift measurements and glimpsed the
outline of something systematic: the farther away the galaxy was
located, the redder its light and the faster its recession. It seemed to
be in proportion. The simplest interpretation of this pattern was that
the universe was gradually expanding, growing larger over a
timescale of billions of years (see ‘Hubble wars’, below). Hubble duly
announced his results to the world in the New York Times on 23
November 1924. It was undoubtedly one the most momentous
discoveries of the twentieth century, and plaudits were soon
showered upon the pipe-smoking astronomer, leaving Vesto Slipher
as the unsung hero of the expanding universe.
With the realization that the universe isn’t just there – an
unchanging collection of glowing oddments – but is a dynamic
system, evolving with time, a host of questions arose about its
trajectory. Where had this system come from and where was it
heading? What determines how fast the universe is expanding?
Could the expansion rate change with time? When did it begin, and
will it go on for ever?
Two implications of the expansion were immediately obvious. First,
if the universe is getting bigger, it must previously have been smaller
and denser. Second, the effect of gravity – a universal attractive
force – would act like a brake on the expansion, slowing the
recession of the galaxies as they pulled on each other. Because of
this deceleration, the expansion must have been faster in the past.
Hubble’s observations weren’t extensive or accurate enough to
detect any change in the rate of expansion over the few million
years that they encompassed. However, the braking effect of gravity
was easy enough to study theoretically, and as early as 1921, the
Russian astronomer Alexander Friedman had calculated precisely
how the expansion rate would gradually slow, but his deliberations
were mostly ignored. Astronomers were flirting with one of the most
far-reaching discoveries in the history of science, but such was the
tentative nature of these early results that no leading scientist was
prepared to take the plunge and state the obvious implication. In the
event, it fell to a young Belgian priest and theoretical physicist, Abbé
Georges Lemaître, to come out and claim explicitly in 1927 that the
universe we now observe must have begun billions of years ago as a
‘cosmic egg’, a state of enormous density that expanded with
explosive force. This was the precursor of the modern ‘big bang’
theory.

Hubble wars
The rate of expansion of the universe is expressed as a
number, known as the Hubble constant, or H. The man
himself assigned H a value of 500. (In the peculiar system
of units that astronomers prefer, that means a galaxy
about 3.3 million light years away is receding on average
at 500 kilometres per second.) Given H, and factoring in
the braking effect of gravity, you can work out the age of
the universe. Hubble’s original value put the universe at
only about 2 billion years old – less than half the age of
Earth! Astronomers lifted their game and produced
estimates of H that progressively pushed back the inferred
age, but they were divided into two warring factions. One
touted a value of H of 180, the other 55, using the same
methods and each insisting that the errors in
measurement were far too small to close the gap. The
discrepancy was an important issue because the smaller
the number, the greater the age of the universe. In the
1980s, data from the Hubble Space Telescope put paid to
the angst. H finally came out at 73 – a nice compromise –
making the universe 13.8 billion years old by current
estimates. But recently a new discrepancy has surfaced.
Measurements of H using cosmic microwave background
(CMB) data yield a value of only 67, implying an age for
the universe of well over 14 billion years. Does this
suggest something seriously amiss with our understanding
of basic cosmology? Time will tell.

You might have expected that a declaration of such profound


importance would be a scientific, not to mention theological,
sensation. Yet again, however, the response was muted. Hubble
himself doubted Lemaître’s conclusion. Albert Einstein, by then the
world’s greatest authority on gravitation and cosmology, was equally
dismissive. ‘Your calculations are correct,’ he wrote to Lemaître, ‘but
your physics is atrocious.’ Einstein had also shrugged Friedman’s
earlier theoretical effort aside, and indeed he only accepted that the
universe was actually expanding after visiting Hubble in California in
1931. After that, he did a U-turn and backed Lemaître’s work. In
spite of this illustrious endorsement, speculation about the origin of
the universe wasn’t taken very seriously in the 1930s; indeed,
cosmology was hardly even a recognized subject.
Happily, the theoretical work of Friedman and Lemaître wasn’t
forgotten, though it took another two decades before it was
revivified by George Gamow, a defector from the Soviet Union
working in the United States. Gamow wasn’t an astronomer; he was
a nuclear physicist. It was he who explained the type of radioactivity
known as alpha decay. Gamow reasoned that the young, highly
compressed universe must have been hot enough to permit nuclear
reactions. Consequently, it would have glowed like a furnace. Which
raised a fascinating possibility. Might a fading remnant of that
primordial heat still pervade the universe today, forming a cosmic
background of microwave radiation?
Figure 4. The antenna that first detected the hiss of microwaves emanating from
the birth of the universe.

Did the big bang really happen?


Not all astronomers accepted the link between the cosmic
expansion and an explosive origin. Ironically, the familiar
moniker ‘big bang’ was initially coined by the British
astronomer Fred Hoyle in 1949, while dismissing the idea.
Hoyle thought Lemaître’s model of a universe exploding
into being was nonsense, and developed a completely
different interpretation of Hubble’s observations, called the
‘steady state’ theory. The basic idea is that as the universe
expands and the galaxies move apart, so new matter is
continually created, gradually aggregating into fresh
galaxies to fill the ever-growing gaps. As a result, on a
very large scale the universe would look much the same
for ever – a mix of old and new galaxies sustained by a
process of continual replenishment. There would be no
beginning and no end, and no hot, dense primordial state.
Hoyle fought fiercely for his theory, marshalling a band
of loyal supporters. For about twenty years the two
theories rivalled each other for support. But then the
knockout blow came. The discovery of the CMB had no
credible explanation within the steady state model, and
support for it rapidly dwindled.
I went to work with Hoyle in Cambridge at this critical
juncture. Here was this world-famous astronomer and
public celebrity, known for his science-fiction novels as
well as his research, strangely isolated, casting around for
some way to rescue the essence of his theory, perhaps by
treating the big bang as an interlude rather than an
absolute origin. He gave up on the idea that particles of
matter were continually created in a thin soup throughout
space in favour of concentrated ‘creation centres’. In the
1960s, highly compact objects called quasars were
discovered that fling out huge quantities of energetic
material. Hoyle believed these energetic sources were
cosmic spigots, pouring brand new matter into the
universe. But the simplicity of the original steady state
concept was lost, and in 1972 Hoyle resigned his Chair at
Cambridge in disillusionment and became something of a
recluse, tucked away in a remote cottage in Cumbria,
where he filled his days fell-walking and sniping at the
scientific establishment.

Gamow was on the right track. In 1964, two scientists working on


satellite communications at the Bell Laboratory in New Jersey
accidentally came across this remnant heat, bathing the universe at
a temperature of about 2.7 degrees above absolute zero (absolute
zero is about −273oC). It showed up as an annoying hiss in their
receiver (see Figure 4). All attempts to explain it away as a defect in
the equipment (including pigeon droppings in the antenna) failed,
and the only explanation left was that the disturbance came from
outer space. This was the big bang’s smoking gun.
Suddenly, cosmology was propelled into the scientific mainstream,
and began to attract some of the brightest minds in physics and
mathematics – the likes of Roger Penrose and Stephen Hawking.
The origin of the universe in a big bang was at last taken seriously
and became the focus of intense theoretical analysis. Astronomers
began to clamour for better observations of the cosmic microwave
background – today referred to simply as the CMB – confident that it
contained crucial clues about the early universe. Because of
atmospheric absorption, a decent view of the CMB needs a satellite,
and that’s what they got. In November 1989 NASA launched COBE,
and, with it, cosmology’s golden age.
5. Where is the Centre of the Universe?

It’s all very well to say the universe is expanding, but what does that
mean? Expanding into what? Coming from where? When I first
learned that the galaxies are rushing away from us it seemed to me
that Earth must be close to cosmic ground zero – the point in space
where the big bang happened. On the other hand, I knew it was
crazy to believe we were at the centre of the universe.
To properly understand the nature of the cosmic expansion, you
have to think about it in a totally different way. Forget all those TV
animations depicting the big bang as the explosion of a glowing
lump, with fragments flung out into the surrounding space. As far as
we can tell, the universe has no centre. The impression that we are
close to it is an illusion. In reality, every galaxy is moving away from
all the others.fn1 The view from any given galaxy would be roughly
the same; there is nothing special about our own location.
A loose analogy is with a dance class where the students form a
ring holding hands while the teacher explains the steps. The teacher
tells all the students to step back five paces. As a result, the ring
expands, and every student is now farther from every other student.
But there is no special, privileged student that the others are fleeing
from. Of course, the galaxies aren’t arranged in a ring, but the same
basic idea is at work.
Actually, it is better to stop thinking about galaxies moving
through space altogether. Instead, imagine the space between the
galaxies to be swelling. This image is not just an aid to visualization:
space really is growing bigger: every day, a hundred billion billion
cubic light years of additional space appears within the observable
universe. It takes a while to get used to the idea that empty space –
a void – can stretch or swell, but it’s true: space is provably elastic.
In fact, not only can space stretch, it can buckle, twist and shudder
too. The cosmic expansion is literally ‘putting space’ between the
galaxies, and as a result they get farther and farther apart. This new
space doesn’t have to ‘come from anywhere’ or expand into
anything. It is space.
Armed with this new insight, we can confront the question ‘Where
did the big bang happen?’ The answer is everywhere. The CMB – the
afterglow of the big bang – doesn’t have an epicentre; it isn’t a
stream of radiation coming from a specific point in space. The entire
universe is immersed in microwaves, as if it were entombed in a
gigantic oven. Observing the CMB, you are witnessing the birth of
the universe at every point in the sky.
The stretching of space also gives a useful alternative way to think
about the red shift. Light waves from a distant galaxy have to travel
through expanding space to reach us. As the intervening space
stretches, so the transiting light waves get stretched with it,
meaning their wavelength increases and their frequency falls: blue
becomes red. For the oldest observable galaxies, the wavelengths
are eleven times longer when they arrive on Earth than when they
set out from the source.
The farther out into the universe we peer, the bigger the red shift
becomes. The stretching factor – the ratio of the wavelength
observed on Earth to what it was when emitted by the source –
escalates with distance, until it shifts the received light well beyond
the visible region of the spectrum into the infrared, microwave and
radio regions, longer and longer, without limit. Theoretically, the
stretching factor would become infinite at a certain critical distance.
In a simple model, that distance corresponds to how far light can
have travelled since the big bang. Obviously, we couldn’t see any
farther than that, because light hasn’t had time to reach us yet
during the age of the universe. There is, therefore, a horizon in
space – a light horizon – restricting our view. And because nothing
can go faster than light there’s no way to know for sure what lies
beyond. No instrument, however powerful, can enlighten us. But just
as a terrestrial horizon doesn’t mark the edge of the world – what
lies beyond is much the same – so the cosmic horizon isn’t ‘the edge
of the universe’. There is no edge. The horizon is simply the
boundary of our visible cosmic patch. The universe may very well be
infinitely extended in space, but if it is finite in time we can’t look to
see. I should mention that, in practice, the horizon is shrouded by
the fiery afterbirth of the big bang. The red shift stretching factor of
the CMB radiation coming from that epoch is about 1,000.
One final point. If a galaxy is, say, 12 billion light years away, we
are seeing where it was located 12 billion years ago – so one can
ask: where is it now? Because of the expansion of the universe it will
be many billions of light years farther away today than it was then.
For example, the horizon is estimated to now be about 47 billion
light years away. It’s important to remember that our telescopes
don’t provide a snapshot of the universe as it is today. Instead, the
amalgamated images form a compilation, a historical time sequence
cascaded together. It’s rather like cutting a movie into frames,
stacking the frames into an ordered pile, and looking down through
them to see the whole stack at once.
6. Why the Cosmos is Actually Fairly
Simple

Browsing in the public library as a teenager, I picked up a little book


purporting to contain nothing less than the equations of the
universe. I confess I was sceptical. Could the great story of the
cosmos be captured in just a few lines of mathematics? Was it really
that simple? The book, by Dennis Sciama, was called The Unity of
the Universe, and it left a deep impression on me. The title says it
all: there is a coherence and unity on the largest scale of size that
lets one distil the basic story of the universe into three or four
equations.
Why do I say that the universe is simple? Mustn’t it be the most
complicated thing we can think about? Well, yes and no. If you tried
to describe everything that exists in individual detail, then of course
it would be incomprehensibly complex. But just as one can recount
the history of, say, the Vietnam War without mentioning the exploits
of every given soldier, so we can determine the basic thrust of
cosmic history without getting too bogged down in local details.
Nevertheless, even this broad-brush narrative wouldn’t be possible
without some very specific features.
The most important of these is the uniformity of nature. As far as
we can tell, identical laws of physics apply on the far side of the
universe as they do in our cosmic neighbourhood. Why this should
be so nobody knows, and it may not even be a proper scientific
question anyway, but it is a critical reason why we can even discuss
‘the universe’ as a single entity and construct a narrative for the
system as a whole. If the laws differed from place to place or time to
time, there would be no common scheme of things.
But that is not all. Even with universal laws, there are countless
ways the universe could be arranged. Imagine an orchestra
scheduled to perform a concert featuring, say, Beethoven’s Fifth
Symphony. It would make perfect sense to attribute this musical
performance to ‘the orchestra’, even though orchestras are made up
of many individual musicians. So long as their playing is
‘orchestrated’ the result is harmonious. If each musician played from
a different score without regard to the others, the upshot would be a
horrible cacophony. It would then be pointless to say the orchestra
was performing anything specific; indeed, it would be pointless to
even speak of ‘the orchestra’ as opposed to an assemblage of
independent musicians.
The universe is organized in an analogous way. Rather than a lot
of complicated parts behaving independently, there is orchestration
and harmony and co-ordination. It shows up most obviously in the
way the universe is observed to expand at the same rate
everywhere. Furthermore, the number of galaxies in a similarly sized
patch of sky is the same in every direction, as is the temperature of
the CMB. Evidently the big bang went ‘bang’ in a remarkably ordered
fashion. Contrast that with explosions on Earth. If a leaking gas main
blows up a house, fragments are scattered haphazardly far and
wide. It is a mess.
There is a third crucial simplifying feature. According to folklore, in
48 BCE the great library of Alexandria was destroyed by fire, one of
the great disasters to befall human civilization. Priceless knowledge
was lost as books were transformed into piles of ash in a tragic
demonstration of the destructive power of heat. As a rule, the hotter
the fire, the more it consumes. Modern incinerators can break down
the most stable chemicals. The interior of the sun is so hot that even
atoms can’t exist; they are smashed into nuclei and electrons. The
big bang was hotter still. At one second after the beginning, it was
ten billion degrees; at one microsecond, ten trillion degrees; at a
picosecond, ten thousand trillion degrees.
Hot = simple is one reason cosmologists can talk so confidently
about the early universe. It really is true that a handful of simple
equations – basic algebra and calculus – accurately captures most of
what happened in the primordial phase. A new-born universe in
which everything is mushed together and broken down to its basic
constituents is far easier to describe theoretically than, say, the
Earth.
7. What is the Speed of Space?

Many people are fearful of flying, so take-off is a stressful time for


them. I suffer from the opposite problem: I’m so stressed about
missing the flight that by the time it leaves I’ve usually fallen asleep
from nervous exhaustion. I often wake up half an hour later with no
idea whether the plane is still stuck on the runway or halfway to Los
Angeles.
The thing is, you can’t tell from within the plane whether it’s
moving or not. Of course, if it banks or dives you’d know it right
away, but steady motion can’t be felt. It was Galileo who first made
this explicit, pointing out that uniform speed can be measured only
relative to something else. My car speedometer, for example,
measures the speed of the car relative to the road. The plane’s
speed is measured relative to the ground, or the air.
Though physicists long ago accepted that uniform motion is only
relative, some of them still wondered how fast Earth moves through
space itself as our planet swings around the sun and glides across
the galaxy. We don’t feel this motion, but we are travelling through
space, which means that from our frame of reference, space is
passing through us, all the time, unnoticed. Well, how fast is it
going? How many litres of space per second slide through my body
with ethereal stealth? After Galileo, it was recognized that no
material device could measure the uniform speed of space. We can’t
dip a metaphorical finger into the void around us to detect the
slipstream as space sweeps by. But what about non-material
methods?
By the late nineteenth century, a new possibility had presented
itself. Perhaps light could be used to measure how fast Earth is
travelling through space? At that time, physicists imagined space as
filled with a sort of ghostly jelly-like substance referred to as ‘the
ether’, and they envisaged light waves as vibrations in the ether
travelling at a definite, fixed speed – the speed of light. Physicists
invented various optical devices to try and measure the speed of the
Earth through the ether. After a few years of effort, the result they
obtained was, precisely, zero. Apparently, the Earth wasn’t moving
through space (i.e. the ether) at all! This made no sense. There was
a flat-out contradiction between Galileo’s relativity of uniform motion
and the theory of light travelling at a fixed speed. The clash called
for a fundamental re-evaluation of the nature of space, time and
motion, and the properties of light. The challenge was taken up by
Einstein, culminating in the first phase of his famed theory of
relativity, published in 1905. Einstein abolished the ether, declared
that the speed of light was an absolute fixed number however the
observer who measures it is moving, and reaffirmed Galileo’s
position that uniform motion is always relative to some other
material thing. It is not only impossible, he said, but meaningless to
talk about the speed of something relative to space itself.
All that’s well and good, but what about non-uniform motion, i.e.
acceleration? We have no trouble detecting that. If my in-flight
coffee spills because the plane hits turbulence I don’t need to look
out of the window to tell that our uniform motion through the air has
been disturbed. Acceleration is a change in velocity, and we can
literally feel it happening. If you slam on the brakes in a car, you get
pitched forward. If you yank the steering wheel and take a tight
corner, you lurch against the door. Both are examples of
acceleration, the first being a change of speed, the second a change
of direction.
A steady rate of rotation represents a constant acceleration, and
Newton used a simple example to illustrate its properties. He pointed
out that water in a spinning bucket creeps up the sides; we normally
attribute this to ‘centrifugal force’. The question, ‘Is the bucket
spinning or not?’ is immediately answered by seeing whether the
surface of the water is flat. It’s not necessary to look at, say, the
ground, to tell, which suggested to Newton that rotation, and
accelerated motion generally, need not be gauged relative to other
material systems, but is absolute. Sometimes this was expressed by
saying that acceleration is relative to absolute space itself.
Not everybody agreed. Right up to the twentieth century some
scientists argued that even rotation must be understood as relative.
Relative to what? Physicists used to say relative to the ‘fixed stars’.
Of course, the stars aren’t really fixed, they are just so far away that
we don’t notice their movement: relative to the distant matter in the
universe is a better description. To visualize this proposal, imagine
riding with your eyes closed on one of those horrid fairground
contraptions where you pay good money to be hurled around in
circles in great discomfort at high speed. Open your eyes, look up.
What do you see? You see the stars whizzing round. When the stars
stop, the ride has stopped, and you can get off.
This basic observation led some people, most notably the Austrian
engineer and philosopher Ernst Mach (the same Mach whose
eponymous numbers are used to describe aircraft speed), to
attribute the feeling of rotation – and acceleration more generally –
to the far-flung stars. When you are pressed against the side of a
whirling fairground seat, or feel your ‘stomach left behind’ when a
lift suddenly descends, it’s because the stars are pulling on you, said
Mach. It was a beguiling theory and became known as Mach’s
principle. It had many adherents, including Einstein, who warmed to
the idea that all motion was relative, not just uniform motion. He
hoped Mach’s principle could be incorporated into a more general
version of his theory of relativity, which he worked on in the early
part of the twentieth century (see ‘The most beautiful theory of all
time’, opposite). His thinking was that the gravity of all the distant
stars and galaxies would combine to produce locally detectable
effects like centrifugal force.
The most beautiful theory of all time
General relativity has been hailed as humanity’s finest
intellectual accomplishment, at once a scientific theory
and an art form of breathtaking elegance. It entwines
space, time, matter and force in an ingenious system of
equations. Unlike many theories of physics, which are built
in stages from the bottom up, general relativity springs
whole from a few grand overarching principles, such as
requiring that all physical properties must be independent
of the co-ordinates used to describe them. Geometrical
forms are sculpted from space itself by the arrangement
of matter and energy. There is no actual force of gravity
as Newton proposed, only restless, evolving geometry
through which matter churns and light glides. The
blandness of Newton’s immutable void is replaced by a
concept of space as a vibrant, dynamic entity, stretching
and shrinking, twisting and curving, pulsating and
convulsing, and vibrating in energy-transporting
undulations that ripple out across the universe at the
speed of light (see ‘Gravitational waves’, p. 41). And
general relativity works brilliantly. More than a century
after Einstein presented his masterpiece, not a single
observation has contradicted it.

Sadly, it didn’t work. Einstein’s general theory of relativity predicts


that a planet spinning in otherwise empty space will bulge round the
equator (as the Earth does) on account of its rotation, even in the
absence of any ‘fixed stars’ with which to gauge its motion.
Einstein’s theory is our best current understanding of gravity, so
when it comes to rotation, it rather vindicates Newton’s view that it
is absolute and not relative. And although we cannot meaningfully
speak of the speed of space, the acceleration of an object in empty
space still makes sense.
That isn’t quite the end of the matter, though. A few physicists
and cosmologists have continued to tinker with subtler formulations
of Mach’s principle, and the subject is not completely settled.
Indeed, many of my colleagues toggle back and forth between
thinking of space as some sort of stuff (as did Newton) one day and
as ‘nothing there’ the next. The lure of Mach’s principle is that it links
everyday human experiences (like the feeling of spinning) with the
organization of the universe, a sentiment captured most beautifully
in the poet Francis Thompson’s lines:

All things by immortal power,


Near or far,
Hiddenly
To each other linkèd are,
That thou canst not stir a flower
Without troubling of a star.
8. What is the Shape of Space?

Hotel rooms sometimes come with those tiny fish-eye lenses in the
door so if somebody knocks you can peer through the little hole and
see a face with a huge nose and tiny ears. Similar effects are
produced with fairground mirrors that are rather unflattering for the
figure. Distorting shapes with lenses and mirrors is trickery, but it
turns out that gravity does it for real. That is, gravity really does
warp space.
The first spacewarp to be observed was in 1919 when the British
astronomer Sir Arthur Eddington led an expedition to West Africa to
study an eclipse. To cast their horoscopes, astrologers follow the
movement of the sun as it migrates through the constellations of the
zodiac, moving against the background of stars. From time to time
the sun’s movement brings it into close alignment with a star. When
that happens it’s of interest to astronomers too. Eddington wanted
to know how the sun’s gravity might affect the light from such a star.
The trouble is, you can’t see stars in the daytime except during a
solar eclipse. Then the moon blots out the sun’s glare for a few
minutes, the sky darkens and the stars become visible. Astronomers
know the positions of the stars very accurately when the sun isn’t in
the way, and Eddington’s plan was to compare those recorded
positions with their positions during the eclipse, to see if they were
shifted. And they were. The stars were indeed very slightly askew
(see Figure 5). It’s as if there is a gigantic fish-eye lens in the middle
of the solar system.
Four years previously, Albert Einstein had used his general theory
of relativity to predict that the sun’s gravity would warp space and
shift the apparent positions of the stars. The very idea that gravity
can ‘bend starlight’ caught the public’s imagination, and Eddington’s
confirmation of the prediction elevated Einstein to international
fame. Curiously, the man himself took it all in his stride. He was so
convinced that his brainchild, general relativity, was right, that when
he was told about Eddington’s measurement he seemed
unimpressed. Suppose Eddington had obtained a contradictory
outcome, Einstein was asked by an assistant. ‘Then I would feel
sorry for the good Lord,’ he replied. ‘The theory is correct anyway.’
Since Eddington’s pioneering expedition, gravitational lensing has
matured into an entire branch of astronomy. Dramatic spacewarps
occur when one galaxy eclipses another, occasionally producing a
complete ring of light (see Figure 6). Black holes produce enormous
spacewarps, leading to spectacular effects on surrounding light
sources.
Figure 5. The bending of light. During an eclipse, stars can be seen near the sun’s
location in the sky. Einstein predicted that the sun’s spacewarp should slightly
displace the position of the stars, and Eddington confirmed this in 1919. Rather
than thinking of the star beam as bent when it passes close to the sun, it’s better
to regard light as taking the shortest, straightest path it can through curved space.

Early on, Einstein realized that the warping or curving of space


would not only occur around massive objects; it would also affect
the shape of the universe as a whole. In 1917, he presented his idea
of what he thought the overall cosmic geometry might be. This was
pre-Hubble and Einstein assumed the universe was static. He ran
into the same conundrum as Newton had, two hundred and fifty
years earlier: why didn’t the cosmos collapse under its own
unsupported weight?
Figure 6. Gravitational lens in which the spacewarp of a foreground
galaxy images a more distant object to produce a ring of light.

The difference was that Einstein came at the problem armed with
a totally new theory of gravitation. Perhaps general relativity could
provide a means to shore up the universe? He soon found that it
could, but only by adding an extra term to his prized gravitational
field equations. This he did reluctantly, feeling that the additional
term was a fudge, compromising the elegance and simplicity of the
original equations. But it seemed to do the job. The fudge factor
described a new kind of repulsive force, or antigravity, with the
strange property that it would become stronger with distance,
whereas normal gravity grows weaker. Einstein picked the strength
of the new force to match the combined gravity of all the matter in
the universe, to produce a static equilibrium.
A distinctive feature of Einstein’s static universe was its shape:
space curves in on itself in such a way that it closes up into a finite
volume, but without having any edge or boundary. To understand
what this means, imagine that a light beam deflected slightly by the
sun goes on to pass close to another star, getting deflected a bit
more, then another, and another … Is it possible for the light to be
bent right around in a closed loop? Einstein’s model universe showed
that it was. The entire universe could turn light beams into circles in
every direction at once. You could look out into space and, with a
big enough telescope, you’d see the back of your own head!
A closed space is difficult to picture in three dimensions, but easy
in two. The spherical surface of the Earth is finite but has no edge or
boundary. Einstein’s universe is an analogous concept, but with one
dimension extra – forming a three-dimensional shape called a
hypersphere. Although it’s hard to visualize, the amazing artwork of
M. C. Escher attempts to portray the geometry of curved space in
aesthetically striking woodcuts. Whether you can imagine it or not, a
hypersphere makes perfect sense and can be described
mathematically. It’s often asked what’s ‘outside’ the hypersphere.
The answer is nothing: there is no outside. All observers are located
in the space itself, so any discussion of a bigger space ‘containing’
the hypersphere is pointless.

Gravitational waves
Imagine that by some magic the sun were to disappear at
noon. We would not know about it until just after 12:08
when the sunshine abruptly stopped. According to
Newton’s theory, in which gravity acts instantaneously
across space, Earth’s orbit would cease curving at
precisely noon. Our planet would sense the sun had gone
even if we couldn’t. But in Einstein’s theory nothing can go
faster than light, including changes in gravity. If the sun
abruptly dematerialized, the resulting gravitational
disturbance would ripple outwards at a finite speed. In
1918 Einstein predicted that ripples of space can indeed
exist, and they travel at the speed of light. These
Another random document with
no related content on Scribd:
had stayed three months with the Scottish King; and no wonder that
when James ascended the throne of England Sir Henry Wotton was
one of the men then in London whom the King desired to see. He
was a favourite at Court; and his lifelong homage to the Princess
Elizabeth, the unhappy Queen of Bohemia, is well known.[40]
He had risen to great things, and might have risen to greater still if
it had not been for one brilliant Latin epigram written in an album.
Even King James, with pleasant memories of a packet of antidotes
and a most delightful guest in Stirling Castle, found it hard to forgive
the Latin epigram—“a merriment,” poor Wotton had called it—written
in an album in an indiscreet moment many years before, and
officiously forwarded from Augsburg to the Court of London: “An
Ambassador is an honest man sent to lie abroad for the good of his
country.” It is said to have ruined Sir Henry Wotton’s diplomatic
chances; and when, after some other missions, he came home in
1624, it was as a penniless man still, with plans of literary work and a
sufficient stock of memories grave and gay. He had consorted with
princes and statesmen, with artists, men of science, and men of
letters. He had worked for Essex and known Raleigh, and Francis
Bacon was his cousin. Among his friends abroad he had counted
Beza, Casaubon, Arminius and Kepler. He had watched Kepler at
work in his laboratory, and he had supplied Bacon with facts. And
when Bacon sent him three copies of his Novum Organum when it
first appeared, Sir Henry sent one of the copies to Kepler.
When Thomas Murray died, and Sir Henry Wotton was selected,
out of many candidates, for the Provostship of Eton, he was so poor
that he was obliged to borrow money to enable him to settle down
there. King James would have granted him a dispensation, but he
preferred to conform to the rule that the Provost of Eton must be a
man in Holy Orders. He had been duly ordained deacon, and, being
a man of liberal views, had steered “a middle way between
Calvinism and Arminianism.”[41]
When the two young sons of the Earl of Cork arrived at Eton, Sir
Henry Wotton had been Provost for ten years, and Eton could
scarcely imagine itself without him. With a royal pension in addition
to his Provostship, and assisted by a strong staff of Fellows of the
College—the learned Hales, John Harrison and the rest—he was
taking life easily, in the evening of his days, among his books and
curios, his Italian pictures, and those manuscripts—biographies of
Donne and Luther, and the History of England,—which he always
meant to finish and never did. He was not quite so active as when he
had first come among them with his new views of teaching, and had
put up the picture of Venice, where he had lived so long as
Ambassador, and had hung on the wooden pillars of the lower
schools his “choicely drawn” portraits of Greek and Latin orators and
poets and historians, for the little Eton boys to gaze at with round
English eyes; but his familiar figure was still a daily presence,
coming and going amongst them in his furred and embroidered
gown, “dropping some choyce Greek or Latin apophthegm” for the
benefit of the youngsters in class. He was still a “constant cherisher”
of schoolboyhood, taking the “hopeful youths” into his own especial
care, having them at his own hospitable table, picking out the
plodding boys and the boys of genius, and himself teaching best in
his own memorable talk. He liked to indulge in reminiscences of Italy
—“that delicat Piece of the Worlde”; and he sometimes looked
wistfully Londonwards, though in his gentle, deprecatory way he
spoke of it, especially in November, as a “fumie citie.” In his last
years he nursed hopes that he might succeed to the mastership of
the Savoy; meantime, from his Provost’s Lodging, he could look
across the “meandering Thames and sweete meadows,”[42] to the
great pile of Windsor Castle in its “antient magnificence”; and he
read and ruminated and smoked—he smoked a little too much,
according to his friend Izaak Walton—and counted his “idle hours not
idly spent” when he could sit quietly fishing with Izaak Walton in the
river-bend above the shooting fields, then, as now, known as Black
Pots. When Robert Boyle went to Eton in 1635, to be an Eton boy
meant not only being “grounded in learning” by such men as Hales
and Harrison, but being “schooled and bred” under the daily
influence of this soft, rich, delightful personality.
The two boys were known in the school as Boyle A and Boyle I:
Robert was Boyle I. According to Carew, they must have grown with
astonishing rapidity during their first months at Eton. Mr. Francis was
not only tall, but “very proportionable in his limbs,” and grew daily
liker to his brother, Lord Dungarvan. He was not so fond of his books
as “my most honoured and affectionate Mr. Robert, who was as
good at his lessons as boys double his age.” An usher, “a careful
man”, was helping them with their lessons, and Carew was keeping
an eye on the usher. Versions and dictamens in French and Latin
filled their time, and Carew could not persuade them to “affect the
Irish,” though Robert seems to have shown a faint, intermittent
interest in that language.[43] As for Mr. Robert, he was “very fatt, and
very jovial, and pleasantly merry, and of ye rarest memory that I ever
knew. He prefers Learninge afore all other virtues and pleasures.
The Provost does admire him for his excellent genius.” They had
acted a play in the College, and Robert had been among those
chosen to take part in it. “He came uppon ye stage,” wrote Carew,
exultant, to the Earl of Cork; “he had but a mute part, but for the
gesture of his body and the order of his pace, he did bravely.”
The little fellow was not yet nine years old, and his stutter must
have made it highly desirable that the part should be “mute”; but “Mr.
Provost” had already made choice of a “very sufficient man” to teach
both the boys to play the viol and to sing, and also to “helpe my
Master Robert’s defect in pronontiation.” Carew was afraid the study
of music, which “elevats the spirits,” might hinder their more serious
lessons; though up to that time the conduct of both boys had been
exemplary. They had said their prayers regularly and been equally
polite to everybody, and were very neat in their “aparelling, kembing,
and washing.” The elder brother had been laid up with “a cowld that
he tooke in the scoole,” which Carew attributed entirely to the fact
that he had outgrown his clothes; and Mr. Perkins, the London tailor,
had been “mighty backward” in sending their new suits. Even with a
bad cold, Frank was his usual pleasant, merry self; and when Mr.
Provost, according to his custom, prescribed “a little phisique,” the
boy drank it cheerfully to the last drop—and “rejected it immediately
after.” Sir Henry Wotton wrote himself to the Earl, describing the
whole episode with an accuracy of detail worthy of Kepler’s
laboratory.
Meantime, Sir Henry himself had assured the Earl that the “spiritay
Robyn’s” voice and pronunciation had been taken in hand by the
Master of the Choristers. Robyn also had caught cold that first winter
at Eton. He had “taken a conceit against his breakfast, being alwaies
curious of his meat, and so going fasting to church.” But on this
occasion, such was the spiritay Robyn’s popularity, that the whole
College seems to have risen in protest against Mr. Provost’s
prescription of “a little phisique.” And Robert recovered without it,
and “continued still increasing in virtues.”
It is somewhat surprising that the younger brother should have
been the favourite, for it was Francis Boyle who had the “quick,
apprehensive wit,” and whose delight was in hunting and
horsemanship; and it was Robyn who dissuaded him, exhorting his
elder brother to learning in his youth, “for,” says he, “there can be
nothing more profitable and honourable.” With his “fayre amiable
countenance,” this child of nine, according to the ebullient Carew,
was “wise, discreet, learned and devout; and not such devotion as is
accustomed in children, but withall in Sincerity he honours God and
prefers Him in all his actions.”[44]
It is very certain that the spiritay Robyn was not fond of games.
There is no enthusiasm for active sports in his Philaretus, not even
of a certain sport that the boys engaged in on winter evenings in the
hall, for which every recent comer was obliged to “find the candles”;
and a very expensive time for candles it was, according to Carew.
But Mr. Robert learnt to “play on music and to sing”, and “to talk Latin
he has very much affected.” And it speaks very well for both Frank
and Robyn that, their tastes being so unlike, they remained such
excellent friends. “Never since they arrived,” according to Carew,
had two ill words passed between them; which he thought was rare
to see, “specially when the younger exceeds the elder in some
qualities.” Some of the noble brothers in the College were continually
quarrelling; but “the peace of God is with my masters.” It had been
noticed even at the Fellows’ table: “Never were sweeter and civiller
gents seen in the Colledge than Mr. Boyles.” The only thing in which
they do not seem to have excelled was in letter-writing. Master Frank
could not write to the Earl because his hand shook; and Master
Robyn could not write because he had hurt his thumb.
And so winter and spring passed, and the summer came, and with
it “breaking-time” at Eton. Mr. Provost, Mr. Harrison, and everybody
else went away. The two boys, and Carew with them, spent their
holidays with their sister Lettice in Sussex. It could not have been a
cheerful visit, though Carew assured the Earl that there was “nothing
wanting to afford a good and pleasant entertainment if my
honourable Lady had not been visited with her continuall guest,
griefe and melancholy.” So extremely melancholy was the Lady
Lettice Goring during this visit that it made the two boys “cry often to
looke upon her.” And yet they must have made a pretty pair to
gladden the eyes of an invalid woman. For Mr. Perkins, the London
tailor, had sent them some fine new clothes—little shirts with laced
bands and cuffs, two scarlet suits without coats, and two cloth-of-
silver doublets.
Robert Boyle’s own recollections of Eton were written a good
many years after he left it. He always remembered with gratitude the
kindness of Mr. John Harrison, in whose house, in that chamber that
was so long in furnishing, the two boys lived—except for some
holidays at “breaking-time,” usually spent in Sussex—from October
1635 to November 1638.
From the very beginning John Harrison must have recognised that
in “Boyle I” he had no ordinary boy to deal with. He saw a “spiritay”
little fellow, with a fair, amiable countenance, a slight stammer, which
the child did his best to amend, and the unstudied civilities of manner
of a little prince. According to Boyle himself, Mr. Harrison saw “some
aptness and much willingness” in him to learn; and this chief
schoolmaster resolved to teach his pupil by “all the gentlest ways of
encouragement.” He began by often dispensing with his attendance
at school in ordinary school hours, and taking the trouble to teach
him “privately and familiarly in his own chamber.”
“He would often, as it were, cloy him with fruit and sweetmeats,
and those little dainties that age is greedy of, that by preventing the
want, he might lessen both his value and desire of them. He would
sometimes give him, unasked, play-days, and oft bestow upon him
such balls and tops and other implements of idleness as he had
taken away from others that had unduly used them. He would
sometimes commend others before him to rouse his emulation, and
oftentimes give him commendations before others to engage his
endeavours to deserve them. Not to be tedious, he was careful to
instruct him in such an affable, kind, and gentle way, that he easily
prevailed with him to consider studying not so much as a duty of
obedience to his superiors, but as a way to purchase for himself a
most delightsome and invaluable good.”[45]
All which means that Mr. Harrison was making a very interesting
experiment, and that his system happened to succeed in the case of
Robert Boyle. The boy learned his “scholar’s task” very easily; and
his spare hours were spent so absorbedly over the books he was
reading that Mr. Harrison was sometimes obliged to “force him out to
play.” And what were the books that were read with such zest? It
was, Robert Boyle says, the accidental perusal of Quintus Curtius
that first made him in love with “other than pedantick books”; and in
after life he used to assert that he owed more to Quintus Curtius
than ever Alexander did: that he had gained more from the history of
Alexander’s conquests than ever Alexander had done from the
conquests themselves.[46]
His other recollections of his Eton schooldays are for the most part
of accidents that happened to him there. He was not so good a
horseman as his brother Frank. Once he fell from his horse, and the
animal trod so near to his throat as to make a hole in his neckband,
“which he long after preserved for a remembrance.” Another time his
nag took fright as he was riding through a town, and reared upright
on his hinder feet against a wall; and the boy just saved himself by
slipping off. Yet a third time he nearly met his death by a “potion”
given him “by an apothecary’s error”; and it is interesting, in the light
of what happened and did not happen in Boyle’s later life, to hear
that “this accident made him long after apprehend more from the
physicians than the disease, and was possibly the occasion that
made him afterwards so inquisitively apply himself to the study of
physick, that he might have the less need of them that profess it.”[47]
The fourth and last of this almost Pauline enumeration of disasters
was the falling, one evening, of the greater part of the wall of the
boys’ bedroom in Mr. Harrison’s house. The two brothers had gone
early to their room; Robyn was already tucked into the big four-post
bed, with its “feather bedd, boulster, and two pillows,” and the
curtains of “blew perpetuana with lace and frenge”,[48] and Frank
was talking with some other boys round the fire when, without a
moment’s warning, the wall of the room fell in, the ceiling with it,
carrying bed, chairs, books and furniture from the room above. A
bigger boy rescued Frank from the debris and dust, the chair in
which he had been sitting broken to pieces, and his clothes torn off
his back; and Robyn, the future chemist, peeping from the blew-
perpetuana curtains, remembered to wrap his head in the sheet, so
that it might serve “as a strainer, through which none but the purer
air could find a passage.”
It is observable that there is no mention of any of those accidents
in the letters to the Earl of Cork from either the Provost or the boys’
personal servant, Carew. Perhaps it was as well that the Earl, much
harassed at home, should not be told everything that was happening
at Eton. As it was, he knew too much. Some go-between—Mr.
Perkins, the tailor, or somebody equally officious—must have told
the Earl in what manner Carew—“poor unmeriting me”, as Carew
called himself in one of his fascinating letters to the Earl—had been
utilising his idle hours by the meandering Thames. Frank and Robyn,
and Carew with them, were spending their holidays with Lady Lettice
Goring, when one morning Sir Henry Wotton, sitting in his study at
Eton, received a letter from the Earl of Cork. The contents came as a
thunderbolt. “Truly, my good Lord,” Sir Henry Wotton wrote back to
the Earl, “I was shaken with such an amazement at the first
percussion thereof, that, till a second perusal, I was doubtfull
whether I had readd aright.” For everybody in the college was so
persuaded of young Mr. Carew’s discretion and temper and zeal in
his charge, and “whole carriadge of himself,” that it would be “harde
to stamp us with any new impression.” However, Mr. Provost had
somewhat reluctantly put away his pipe and “bestowed a Daye in a
little Inquisitiveness.” And he had found that the Earl, in Dublin, was
quite right; that between Carew and a certain “yonge Mayed,
dawghter to our under baker—” and Mr. Provost could not but own
that she was pretty—there had passed certain civil, not to say
amorous, language. The old Provost was evidently disposed to look
leniently on this particular foolish pair. Had he not himself once, in
his youth, written a little poem which began—

O faithless world, and thy most faithless part


A woman’s heart!
...
Why was she born to please, or I to trust
Words writ in dust?[49]

However, Sir Henry told the Earl he was going to talk to Carew on
his return from Sussex, and warn him how careful, in his position, he
ought to be; and he would write again to the Earl after seeing Carew.
But, in the meantime he wished to reserve judgment: “For truely
theare can not be a more tender attendant about youre sweete
children.”
And after all news travelled slowly. Those little love passages were
already six months old: “Tyme enough, I dare swere”—wrote the old
diplomatist, sitting alone in his study, with his Titian and his
Bassanos looking down upon him—“to refrigerat more love than was
ever betweene them.”
CHAPTER IV
THE MANOR OF STALBRIDGE
“... He would very often steal away from all company, and spend four or five
hours alone in the fields, and think at random: making his delighted imagination
the busy scene where some romance or other was daily acted.”—Robert Boyle’s
Philaretus.

After the boys went to Eton, the Earl had very unpleasant things to
think about. Wentworth was pressing him hard. It is true that the little
dinner-parties and card-parties and private theatricals at the Castle
were going on as if there were no Star Chamber behind them. In
January 1636 the Lord Deputy was inviting himself to supper at the
Earl’s Dublin house, and bringing Lady Wentworth with him. Lady
Dungarvan’s baby was born in March, a “ffair daughter”, to be
christened Frances, and to figure in the old Earl’s diary as “lyttle
ffranck”; and the Lord Deputy himself stood sponsor, though he had
just lost his own little son, and the Dungarvan christening had been
postponed till the Wentworth baby had been buried. But the Lord
Deputy’s “sharp pursuit” of men was going on all the same. In
February, before the death of Wentworth’s child and the Dungarvan
christening, Lord Mountmorris had been degraded from the office of
Vice-Treasurer, “tried by a Commission and sentenced to be shot, for
no other crime than a sneer” against Wentworth’s government.[50]
The sentence was not to be carried out; but it became every day
more evident that “whatever man of whatever rank” opposed
Wentworth, or even spoke disrespectfully of his policy, “that man he
pursued to punishment like a sleuth-hound”.[51]
At the beginning of that year, the Earl of Cork had made his “Great
Conveighance,” by which he entailed all his lands upon his five sons.
Wentworth had taken exception to the conveyance of some of these
lands to the Earl’s eldest son, Lord Dungarvan; and in February a
“sharp and large discourse” had taken place between the Lord
Deputy and the Earl. In April the Star Chamber Bill against the Earl,
dealing with his titles to the churchlands of Youghal, was still under
discussion; and Wentworth was now pressing for the payment of
money, by way of ransom, which was at first to be £30,000, but was
afterwards reduced to £15,000.
The Earl was still asserting his right to his lands, and unwilling to
compound—no-one had ever heard the Earl of Cork, he said,
“enclyned to offer anything.” Things were at this pass when at the
end of April Lady Dungarvan, six weeks after her baby’s birth, fell
sick; and the next day, “the smallpockes brake owt uppon her.” On
that very day, under pressure from his friends and from his son
Dungarvan, who went down on his knees before his father, the Earl
of Cork gave way. Very unwillingly, on May 2, he agreed to pay the
£15,000 “for the King’s use,” and for his own “redemption out of
Court”—though his “Innocencie and Intigritie” he declared, writing in
his own private diary, were “as cleer as the son at high noon.” The
old Royalist, even then, believed that if his King only knew how
undeservedly the mighty fine had him imposed, “he would not accept
a penny of it.” The Earl was hard hit, though his great Conveyance
was at last signed and sealed, and he could talk of drinking a cup of
sack “to wash away the care of a big debt.”[52] It is comforting to note
that he had meantime cash in hand not only to tip Archie Armstrong,
the King’s Jester, who seems to have passed through Dublin, but to
pay for two knitted silk waistcoats for his own “somer wearings.”
While all this was going on, Kynalmeaky and Broghill were
enjoying what the Earl called their “peregrination.” A tutor had been
found to accompany them on their foreign travels; a M. Marcombes,
highly recommended to the Earl by Sir Henry Wotton, as a man
“borne for your purpose.” Sir Henry wrote from London, where he
had been spending a week or two, and was returning next day “to
my poore Cell agayne at Eton”;[53] but he gave the Earl a careful
account of Marcombes, whom he had seen in London. He was “by
birthe French; native in the Province of Auvergne; bredd seaven
years in Geneve, verie sounde in Religion, and well conversant with
Religious Men. Furnished with good literature and languages,
espetially with Italian, which he speaketh as promptly as his owne.
And wilbe a good guide for your Sonns in that delicat Piece of the
Worlde. He seemeth of himself neither of a lumpish nor of a light
composition, but of a well-fixed meane.”
M. Marcombes had already won golden opinions in the family of
Lord Middlesex, a former Lord Mayor of London; and was well
known to the then Lord Mayor, Mr. Burlamachy, who also wrote to
the Earl about him. And Mr. Perkins, the tailor, seems to have put in
a word; for there had been a meeting in the “fumie citie” between Sir
Henry Wotton and M. Marcombes and Mr. Perkins, at which Sir
Henry had found the French tutor’s conversation “very apposite and
sweet.”
So in the early spring of 1636 Kynalmeaky and Broghill, with their
governor M. Marcombes, had set out from Dublin on their foreign
travels, stopping long enough in London to kiss the King’s and
Queen’s hands, and obtain the royal licence and passport to travel;
and they took letters also to Sir Henry Wotton at Eton, and to Frank
and Robyn, and poor unmeriting Carew.
The Earl of Cork himself, in the early stages of his struggle with
Wentworth, had thought of going to London, to “justify himself” once
again, as he had done when he was a young man, and Elizabeth
was Queen. But he was no longer a young man, and Charles I was
not Queen Elizabeth, and the Lord Deputy, when he found it out, had
objected strongly to the Earl’s little plan. On the contrary, the Lord
Deputy had gone to England himself, in the summer of 1636; and
though Sir Henry Wotton was under “a kind of hovering conceypt”
that the Earl of Cork was coming over, and there was even a rumour
that he was to be offered the Lord Chancellorship of England, the old
Earl was to remain for two more years in Ireland. He was busy as
usual, moving about, on assize and other duties, between Dublin
and Lismore and Cork; paving the terrace at Lismore with hewn
stones, dedicating the free schools and almshouses there, setting up
an old servant in Dublin in a “tobacko” business, and paying Mr.
Perkins’s bill for those little scarlet suits and cloth-of-silver doublets
that Frank and Robyn were wearing in their Whitsuntide holidays. Sir
Henry Wotton was able to tell the Earl that Lady Lettice would see
Frank in better health and strength than he had been in either
kingdom before, while Robert would “entertayne her with his pretie
conceptions, now a greate deale more smoothely than he was
wonte.”
The Earl had not given up his English project; on the contrary, it
was to mature into the purchase of a little bit of England for his very
own; and his choice had fallen on a “capitall howse, demesne, and
lands” in Dorsetshire. Accordingly in the autumn of 1636 he bought
the Manor of Stalbridge, and sent over a steward, Thomas Cross, to
take possession. At Stalbridge the Earl would be a near neighbour of
the Earl of Bristol—his son-in-law Digby’s uncle—at Sherborne
Castle.
The year 1636 had been a trying year; and one of the first
expenses in the New Year 1637 was a fee to Mr. Jacob Longe, of
Kinsale, “my Jerman physician,” for plaisters and prescriptions, “to
stay the encrease of the dead palsy which hath seized uppon all the
right side of my boddy (God helpe me) £5.” And though the returns
for the year shewed a “Lardge Revenew,” and the diary record for
the year ended in a note of triumph, with a triple “Amen, Amen,
Amen,” there was yet sorrow in store that no revenue, however
large, could avert. For Peggie, the Earl’s youngest daughter, was ill.
The Earl had paid £5 to Mr. Higgins, the Lismore doctor, to give her
“phisick, which he never did”; and either because of this, or in spite
of this, Little Peggie did not get well. She died in June 1637, in Lady
Clayton’s house in Cork, where she and Mary had lived all this time
together. The Lady Margaret Boyle, youngest daughter of the Earl of
Cork—eight years old when she died—was buried in the family tomb
at Youghal.
It was not till Midsummer 1638, when the last instalment of the
mighty fine had been paid, that the Earl began his preparations for a
prolonged visit to England. He revoked all other wills, and again
made a last will and testament; and at the end of July he actually set
out for England, taking with him his daughter Mary, Lord and Lady
Barrymore, and several of the grandchildren.
The parting was a sad one between Mary Boyle and Lady Clayton,
who had just lost her husband, and, a childless woman herself, had
been a real mother to “Moll” and “Peggie.” But the Earl had a grand
marriage in view for his daughter Mary; and he had yet to discover
that Lady Mary had a will of her own: that of all his daughters it was
she who had inherited his own indomitable pride. Hitherto, she had
been a child, brought up away from him; to be gladdened from time
to time by a happy visit or a New Year’s gift. But even these are
indications of the little lady’s tastes and character. It was to Mary the
Earl gave the “ffether of diamonds and rubies that was my wive’s,”
long before he could have known how defiantly she would toss that
little head of hers. She must have been a fair horsewoman already at
nine years old; for it was to her that the Earl sent the dead mother’s
saddle and saddle-cloth of green velvet, laced and fringed with silver
and green silk; and it is certain she inherited the Earl’s love of fine
dressing, from the choice of various small gowns of figured satins
and rich stuffs of scarlet dye. Of even more significance is the old
Earl’s gift of Sir Philip Sidney’s Arcadia, “To my daughter, Mary
Boyle,” when this imperious young creature was only twelve years
old. Do little girls of twelve read Sir Philip Sidney’s Arcadia to-day?
There was to come a moment when, if the Earl had ever read it
himself, he must have heard in “Moll’s” voice, as she answered him,
some echo of Sidney’s teaching—

“... but a soule hath his life


Which is held in loue: loue it is hath ioynd
Life to this our Soule.”

After the usual delays at starting, the Ninth Whelp made a good
passage; and the Earl and his party reached Bristol safely on
Saturday August 4. As usual, presents were dispensed to the ship’s
captain and company, together with what remained of a hogshead of
claret wine. Next day, Sunday, the whole family went obediently to
church; and on Monday morning, leaving the others to follow with the
servants and luggage, the old Earl, riding a borrowed horse, set off
by himself to find his way to Stalbridge.[54]
A wonderful peace and stillness falls on the Dorsetshire uplands at
evening after a long, hot summer day. Up hill and down dale and up
hill again go the Dorsetshire lanes, between their tangled hedges,
through a country of undulating woods and downs and soft green
pastures. The lark sings, high up, invisible: a far-away, sleepy cock-
crow or faint bark of sheepdog breaks the silence; the grazing cattle
bend their brown heads in the fields.
The Earl was in England again, the land of his birth. It was
perhaps not altogether a prosperous and satisfied England, in
August 1638. The heavy hand of taxation was on even these
pastoral uplands. The heart of England was throbbing with political
unrest. But on that evening, at least, there could have been only the
lark’s ecstasy, and the sweet smell of wild thyme and woodsmoke in
the air. Ireland, the distressful country of his adoption, lay behind the
old man, and with it the memories of fifty strenuous years;—all that
was hardest and proudest and tenderest in a lifetime.
Lord and Lady Dungarvan were already at Stalbridge with “lyttle
ffrancke.” There was another baby-daughter now, but it had
apparently been left at Salisbury House, in London. Dungarvan had
ridden some six miles upon the road to meet his father. It was still
daylight when, riding together—the old man must have been pretty
stiff in the saddle, for he had ridden nearly sixty miles that day—they
came in sight of the Elizabethan manor standing among elms and
chestnut trees, surrounded by park lands and hayfields and
orchards: “My owne house of Stalbridge in Dorcetshier; this being
the firste tyme that ever I sawe the place.”[55]
After this, the movements of the Earl and his family read rather like
a Court Circular. Not much is heard of the life that must have been
going on in the little town itself, with its Church and market Cross;
but the mere presence of this great Irish family among them must, by
the laws of supply and demand, have wrought many changes in the
little market town. The Earl paid his love and service to his neighbour
and kinsman, the Earl of Bristol, at Sherborne Castle, and the Earl
and Countess of Bristol, with all their house-party, immediately
returned the visit; after which the whole family at the Manor were
“feasted” for two days at Sherborne Castle. The Earl of Cork and his
house-party rode to “the Bathe”, and return visits were received at
Stalbridge from friends at “the Bathe”. And a week or two later the
Earl, attended by Dungarvan and Barrymore, rode to London, and
was graciously received by the King and “all the Lords at Whitehall.”
The King praised the Earl’s government of Ireland, and the
Archbishop of Canterbury was particularly friendly.
Lady Barrymore and Lady Dungarvan had between them
undertaken to ease the Earl from the “trowble of hows-keeping,” and
for this purpose were allowed £50 a week, and more when they
wanted it; and the cellars and larders at Stalbridge were replenished
from time to time with gifts. A ton of claret wine and six gallons of
aqua vitæ arrived as a New Year’s gift from Munster, and “veary fatt
does” from English friends; while among his assets the Earl counted,
besides the produce of his Stalbridge lands and woods, the twenty
stalled oxen, the powdered beef, the bacon and salted salmon that
were sent from his Irish estates.
Thomas Cross, his steward, became “seneschal”;—perhaps there
was a seneschal at Sherborne Castle; and there was a Clerk of the
Kitchen and a large staff of household servants, men and women,
and a long list of rules for the management of the household drawn
up and signed by the Earl himself.[56] And of course the “re-
edification” of the Manor House began at once. There was water to
be carried in leaden pipes; new furniture to hasten home from the
London upholsterer, who dwelt at the sign of the Grasshopper; a red
embroidered bed, a tawny velvet carpet, couch and chairs. There
was a new coach to buy, and the paths and terraces at Stalbridge
were to be stone paved exactly like the paths and terraces at
Sherborne Castle. Stairs with a stone balustrade, and carved stone
chimney-pieces were to be added to the Manor;—one at least
carved with the Earl’s coat of arms “compleate,” and reaching nearly
to the ceiling, “fair and graceful in all respects.” There was a limekiln
to build, and pit coal to procure and cane apples[57] to be planted in
the orchard. But charity only began at home; and in this case it did
not prevent a subscription being sent—“a myte” of £100—to help the
Archbishop of Canterbury in his scheme of “re-edifying Pawle’s
Church in London.”
Meantime, all the Earl’s daughters and sons-in-law, except
Dorothy and Arthur Loftus, who remained in Ireland, seem to have
found their way, separately or together, to the Manor of Stalbridge;
while grandchildren, nieces and nephews and even “cozens” were
welcomed under its roof. The Dungarvans made their headquarters
there, and the Barrymores, and the little Lady Mary, who was now
fourteen, and to be considered a grown-up young lady, with an
allowance of £100 a year “to fynde herself.” And they were presently
joined by the Kildares, and Katharine and Arthur Jones. Even the
plaintive Lettice and her lord stayed for some time under the Earl’s
roof.[58] And in March 1639, after an absence of three years,
Kynalmeaky and Broghill, with M. Marcombes, returned from their
“peregrination”. They found Frank and Robyn already at Stalbridge,
though not in the great house itself. For their father had taken the
boys away from Eton on his return journey from London in
November 1638, and since then they had been boarded out with the
Rev. Mr. Dowch at the Parsonage, scarcely “above twice a musket-
shot” distant from their father’s house. Their three years at Eton had
cost, “for diett, tutaradge and aparell,” exactly £914 3s. 9d.
When the Earl of Cork visited Eton and took his two boys away, Sir
Henry Wotton must have been already ill. Since his return after the
summer breaking-time of 1638 the old Provost had suffered from a
feverish distemper, which was to prove the beginning of the end.[59]
It is possible that during the Provost’s illness extra duties had
fallen on Mr. Harrison, the Rector; in any case the two boys had
been removed from the care of their “old courteous schoolmaster,”
and handed over to “a new, rigid fellow;” and things were not going
quite so happily for them at Eton as heretofore. Moreover, poor
Carew, the romance of the underbaker’s daughter nipped in the bud,
had, from overmuch fondness for cards and dice, come utterly to
grief.
It was during this last year of Robert Boyle’s schooldays—in the
April of 1638, before Sir Henry Wotton’s illness, and while all was
going on as usual at Eton—that Mr. Provost had entertained at his
hospitable table a guest whose life was to be strangely linked in after
years with that of some members of the great Boyle family. This was
John Milton, then a young poet, living with his father at Horton—not
far from Eton—and just about to set out on his Italian journey.[60]
Was Robert Boyle one of the “hopeful youths” selected by the
Provost to dine at his table that day when Milton dined there? And
did Robert Boyle listen to the talk that went on at table between
Milton and his friend, the learned Hales, and Sir Henry Wotton? It
was very pleasant talk. When Milton returned to Horton he ventured
to send the Provost a little letter of thanks and a copy of his Comus
as a parting gift; and Sir Henry sent his own footboy post-haste to
Horton, to catch Milton before he started, with a pretty letter of
acknowledgment and an introduction to the British Agent in Venice. It
is noteworthy that the advice Sir Henry Wotton gave to Milton, and
the advice he always gave to his own pupils when they were setting
out on a career of diplomacy abroad, showed that, while the old man
had not forgotten his experience of the Augsburg album, his kindly
cynicism remained unchanged. I pensiori stretti, was the advice he
handed on in his charming letter to Milton,—ed il viso sciolto; while to
all young Etonians travelling in diplomacy he used to say, Always tell
the truth; for you will never be believed.
It is hard to say how much Robert Boyle may have owed to the
guidance and talk of Sir Henry Wotton. Boyle remembered him as a
fine gentleman who possessed the art of making others so; and it
was John Harrison’s methods of teaching that had impressed the
boy. Yet it must not be forgotten that the Provost’s tastes were not
only literary and scholarly; that he had not only surrounded himself
with a library of books that Robert Boyle in his boyhood must have
envied—Sir Henry Wotton was of a scientific turn of mind: he was
fond of experimenting. Ever since the days when he had watched
Kepler at work in his laboratory and supplied his cousin Lord Bacon
with facts, he had been accustomed to occupy himself, in more or
less dilletante fashion, with such little experiments as the distilling of
medicinal herbs and the measurement of time by allowing water to
pass through a filter, drop by drop; and it was Sir Henry Wotton
whom Izaak Walton consulted about the preparation of “seductive-
smelling oils” in the catching of little fishes. And who could it have
been, in that last year that Robyn spent at Eton, who lent him the
books that “meeting in him with a restless fancy” gave his thoughts
such a “latitude of roving”? Robyn had been away from school on a
visit to London, and there had fallen ill of a “tertian ague”, and had
been sent back to Eton to see if good air and diet might not do more
for him than all “the Queen’s and other doctors’ remedies” had done.
His own phrase[61] is that “to divert his melancholy they made him
read the State Adventures of Amadis de Gaule, and other fabulous
and wandering stories.” Who was the “they” at Eton? It could not
have been the “new, rigid fellow”. Amadis de Gaule may have been
part of Mr. John Harrison’s system of education, but one would like
to believe that Sir Henry Wotton had some hand in fashioning Robert
Boyle—that his whole library was open to the boy, not only the books
of romance and adventure in it that gave Robyn’s thoughts such a
“latitude of roving.” One would like to believe that the torch was
indeed passed on from Kepler’s laboratory, and by the study of one
of those three copies of Bacon’s Novum Organum, into the hands of
England’s first great experimental chemist.
Be that as it may, Sir Henry Wotton was already ill when Frank and
Robyn were removed from Eton in November 1638; and it was Mr.
Harrison who duly sent after them to Stalbridge the furniture of their
chamber—the blew perpetuana curtains and all the other things so
carefully inventoried by poor Carew. And Carew himself no longer
served his sweet young masters: he had been succeeded by a
manservant with the suggestive name of Rydowt, who appears to
have been a married man, and was accommodated with a little
cottage of his own at Stalbridge, with a garden which the Earl
planted with “cane apples” from Ireland. The boys were to live and
learn their lessons with Mr. Dowch at the Parsonage; and that “old
divine” was very soon to discover that Robyn had not learnt much
Latin at Eton after all, and “with great care and civility” to proceed to
read with him the Latin poets as well as the Latin prose-writers. And
while the Earl gave Frank a horse of his own, and knew him to be
happy and gallant in the saddle, it was Robyn who was the old Earl’s
Benjamin, most loved of all his sons. The family saw a likeness in
Robyn to his father—a likeness both in body and mind. It is difficult
to credit the Earl of Cork with any of his youngest son’s habit of
“unemployed pensiviness,” but there must have been something in
Robyn when he was quite a little fellow—a quiet self-reliance—that
impressed the old Earl strangely. Robyn was only twelve years old;
he had as yet shown none of the traits of character that the Earl so
“severely disrelished” in some of his sons and sons-in-law. And so,
when he gave Frank the horse, the Earl listened perhaps with some
wonderment to Robyn’s “pretie conceptions” in excellent language,
spoken still not quite smoothly; and he was content to let the boy
wander as he liked. He gave his Benjamin the keys of his orchard,
not afraid to leave him in a very paradise of unplucked apples,
“thinking at random.”
CHAPTER V
THE HOUSE OF THE SAVOY
“You shall have all this winter att the Savoy, in Sʳ. Tho. Stafford’s howse, the
greatest familie that will be in London (I pray God the ould man houlds out).”—
Letter from Sir John Leeke to Sir Edmund Verney, 1639: Verney Memoirs, vol. i.

The Earl’s gift of a horse to his son Frank had been made at a
psychological moment. For Frank, at sixteen, was not over robust,
and Robyn, not much more than twelve, was scarcely fitted to
defend his King and Country; and they must have just watched their
three elder brothers, Dungarvan, Kynalmeaky and Broghill, ride off in
great spirits from Stalbridge to join the King at Newcastle or York.
War was in the air, and rumours of war; and Stalbridge Manor and
Sherborne Castle, and the villages of Dorsetshire and all the great
families in England, were astir, as day by day and week by week the
troops of English horse and foot were moving northwards to engage
against what the old Earl turned off so lightly in his diary as the
“Covenanting, rebelleows Skots.”
In January 1639 a circular letter had been sent out in King
Charles’ name to all the English nobility, asking them to state how far
and in what manner they were ready to assist the Scottish
Expedition; and week by week, according to their means and their
political inclinations, the English nobles—Laudians and Puritans
alike—had been sending in their answers: £1000, and twenty horse;
twenty horse and attendance in person or by substitute; £1000 in lieu
of horse; £500 and twelve horse; and so on.[62] And some gave
willingly, and some gave grudgingly, and some evaded promising to
give anything; and one or two were brave enough to refuse—and to
give their reasons why. Even at Court there was “much contrarity”; it
was a case of “soe many men soe many opinions.”[63] Wentworth,
over in Ireland, had written offering his King £2000, and if necessary

You might also like